Skip to main content

Advertisement

Log in

Multiscale two-way embedding schemes for free-surface primitive equations in the “Multidisciplinary Simulation, Estimation and Assimilation System”

  • Published:
Ocean Dynamics Aims and scope Submit manuscript

Abstract

We derive conservative time-dependent structured discretizations and two-way embedded (nested) schemes for multiscale ocean dynamics governed by primitive equations (PEs) with a nonlinear free surface. Our multiscale goal is to resolve tidal-to-mesoscale processes and interactions over large multiresolution telescoping domains with complex geometries including shallow seas with strong tides, steep shelfbreaks, and deep ocean interactions. We first provide an implicit time-stepping algorithm for the nonlinear free-surface PEs and then derive a consistent time-dependent spatial discretization with a generalized vertical grid. This leads to a novel time-dependent finite volume formulation for structured grids on spherical or Cartesian coordinates, second order in time and space, which preserves mass and tracers in the presence of a time-varying free surface. We then introduce the concept of two-way nesting, implicit in space and time, which exchanges all of the updated fields values across grids, as soon as they become available. A class of such powerful nesting schemes applicable to telescoping grids of PE models with a nonlinear free surface is derived. The schemes mainly differ in the fine-to-coarse scale transfers and in the interpolations and numerical filtering, specifically for the barotropic velocity and surface pressure components of the two-way exchanges. Our scheme comparisons show that for nesting with free surfaces, the most accurate scheme has the strongest implicit couplings among grids. We complete a theoretical truncation error analysis to confirm and mathematically explain findings. Results of our discretizations and two-way nesting are presented in realistic multiscale simulations with data assimilation for the middle Atlantic Bight shelfbreak region off the east coast of the USA, the Philippine archipelago, and the Taiwan–Kuroshio region. Multiscale modeling with two-way nesting enables an easy use of different sub-gridscale parameterizations in each nested domain. The new developments drastically enhance the predictive capability and robustness of our predictions, both qualitatively and quantitatively. Without them, our multiscale multiprocess simulations either were not possible or did not match ocean data.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17

Similar content being viewed by others

References

  • Agarwal A (2009) Statistical field estimation and scale estimation for complex coastal regions and archipelagos. Master’s thesis, Massachusetts Institute of Technology, Cambridge, MA

  • Agarwal A, Lermusiaux PFJ (2010) Statistical field estimation for complex coastal regions and archipelagos. Ocean Model. Sub Judice

  • Arakawa A, Lamb VR (1977) Computational design of the basic dynamical processes of the UCLA general circulation model. Methods Comput Phys 17:173–265

    Google Scholar 

  • Arango HG, Ezer T, Shchepetkin AF (2010) Terrain-following ocean modeling system (TOMS). http://www.myroms.org/index.php?model=toms

  • Barth A, Alvera-Azcárate A, Rixen M, Beckers JM (2005) Two-way nested model of mesoscale circulation features in the Ligurian Sea. Prog Oceanogr 66(2–4):171–189

    Article  Google Scholar 

  • Barth A, Alvera-Azcárate A, Beckers JM, Rixen M, Vandenbulcke L (2007) Multigrid state vector for data assimilation in a two-way nested model of the Ligurian Sea. J Mar Syst 65(1–4):41–59

    Article  Google Scholar 

  • Beckers JM (1991) Application of the GHER 3D general circulation model to the western Mediterranean. J Mar Syst 1(4):315–332

    Article  Google Scholar 

  • Beckers JM, Brasseur P, Nihoul JCJ (1997) Circulation of the western Mediterranean: from global to regional scales. Deep-Sea Res II 44(3–4):531–549

    Article  Google Scholar 

  • Berntsen J, Svendsen E (1999) Using the SKAGEX dataset for evaluation of ocean model skills. J Mar Syst 18(4):313–331

    Article  Google Scholar 

  • Blayo E, Debreu L (2005) Revisiting open boundary conditions from the point of view of characteristic variables. Ocean Model 9(3):231–252

    Article  Google Scholar 

  • Bleck R (2002) An oceanic general circulation model framed in hybrid isopycnic-Cartesian coordinates. Ocean Model 4(1):55–88

    Article  Google Scholar 

  • Blumberg AF, Mellor GL (1987) A description of a three-dimensional coastal ocean circulation model. In: Heaps NS (ed) Three-dimensional coastal ocean models. American Geophysical Union, Washington, DC, pp 1–16

    Google Scholar 

  • Bryan K (1969) A numerical method for the study of the circulation of the world ocean. J Comput Phys 4(3):347–376

    Article  Google Scholar 

  • Bunya S, Dietrich JC, Westerink JJ, Ebersole BA, Smith JM, Atkinson JH, Jensen R, Resio DT, Luettich RA, Dawson C, Cardone VJ, Cox AT, Powell MD, Westerink HJ, Roberts HJ (2010) A high-resolution coupled riverine flow, tide, wind, wind wave, and storm surge model for southern Louisiana and Mississippi. Part I: model development and validation. Mon Weather Rev 138(2):345–377

    Google Scholar 

  • Cailleau S, Fedorenko V, Barnier B, Blayo E, Debreu L (2008) Comparison of different numerical methods used to handle the open boundary of a regional ocean circulation model of the Bay of Biscay. Ocean Model 25(1–2):1–16

    Article  Google Scholar 

  • Carnes MR, Fox DN, Rhodes RC, Smedstad OM (1996) Data assimilation in a north Pacific Ocean monitoring and prediction system. In: Malanotte-Rizzoli P (ed) Modern approaches to data assimilation in ocean modeling. Elsevier, Amsterdam, pp 319–345

    Chapter  Google Scholar 

  • Carter EF, Robinson AR (1987) Analysis models for the estimation of oceanic fields. J Atmos Ocean Technol 4(1):49–74

    Article  Google Scholar 

  • Chao Y, Li Z, Farrara J, McWilliams JC, Bellingham J, Capet X, Chavez F, Choi JK, Davis R, Doyle J, Fratantoni DM, Li P, Marchesiello P, Moline MA, Paduan J, Ramp S (2009) Development, implementation and evaluation of a data-assimilative ocean forecasting system off the central California coast. Deep-Sea Res Part 2 Top Stud Oceanogr 56(3–5):100–126

    Article  Google Scholar 

  • Chapman NR, Lynch JF (2010) Special issue on the 2006 shallow water experiment. IEEE J Oceanic Eng 35(1):1–2

    Article  Google Scholar 

  • Chassignet EP, Malanotte-Rizzoli P (eds) (2000) Ocean circulation model evaluation experiments for the North Atlantic basin. Dyn Atmos Ocean 32(3–4)

    Google Scholar 

  • Chassignet EP, Arango H, Dietrich D, Ezer T, Ghil M, Haidvogel DB, Ma CC, Mehra A, Paiva AM, Sirkes Z (2000) DAMEE-NAB: the base experiments. Dyn Atmos Ocean 32(3–4):155–183

    Article  Google Scholar 

  • Chassignet EP, Smith L, Halliwell G, Bleck R (2003) North Atlantic simulation with the hybrid coordinate ocean model (HYCOM): impact of the vertical coordinate choice, reference density, and thermobaricity. J Phys Oceanogr 33(12):2504–2526

    Article  Google Scholar 

  • Chassignet EP, Wallcraft A, Halliwell G, Bleck R, McDonald M, Ryan E, Srinivasan A (2009) HYCOM consortium for data assimilative modeling. http://www.hycom.org/

  • Courant R, Friedrichs K, Lewy H (1928) Über die partiellen differenzengleichungen der mathematischen physik. Math Ann 100(1):32–74

    Article  Google Scholar 

  • Cushman-Roisin B, Beckers JM (2010) Introduction to geophysical fluid dynamics: physical and numerical aspects. Academic, New York

    Google Scholar 

  • Debreu L, Blayo E (2008) Two-way embedding algorithms: a review. Ocean Dyn 58(5–6):415–428

    Article  Google Scholar 

  • Deleersnijder E, Campin JM (1995) On the computation of the barotropic mode of a free-surface world ocean model. Ann Geophys 13(6):675–688

    Article  Google Scholar 

  • Deleersnijder E, Lermusiaux PFJ (2008a) Multi-scale modeling: nested grid and unstructured mesh approaches. Ocean Dyn 58(5–6):335–336. doi:10.1007/s10236-008-0170-5

    Article  Google Scholar 

  • Deleersnijder E, Lermusiaux PFJ (eds) (2008b) Multi-scale modeling: nested grid and unstructured mesh approaches. Ocean dynamics, vol 58. Springer, Berlin, pp 335–336

    Google Scholar 

  • Dukowicz JK, Smith RD (1994) Implicit free-surface method for the Bryan–Cox–Semtner ocean model. J Geophys Res 99(C4):7991–8014

    Article  Google Scholar 

  • Dutay JC, Bullister JL, Doney SC, Orr JC, Najjar R, Caldeira K, Campin JM, Drange H, Follows M, Gao Y (2002) Evaluation of ocean model ventilation with CFC-11: comparison of 13 global ocean models. Ocean Model 4(2):89–120

    Article  Google Scholar 

  • Egbert GD, Erofeeva SY (2002) Efficient inverse modeling of barotropic ocean tides. J Atmos Ocean Technol 19(2):183–204

    Article  Google Scholar 

  • Estournel C, Auclair F, Lux M, Nguyen C, Marsaleix P (2009) Scale oriented embedded modeling in the north-western Mediterranean in the frame of MFSTEP. Ocean Sci 5(2):73–90

    Article  Google Scholar 

  • Ezer T, Arango H, Shchepetkin AF (2002) Developments in terrain-following ocean models: intercomparisons of numerical aspects. Ocean Model 4(3–4):249–267

    Article  Google Scholar 

  • Ferziger JH, Perić M (1996) Computational methods for fluid dynamics. Springer, Berlin

    Google Scholar 

  • Fox AD, Maskell SJ (1995) Two-way interactive nesting of primitive equation ocean models with topography. J Phys Oceanogr 25(12):2977–2996

    Article  Google Scholar 

  • Gangopadhyay A, Robinson AR, Haley PJ, Leslie WG, Lozano CJ, Bisagni JJ, Yu Z (2003) Feature-oriented regional modeling and simulations in the Gulf of Maine and Georges Bank. Cont Shelf Res 23(3–4):317–353

    Article  Google Scholar 

  • Gawarkiewicz G (2008) Quantifying, predicting, exploiting uncertainty (QPE), 2008 pilot experiment; Aug. 22nd–Sep. 11th, 2008. Tech. rep., Woods Hole Oceanographic Institute, Woods Hole

    Google Scholar 

  • Gent PR, Bryan FO, Danabasoglu G, Doney SC, Holland WR, Large WG, McWilliams JC (1998) The NCAR climate system model global ocean component. J Climate 11(6):1287–1306

    Article  Google Scholar 

  • Ginis I, Richardson RA, Rothstein LM (1998) Design of a multiply nested primitive equation ocean model. Mon Weather Rev 126(4):1054–1079

    Article  Google Scholar 

  • Gordon AL (2009) Regional cruise intensive observational period 2009. Tech. rep., Lamont Doherty Earth Observatory. http://mseas.mit.edu/Sea_exercises/Philex_IOP09/RIOP09_LEG2_rept.pdf

  • Griffies SM, Böning C, Bryan FO, Chassignet EP, Gerdes R, Hasumi H, Hirst A, Treguier AM, Webb D (2000) Developments in ocean climate modelling. Ocean Model 2(3–4):123–192

    Article  Google Scholar 

  • Griffies SM, Harrison J, Pacanowski RC, Rosati T (2007) Ocean modelling with MOM. CLIVAR Exchanges 12(3):3–5

    Google Scholar 

  • Griffies SM, Schmidt M, Herzfeld M (2010) Elements of MOM4p1. GFDL Ocean Group Technical Report 6. Geophysical Fluid Dynamics Laboratory, Princeton. http://www.gfdl.noaa.gov/cms-filesystem-action/model_development/ocean/mom-guide4p1.pdf

  • Guo X, Hukuda H, Miyazawa Y, Yamagata T (2003) A triply nested ocean model for simulating the Kuroshio—roles of horizontal resolution of JEBAR. J Phys Oceanogr 33(1):146–169

    Article  Google Scholar 

  • Haidvogel DB, Arango HG, Hedstrom K, Beckmann A, Malanotte-Rizzoli P, Shchepetkin AF (2000) Model evaluation experiments in the North Atlantic Basin: simulations in nonlinear terrain-following coordinates. Dyn Atmos Ocean 32(3–4):239–281

    Article  Google Scholar 

  • Haley PJ Jr (1999) GRIDS user’s guide. Harvard Open Ocean Model Reports 54. Harvard University, Cambridge. http://mseas.mit.edu/archive/HOPS/Grids/grids.ps.gz

  • Haley PJ Jr, Lozano CJ (2001) COND_TOPO: topography conditioning in MATLAB. http://mseas.mit.edu/archive/HOPS/Cond_Topo/cond_topo.ps.gz

  • Haley PJ Jr, Lermusiaux PFJ, Leslie WG, Robinson AR (1999) Harvard ocean prediction system (HOPS). http://oceans.deas.harvard.edu/HOPS/

  • Haley PJ Jr, Lermusiaux PFJ, Robinson AR, Leslie WG, Logoutov O, Cossarini G, Liang XS, Moreno P, Ramp SR, Doyle JD, Bellingham J, Chavez F, Johnston S (2009) Forecasting and reanalysis in the Monterey Bay/California current region for the autonomous ocean sampling network-II experiment. Deep-Sea Res II 56(3–5):127–148. doi:10.1016/j.dsr2.2008.08.010

    Article  Google Scholar 

  • Haney RL (1991) On the pressure gradient force over steep topography in sigma coordinate ocean models. J Phys Oceanogr 21(4):610–619

    Article  Google Scholar 

  • Kaufman M (2010) Upwelling dynamics off Monterey Bay: heat flux and temperature variability, and their sensitivities. Senior thesis

  • KleptsovaO, StellingGS, Pietrzak JD (2010) On conservative advection schemes in an unstructured grid finite volume threedimensional hydrostatic model. Ocean Dyn (in press, this special issue)

  • Lam FP, Haley PJ Jr, Janmaat J, Lermusiaux PFJ, Leslie WG, Schouten MW, te Raa L, Rixen M (2009) At-sea real-time coupled four-dimensional oceanographic and acoustic forecasts during battlespace preparation 2007. J Mar Syst 78:306–320. doi:10.1016/j.jmarsys.2009.01.029

    Article  Google Scholar 

  • le Roux DY, Sène A, Rostand V, Hanert E (2005) On some spurious mode issues in shallow-water models using a linear algebra approach. Ocean Model 10(1–2):83–94

    Article  Google Scholar 

  • Leben RR, Born GH, Engebreth BR (2002) Operational altimeter data processing for mesoscale monitoring. Marine Geodesy 25(1–2):3–18

    Article  Google Scholar 

  • Lermusiaux PFJ (1997) Error subspace data assimilation methods for ocean field estimation: theory, validation and applications. Ph.D. thesis, Harvard University, Cambridge

  • Lermusiaux PFJ (1999) Data assimilation via error subspace statistical estimation. Part II: middle Atlantic Bight shelfbreak front simulations and ESSE validation. Mont Weather Rev 127(8):1408–1432

    Article  Google Scholar 

  • Lermusiaux PFJ (2002) On the mapping of multivariate geophysical fields: sensitivity to size, scales and dynamics. J Atmos Ocean Technol 19(10):1602–1637

    Article  Google Scholar 

  • Lermusiaux PFJ (2006) Uncertainty estimation and prediction for interdisciplinary ocean dynamics. J Comput Phys 217(1):176–199. (special issue of on “Uncertainty quantification”. J Glimm, G Karniadakis, eds)

    Article  Google Scholar 

  • Lermusiaux PFJ (2007) Adaptive sampling, adaptive data assimilation and adaptive modeling. Physica D 230:172–196. (special issue on “Mathematical issues and challenges in data assimilation for geophysical systems: interdisciplinary perspectives”. Christopher KRT Jones and Kayo Ide, eds)

    Article  Google Scholar 

  • Lermusiaux PFJ, Xu J (2010) Coupled ocean-acoustic prediction of transmission loss in a continental shelfbreak region: predictive skill, uncertainty quantification and dynamical sensitivities. IEEE Trans J Ocean Eng. doi:10.1109/JOE.2010.2068611

    Google Scholar 

  • Lermusiaux PFJ, Anderson DGM, Lozano CJ (2000) On the mapping of multivariate geophysical fields: error and variability subspace estimates. Q J R Meteorol Soc 126(565):1387–1429

    Article  Google Scholar 

  • Lermusiaux PFJ, Robinson AR, Haley PJ, Leslie WG (2002) Filtering and smoothing via error subspace statistical estimation. In: Advanced interdisciplinary data assimilation. The OCEANS 2002 MTS/IEEE. Holland Publications, pp 795–802

  • Lermusiaux PFJ, Haley PJ Jr, Leslie WG, Logoutov O, Robinson AR (2006) Autonomous wide aperture cluster for surveillance (AWACS): adaptive sampling and search using predictive models with coupled data assimilation and feedback—Harvard page. http://mseas.mit.edu/archive/AWACS/index_AWACS.html

  • Lermusiaux PFJ, Haley PJ Jr, Leslie WG, Logutov O (2009a) Philippines Strait Dynamics Experiment—MSEAS Home Page. http://mseas.mit.edu/Research/Straits/

  • Lermusiaux PFJ, Haley PJ Jr, Leslie WG, Logutov O, Agarwal A, Burton L (2009b) MSEAS PHILEX IOP09 Real-Time Home Page. http://mseas/Sea_exercises/Philex_IOP09/

  • Lermusiaux PFJ, Haley PJ Jr, Leslie WG, Logutov O, Xu J, Agarwal A (2009c) MSEAS QPE IOP 2009 Real-Time Results. http://mseas/Sea_exercises/QPE_IOP09/index_iop2009.html

  • Leslie WG, Robinson AR, Haley P, Logoutov O, Moreno P, Lermusiaux PFJ, Coehlo E (2008) Verification and training of real-time forecasting of multi-scale ocean dynamics for maritime rapid environmental assessment. J Mar Syst 69(1–2):3–16

    Article  Google Scholar 

  • Liang WD, Tang TY, Yang YJ, Ko MT, Chuang WS (2003) Upper-ocean currents around Taiwan. Deep-Sea Res Part 2 50(6–7):1085–1105

    Article  Google Scholar 

  • Lin Y-T, Newhall AE, Duda TF, Lermusiaux PFJ, Haley PJ (2010) Merging multiple-partial-depth data time series using objective empirical orthogonal function fitting. IEEE Trans J Ocean Eng. doi:10.1109/JOE.2010.2052875

    Google Scholar 

  • Locarnini RA, Mishonov AV, Antonov JI, Boyer TP, Garcia HE (2006) NOAA NODC WOA05. http://iridl.ldeo.columbia.edu/SOURCES/.NOAA/.NODC/.WOA05/

  • Logutov OG (2007) Multi-model fusion and uncertainty estimation for ocean prediction. Ph.D. thesis, Harvard University, Cambridge

  • Logutov OG (2008) A multigrid methodology for assimilation of measurements into regional tidal models. Ocean Dyn 58(5–6):441–460

    Article  Google Scholar 

  • Logutov OG, Lermusiaux PFJ (2008) Inverse barotropic tidal estimation for regional ocean applications. Ocean Model 25(1–2):17–34

    Article  Google Scholar 

  • Lozano CJ, Haley PJ, Arango HG, Sloan NQ, Robinson AR (1994) Harvard coastal/deep water primitive equation model. Harvard Open Ocean Model Reports 52. Harvard University, Cambridge

  • Lynch DR, Davies AM (eds) (1995) Quantitative skill assessment for coastal ocean models, coastal and estuarine studies, vol 47. American Geophysical Union, Washington, DC

    Google Scholar 

  • Lynch DR, Naimie CE, Ip JT, Lewis CV, Werner FE, Luettich R, Blanton BO, Quinlan J, McGillicuddy DJ Jr, Ledwell JR, Churchill J, Kosnyrev V, Davis CS, Gallager SM, Ashjian CJ, Lough RG, Manning J, Flagg CN, Hannah CG, Groman RC (2001) Real-time data assimilative modeling on Georges Bank. Oceanography 14(1):65–77

    Google Scholar 

  • Maderich V, Heling R, Bezhenar R, Brovchenko I, Jenner H, Koshebutskyy V, Kuschan A, Terletska K (2008) Development and application of 3D numerical model THREETOX to the prediction of cooling water transport and mixing in the inland and coastal waters. Hydrol Process 22(7):1000–1013

    Article  Google Scholar 

  • Marshall J, Hill C, Perelman L, Adcroft A (1997) Hydrostatic, quasi-hydrostatic, and nonhydrostatic ocean modeling. J Geophys Res 102(C3):5733–5752

    Article  Google Scholar 

  • Martin V (2003) Méthodes de décomposition de domaine de type relaxation dondes pour des équations de locéanographie. Ph.D. thesis, Université Paris, Paris

  • Martin V (2004) An optimized Schwarz waveform relaxation method for the unsteady convection diffusion equation in two dimensions. Comput Fluids 33(5–6):829–837

    Article  Google Scholar 

  • Mason E, Molemaker J, Shchepetkin AF, Colas F, McWilliams JC, Sangrà P (2010) Procedures for offline grid nesting in regional ocean models. Ocean Model 35(1–2):1–15. doi:10.1016/j.ocemod.2010.05.007

    Article  Google Scholar 

  • Mellor GL (2004) Users guide for a three-dimensional, primitive equation, numerical ocean model. Program in atmospheric and oceanic sciences, Princeton University. http://www.aos.princeton.edu/WWWPUBLIC/htdocs.pom/FTPbackup/usersguide0604.pdf

  • Mooers CNK (ed) (1999) Coastal ocean prediction, coastal and estuarine studies, vol 56. American Geophysical Union, Washington, DC

    Google Scholar 

  • MSEAS Group (2010) The Multidisciplinary Simulation, Estimation, and Assimilation Systems (http://mseas.mit.edu/. http://mseas.mit.edu/codes). Reports in Ocean Science and Engineering 6, Department of Mechanical Engineering, Massachusetts Institute of Technology, Cambridge

  • Newhall AE, Gawarkiewicz GG, Lynch JF, Duda TF, McPhee NM, Bahr FB, Marquette CD, Lin YT, Jan S, Wang J, Chen CF, Chiu LYS, Yang YJ, Wei RC, Emerson C, Morton D, Abbot T, Abbot P, Calder B, Mayer L, Lermusiaux PFJ (2010) Acoustics and oceanographic observations collected during the QPE experiment by research vessels OR1, OR2 and OR3 in the east China Sea in the summer of 2009. WHOI Technical Report WHOI-2010-06. Woods Hole Oceanographic Institution, Woods Hole

    Book  Google Scholar 

  • NOAA (2006) National Geophysical Data Center, NGDC coastal relief model. http://www.ngdc.noaa.gov/mgg/coastal/coastal.html

  • Oddo P, Pinardi N (2008) Lateral open boundary conditions for nested limited area models: a scale selective approach. Ocean Model 20(2):134–156

    Article  Google Scholar 

  • Onken R, Robinson AR, Lermusiaux PFJ, Haley PJ Jr, Anderson LA (2003) Data-driven simulations of synoptic circulation and transports in the Tunisia–Sardinia–Sicily region. J Geophys Res 108(C9):8123–8136

    Article  Google Scholar 

  • Onken R, Álvarez A, Fernández V, Vizoso G, Basterretxea G, Tintoré J, Haley P Jr, Nacini E (2008) A forecast experiment in the Balearic Sea. J Mar Syst 71(1–2):79–98

    Article  Google Scholar 

  • Orlanski I (1976) A simple boundary condition for unbounded hyperbolic flows. J Comput Phys 21(3):251–269

    Article  Google Scholar 

  • Penven P, Debreu L, Marchesiello P, McWilliams JC (2006) Evaluation and application of the ROMS 1-way embedding procedure to the central California upwelling system. Ocean Model 12(1–2):157–187. doi:10.1016/j.ocemod.2005.05.002

    Article  Google Scholar 

  • Perkins AL, Smedstad LF, Blake DW, Heburn GW, Wallcraft AJ (1997) A new nested boundary condition for a primitive equation ocean model. J Geophys Res 102(C2):3483–3500

    Article  Google Scholar 

  • Phillips NA (1957) A coordinate system having some special advantages for numerical forecasting. J Meteorol 14(2):184–185

    Google Scholar 

  • Pinardi N, Woods JD (eds) (2002) Ocean forecasting: conceptual basis and applications. Springer, Berlin

    Google Scholar 

  • Pinardi N, Auclair F, Cesarini C, Demirov E, Fonda-Umani S, Giani M, Montanari G, Oddo P, Tonani M, Zavatarelli M (2002) Toward marine environmental prediction in the Mediterranean Sea coastal areas: a monitoring approach. In: Pinardi N, Woods JD (eds) Ocean forecasting: conceptual basis and applications. Springer, Berlin, pp 339–373

    Google Scholar 

  • Rixen M, Le Gac JC, Hermand JP, Peggion G, Coelho E (2009) Super-ensemble forecasts and resulting acoustic sensitivities in shallow waters. J Mar Syst 78:S290–S305

    Article  Google Scholar 

  • Robinson AR (1999) Forecasting and simulating coastal ocean processes and variabilities with the Harvard Ocean Prediction System. In: Mooers CNK (ed) Coastal ocean prediction, coastal and estuarine studies, vol 56. American Geophysical Union, Washington, DC, pp 77–100

    Google Scholar 

  • Robinson AR, Bellingham JG, Chryssostomidis C, Dickey TD, Levine E, Patrikalakis N, Porter DL, Rothschild BJ, Schmidt H, Sherman K, Holliday DV, Atwood DK, the LOOPS Group (1999) Realtime forecasting of the multidisciplinary coastal ocean with the littoral ocean observing and predicting system (loops). In: Third conference on coastal atmospheric and oceanic prediction and processes. American Meteorological Society, Boston

    Google Scholar 

  • Robinson AR, Rothschild BJ, Leslie WG, Bisagni JJ, Borges MF, Brown WS, Cai D, Fortier P, Gangopadhyay A, Haley PJ Jr, Kim HS, Lanerolle L, Lermusiaux PFJ, Lozano CJ, Miller MG, Strout G, Sundermeyer MA (2001) The development and demonstration of an advanced fisheries management information system. In: 17th international conference on interactive information and processing systems for meteorology, oceanography and hydrology. American Meteorological Society, Boston, p 6

    Google Scholar 

  • Robinson AR, Haley PJ, Lermusiaux PFJ, Leslie WG (2002)Predictive skill, predictive capability and predictability in ocean forecasting. In: Proceedings of the “The OCEANS 2002 MTS/IEEE” conference. Holland Publications, pp 787–794

  • Saad Y (2009) SPARSKIT: a basic tool-kit for sparse matrix computations. http://www-users.cs.umn.edu/~saad/software/SPARSKIT/sparskit.html

  • Sapsis TP, Lermusiaux PFJ (2009) Dynamically orthogonal field equations for continuous stochastic dynamical systems. Physica D 238:2347–2360. doi:10.1016/j.physd.2009.09.017

    Article  Google Scholar 

  • Sarkisya AS, Ivanov VF (1971) Joint effect of baroclinicity and bottom relief as an important factor in dynamics of sea currents. Izv Akad Nauk SSSR Fiz Atmos Okeana 7(2):173–188

    Google Scholar 

  • Semtner AJ Jr (2000) Very high-resolution estimates of global ocean circulation, suitable for carbon-cycle modeling. In: Wigley TML, Schimel DS (eds) The carbon cycle. Cambridge University Press, Cambridge, pp 212–228

    Chapter  Google Scholar 

  • Shapiro R (1970) Smoothing, filtering and boundary effects. Rev Geophys Space Phys 8(2):359–387

    Article  Google Scholar 

  • Shchepetkin AF, McWilliams JC (2005) The regional ocean modeling system (roms): A split-explicit, free-surface, topography following coordinates ocean model. Ocean Model 9(4):347–404

    Article  Google Scholar 

  • Shen CY, Evans TE (2004) A free-surface hydrodynamic model for density-stratified flow in the weakly to strongly non-hydrostatic regime. J Comput Phys 200(2):695–717

    Article  Google Scholar 

  • Signell RP, Jenter HL, Blumberg AF (2000) Predicting the physical effects of relocating Boston’s sewage outfall. J Estuar Coast Shelf Sci 50(1):59–72

    Article  Google Scholar 

  • Sloan NQ III (1996) Dynamics of a shelf-slope front: process studies and data-driven simulations in the middle Atlantic Bight. Ph.D. thesis, Harvard University, Cambridge

  • Smith RD, Dukowicz JK, Malone RC (1992) Parallel ocean general circulation modeling. Physica D 60(1–4):38–61

    Article  Google Scholar 

  • Smith WHF, Sandwell DT (1997) Global seafloor topography from satellite altimetry and ship depth soundings. Science 277(5334):1957–1962

    Article  Google Scholar 

  • Spall MA, Holland WR (1991) A nested primitive equation model for oceanic applications. Math Comput Simul 21(2):205–220

    Google Scholar 

  • Spall MA, Robinson AR (1989) A new open ocean, hybrid coordinate primitive equation model. Math Comput Simul 31(3):241–269

    Article  Google Scholar 

  • Stammer D, Chassignet EP (2000) Ocean state estimation and prediction in support of oceanographic research. Oceanography 13(2):51–56

    Google Scholar 

  • de Szoeke RA, Springer SR, Oxilia DM (2000) Orthobaric density: a thermodynamic variable for ocean circulation studies. J Phys Oceanogr 30(11):2830–2852

    Article  Google Scholar 

  • Tanaka S, Bunya S, Westerink JJ, Dawson C, Luettich RA (2010) Scalability of an unstructured grid continuous Galerkin based hurricane storm surge model. J Sci Comput. doi:10.1007/s10915-010-9402-1

    Google Scholar 

  • Tessler ZD, Gordon AL, Pratt LJ, Sprintall J (2010) Transport and dynamics of the Panay Sill overflow in the Philippine seas. J Phys Oceanogr. doi:10.1175/2010JPO4395.1

    Google Scholar 

  • Tian RC, Lermusiaux PFJ, McCarthy JJ, Robinson AR (2004) A generalized prognostic model of marine biogeochemical-ecosystem dynamics: structure, parametrization and adaptive modeling. Harvard Reports in Physical/Interdisciplinary Ocean Science 67, Harvard University, Cambridge

    Google Scholar 

  • Ueckermann MP (2009) Towards next generation ocean models: novel discontinuous Galerkin schemes for 2D unsteady biogeochemical models. Master’s thesis, Massachusetts Institute of Technology, Cambridge

  • Ueckermann MP, Lermusiaux PFJ (2010) High order schemes for 2D unsteady biogeochemical models. Ocean Dyn (this special issue). doi:10.1007/s10236-010-0351-x

    Google Scholar 

  • Wallcraft AJ, Kara AB, Hurlburt HE, Rochford PA (2003) The NRL layered global ocean model (NLOM) with an embedded mixed layer submodel: formulation and tuning. J Atmos Ocean Technol 20(11):1601–1615

    Article  Google Scholar 

  • Warner JC, Sherwood CR, Signell RP, Harris CK, Arango HG (2008) Development of a three-dimensional, regional, coupled wave, current, and sediment-transport model. Comput Geosci 34(10):1284–1306

    Article  Google Scholar 

  • Webb DJ, de Cuevas BA, Coward A (1998) The first main run of the OCCAM global ocean model. Internal Document 34, Southampton Oceanography Centre. http://citeseerx.ist.psu.edu/viewdoc/download?doi=10.1.1.17.5811&rep=rep1&type=pdf

  • WHOI (2006) Shallow water experiment 2006. http://acoustics.whoi.edu/sw06/

  • Wickett ME (1999) A reduced grid method for a parallel global ocean general circulation model. Ph.D. thesis, University of California, Davis, Davis

  • Wubs FW, de Niet AC, Dijkstra HA (2006) The performance of implicit ocean models on B- and C-grids. J Comput Phys 211(1):210–228

    Article  Google Scholar 

  • Xu J, Lermusiaux PFJ (2010) Three dimensional sound propagation in the north Mien Hua Canyon using coupled 4D ocean and 3D acoustics models. JASA Express Lett. Sub Judice

  • Xu J, Lermusiaux PFJ, Haley PJ Jr, Leslie WG, Logutov OG (2008) Spatial and temporal variations in acoustic propagation during the PLUSNet07 exercise in Dabob Bay. In: Proceedings of meetings on acoustics (POMA), vol 4. 155th Meeting Acoustical Society of America. doi:10.1121/1.2988093

Download references

Acknowledgements

We are thankful to W.G. Leslie for his collaboration in the real-time experiments that formed the test-bed for this research and for his help in preparing the figures for this manuscript. We would like to thank C. Lozano and L. Lanerolle for many fruitful discussions during the development of our system. We thank O. Logutov for his tide research and discussions. We are very thankful to the reviewers for their careful and detailed reviews and their very useful suggestions. For our AWACS-SW06 effort, we thank G. Gawarkiewicz, P. Abbot and T. Duda for their ocean data and M. Taylor and J. Hare for their NMFS survey data. We also thank J. Evans, S. Glenn, and J. Wilkin for their real-time WRF atmospheric fluxes and the FNMOC teams for their own products. For our PhilEx effort, we thank A. Gordon, C. Villanoy, C. Lee, and J. Sprintall for their ocean data; H. Hurlburt and J. Metzger for large-scale boundary conditions; and H. Arango and J. Levin for discussions. We also thank B. Leben and CCAR for providing SSH anomaly data. For our QPE effort, we thank G. Gawarkiewicz, T. Duda, J. Sen, B. Cornuelle, J. Lynch, P. Niiler, L. Centurioni, C. Lee, R.-C. Lien, T. Sanford, L. Mayer, B. Calder, and Y.-T. Lin for data and discussions. We also thank J. Doyle, D. Marble, J. Nachimknin, and J. Cook as well as the FNMOC for providing us with atmospheric fluxes.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Pierre F. J. Lermusiaux.

Additional information

Responsible Editor: Eric Deleersnijder

We thank the Office of Naval Research for research support under grants N00014-07-1-1061 and N00014-08-1-1097 (6.1), N00014-07-1-0501 (AWACS-SW06), N00014-07-1-0473 (PHILEX), and N00014-07-1-0241 and N00014-08-1-0586 (QPE) and to the Massachusetts Institute of Technology.

Appendices

Appendix 1: Additional details of the conservative discrete equations

Our horizontal grids are structured Cartesian or spherical grids, either of which can be rotated from the standard geographic orientation (for more details, see Haley 1999, Section 2.2). To obtain the spherical representation of our Eqs. 3744 and “Appendix  1.2”, identify x with longitude and y with latitude. Then multiply all Δx terms by the radius of the earth (R earth) times the cosine of the latitude and all Δy terms by R earth. Our model also includes options for atmospheric forcing, assimilation, tidal forcing, and a river forcing that employs relaxation time constants which can be tuned to reproduce the desired mass, salt, and internal energy (heat) transports. All of these options are compatible with the two-way nesting scheme.

1.1 Appendix 1.1: Vertical grid

In Section 2.4, we introduced our vertical discretization, defining first a set of terrain-following depths for the undisturbed mean sea level, \(z^{\rm MSL}_{i,j,k}\). Here we present the details of \(z^{\rm MSL}_{i,j,k}\). We can currently employ five different schemes for defining these vertical levels, two of which are new:

  1. (a)

    σ-Coordinates (Phillips 1957)

    $$ \label{Eq:SigmaCoords} z^{\rm MSL}_{i,j,k} = - \sigma_k H_{i,j} $$
    (56)

    where 0 ≤ σ k  ≤ 1

  2. (b)

    Hybrid coordinates (Spall and Robinson 1989)

    $$ \label{Eq:HybridCoords} z^{\rm MSL}_{i,j,k} = \left\{ \begin{array}{lcc} \tilde{z}_k & {\rm if} & k \le k_c \\ -h_c - \sigma_k \left( H_{i,j} - h_c\right) & {\rm if} & k > k_c \end{array} \right. $$
    (57)

    where \(\tilde{z}_k\) are a set of constant depths and h c is the sum of the top k c flat level depths

  3. (c)

    Double σ-coordinates (Lozano et al. 1994)

    $$ \label{Eq:DoubleSigmaCoords} z^{\rm MSL}_{i,j,k} = \left\{ \begin{array}{lcc} - \sigma_k \tilde{f}_{i,j} & {\rm if} & k \le k_c \\ -\tilde{f}_{i,j} - (\sigma_k - 1) \left( H_{i,j} - \tilde{f}_{i,j}\right) & {\rm if} & k > k_c \end{array} \right. $$
    (58)
    $$\begin{array}{rll} \tilde{f}_{i,j} &=& \frac{z_{c_1} + z_{c_2}}{2} + \frac{z_{c_2} - z_{c_1}}{2} \\&&\times \tanh\left[\frac{2 \alpha}{z_{c_2} - z_{c_1}} \left(H_{i,j} - h_{ref} \right)\right] \\ \sigma_k &\in& \left\{ \begin{array}{lcc} [0, 1] & {\rm if} & k \le k_c \\ {[1, 2]} & {\rm if} & k > k_c \end{array} \right. \end{array} $$

    where \(\tilde{f}_{i,j}\) is the variable interface depth between the upper and lower σ-systems, \(z_{c_1}\) and \(z_{c_2}\) are the shallow and deep bounds for \(\tilde{f}_{i,j}\), h ref is the reference topographic depth at which the hyperbolic tangent term changes sign, and α is a nondimensional slope parameter (\(|| \nabla \tilde{f} || \le \alpha || \nabla H ||\)).

  4. (d)

    Multi-σ-coordinates This new system is a generalization of the double σ system in which, for Pσ-systems, we define P + 1 nonintersecting interface surfaces. Then the depths are found from

    $$ \begin{array}{rll}\label{Eq:MultiSigmaCoords} z^{\rm MSL}_{i,j,k} &=& -\tilde{f}^{p-1}_{i,j} - (\sigma_k - p + 1) \tilde{f}^p_{i,j} \quad \\&&\quad {\rm for}\quad k_{p-1} < k \le k_p \end{array} $$
    (59)
    $$ \begin{array}{lll}&&\tilde{f}^0_{i,j} = 0 ; \quad \tilde{f}^P_{i,j} = H_{i,j} \\ &&\sigma_k \in {[(p-1),\ p]} \quad {\rm for} \quad k_{p-1} < k \le k_p \end{array} $$

    The intermediate interfaces are free to be chosen from many criteria, including key σ θ surfaces (e.g., top of mean thermocline) or large mean vertical gradients.

  5. (e)

    General coordinates For this new system, we provide a 3D field of level thicknesses, \(\Delta z^{\rm MSL}_{i,j,k}\), under the constraint

    $$ \sum\limits^K_{k=1} \Delta z^{\rm MSL}_{i,j,k} = H_{i,j} \, . $$

    The unperturbed levels are then found from

    $$ \label{Eq:GeneralVerticalCoords} z^{\rm MSL}_{i,j,k} = \left\{ \begin{array}{lcc} \displaystyle\frac{-1}{2} \Delta z^{\rm MSL}_{i,j,1} & {\rm if} & k = 1 \\ z^{\rm MSL}_{i,j,k-1} - \displaystyle\frac{1}{2} \left( \Delta z^{\rm MSL}_{i,j,k-1} + \Delta z^{\rm MSL}_{i,j,k} \right) & {\rm if} & k > 1 \end{array} \right. $$
    (60)

    Note that our new general coordinate scheme contains schemes (a–d) as special cases. Hence, schemes (a–d) are now implemented by specifying \(\Delta z^{\rm MSL}_{i,j,k}\) outside the model, according to their respective rules, and using the resulting \(\Delta z^{\rm MSL}_{i,j,k}\) as input to the general coordinate scheme.

1.2 Appendix 1.2: Fluxes through boundaries of computational cells

To complete the conservative spatial discretizations of Section 2.4, we first establish some notation. Values taken at the centers of tracer volumes have integer indices, e.g., T i,j,k, while values taken at the centers of velocity volumes have odd-half integer indices, e.g., \(\mathbf u_{i+\frac{1}{2},j+\frac{1}{2},k}\). In the vertical, values taken at either the centers of tracer or velocity volumes have integer indices while those at the tops or bottoms of the computational volumes have odd-half integer indices, e.g., \(\omega_{i,j,k+\frac{1}{2}}\). Using these rules, we define the following averaging and differencing operators:

$$ \begin{array}{rll} \label{Eq:avgop} \langle \phi \rangle^x_{i,j,k} &=& \frac{1}{2} \big(\phi_{i+\frac{1}{2},j,k} + \phi_{i-\frac{1}{2},j,k} \big) \\ \langle \phi \rangle^y_{i,j,k} &=& \frac{1}{2} \big(\phi_{i,j+\frac{1}{2},k} + \phi_{i,j-\frac{1}{2},k} \big)\\ \langle \phi \rangle^z_{i,j,k} &=& \frac{1}{2} \big(\phi_{i,j,k+\frac{1}{2}} + \phi_{i,j,k-\frac{1}{2}} \big) \\ \delta^x(\phi)_{i,j,k} &=& \phi_{i+\frac{1}{2},j,k} - \phi_{i-\frac{1}{2},j,k} \\ \delta^y(\phi)_{i,j,k} &=& \phi_{i,j+\frac{1}{2},k} - \phi_{i,j-\frac{1}{2},k}\\ \delta^z(\phi)_{i,j,k} &=& \phi_{i,j,k-\frac{1}{2}} - \phi_{i,j,k+\frac{1}{2}}. \end{array} $$

Note that in the above, i and j increase with increasing x and y while k increases with decreasing depth (negative below sea level).

Now we can define the fluxes through the sides of the computational cells. We start with the “flux velocities” evaluated at the centers of the sides. Following Dukowicz and Smith (1994, Appendix E), we define the integrated flows through the “east” and “north” lateral walls of the tracer volumes as

$$ \upsilon^n_{i+\frac{1}{2},j,k} = \Delta y_j \left\langle \Delta z^n {u^\prime} + \frac{1}{2} \Delta z^{n-1} \left(U^n + U^{n-1} \right) \right\rangle^y_{i+\frac{1}{2},j,k} \, , $$
$$ \nu^n_{i,j+\frac{1}{2},k} = \Delta x_i \left\langle \Delta z^n {v^\prime} + \frac{1}{2} \Delta z^{n-1} \left(V^n + V^{n-1} \right) \right\rangle^x_{i,j+\frac{1}{2},k} \, , $$

while at the velocity boxes, we define the integrated flows through the “east” and “north” lateral walls as

$$ \begin{array}{lll} &&{\kern-6pt} \upsilon^n_{i+1,j+\frac{1}{2},k} \\ && = \Delta y_{j+\frac{1}{2}} \Bigg\langle \Bigg\langle \Bigg\langle \Delta z^n {u^\prime} + \frac{1}{2} \Delta z^{n-1} \left(U^n + U^{n-1} \right) \Bigg\rangle^y \Bigg\rangle^x \Bigg\rangle^y_{i+1,j+\frac{1}{2},k} \, , \end{array} $$
$$\begin{array}{lll} &&{\kern-6pt} \nu^n_{i+\frac{1}{2},j+1,k} \\ &&= \Delta x_{i+\frac{1}{2}} \Bigg\langle \Bigg\langle \Bigg\langle \Delta z^n {v^\prime} + \frac{1}{2} \Delta z^{n-1} \left(V^n + V^{n-1} \right) \Bigg\rangle^x \Bigg\rangle^y \Bigg\rangle^x_{i+\frac{1}{2},j+1,k} \, . \end{array} $$

These particular spatial averagings are chosen to match the discrete transport constraint (Eq. 65 in “Appendix 1.4”). The new aspect here is the temporal evaluations. The baroclinic velocity components are evaluated at time n. However, the timings for the barotropic components are, again, chosen to match the transport constraint (Eq. 65). Also note that these timings assume θ = 1. To get the corresponding flows through the “west”(“south”) lateral walls, simply decrement i(j ) by one.

To evaluate the fluxes through the tops of the computational volumes, we use the above definitions in Eq. 37. At tracer volumes this yields

$$ \begin{array}{lll} \label{Eq:VertFluxTbox} &&{\kern-6pt} \delta^z (\omega^n)_{i,j,k} \Delta x_i \Delta y_j + \delta^x (\upsilon^n)_{i,j,k} + \delta^y(\nu^n)_{i,j,k}\\ &&{\kern2pt} + \frac{\Delta {\cal V}^n_{i,j,k}}{H_{i,j}+\eta^n_{i,j}} \frac{\delta (\eta_{i,j})^{n,n-2}}{\tau} = 0 \end{array} $$
(61)

while at velocity volumes, we get

$$ \begin{array}{lll} &&{\kern-6pt}\delta^z(\omega^n)_{i+\frac{1}{2},j+\frac{1}{2},k} \Delta x_{i+\frac{1}{2}} \Delta y_{j+\frac{1}{2}}\\ &&{\kern2pt} +\, \delta^x(\upsilon^n)_{i+\frac{1}{2},j+\frac{1}{2},k} + \delta^y(\nu^n)_{i+\frac{1}{2},j+\frac{1}{2},k}\\ \label{Eq:VertFluxVbox} &&{\kern2pt} +\, \frac{\Delta {\cal V}^n_{i+\frac{1}{2},j+\frac{1}{2},k}} {H_{i,j}+\langle\langle\eta^n\rangle^x\rangle^y_{i+\frac{1}{2},j+\frac{1}{2}}} \frac{\delta \left(\langle\langle\eta\rangle^x\rangle^y_{i+\frac{1}{2},j+\frac{1}{2}}\right)^{n,n-2}} {\tau}= 0 \, .\\ \end{array} $$
(62)

Using these definitions of the fluxes through the boundaries of the computational volumes, we can now simply write the discrete advection operator as

$$ \begin{array}{rll} \breve{\Gamma}(\phi)^n_{i,j,k} &=& \delta^x\left(\langle\phi^n\rangle^x\upsilon^n\right)_{i,j,k} + \delta^y\left(\langle\phi^n\rangle^y\nu^n\right)_{i,j,k}\\ &&+\; \delta^z\left(\langle\phi^n\rangle^z\omega^n\right)_{i,j,k} \Delta x_i \Delta y_j \, . \end{array} $$

This formulation is valid for both tracer and velocity computational volumes, with the understanding that for velocity volumes the i,j indices are shifted by one half.

We have evaluated the pressure force term, \(-{1 \over \rho_0} \int_{{\cal S}^n} p^n_h \, \hat n_h \cdot d{\boldsymbol{\cal A}}\), both by directly discretizing the integrals of pressure along the cell walls (including the horizontal contributions from the sloping cell tops and bottoms) and by interpolating the pressure to the corresponding velocity depths and evaluating the differential gradient. Both give similar results, but the integral evaluation is conservative and produces less noise in the resulting velocities (especially near sloping bottoms).

1.3 Appendix 1.3: Open boundary conditions

For u , T, S, and η, the application of boundary conditions is straightforward. Our options (see Haley et al. 2009; Lermusiaux 1997) include using values based on data, applying radiation conditions (Orlanski 1976; Spall and Robinson 1989), or, following Perkins et al. (1997), using radiation conditions to correct the provided values. For nested subdomains, we have first used the interpolated values directly or with Perkins et al. (1997) corrections. Some other promising options we have explored with nested subdomains include using the coarse grid values in a narrow buffer zone around the fine domain, which reduces discontinuities. Another important multiscale conservative boundary condition option is to feedback the averages of the fluxes across the boundary walls shared with the large domain (Fig. 3). These include the advective fluxes of momentum and tracers, the pressure force, and the diffusive fluxes of momentum and tracers.

We still need an additional boundary condition for \(\overline{\hat{\cal F}^{n,n-1}}\) since we are unable to directly evaluate Eq. 42 at the boundaries. To derive this boundary condition, we recast Eq. 42 in the form of Eq. 24 and solve for \(\overline{\hat{\cal F}^{n,n-1}}\):

$$ \label{Eq:SetUpUbarBC} \overline{\hat{\cal F}^{n,n-1}} = \frac{\delta(\mathbf{U})}{\tau} + f \hat k \times {\mathbf{U}}^\alpha + g \nabla \eta^\alpha_{i,j} \, . $$
(63)

Now, the right-hand side of Eq. 63 is made up entirely of quantities that can be directly evaluated at the boundary of the velocity grid. For the free surface, we have found that it is more stable to rewrite Eq. 63 in terms of transports:

$$ \begin{array}{rll} \label{Eq:UbarotropicBC} \overline{\hat{\cal F}^{n,n-1}} &=& \frac{1}{H+\eta^n} \left\{\frac{\delta\left[\left(H+\eta\right)\mathbf{U}\right]}{\tau} + f \hat k \times \left[\left(H+\eta\right){\mathbf{U}}\right]^\alpha\right\}\\ &&+\, g \nabla \eta^\alpha_{i,j} \, . \end{array} $$
(64)

Note: when evaluating Eq. 64, only values at time t n + 1 are taken from the provided fields (or nesting interpolations). The fields at times t n and t n − 1 are both already in memory and in primitive equation balance. They are combined with the t n + 1 fields to evaluate (Eq. 64).

Applying Perkins et al. ( 1997 ) boundary conditions:

Following the algorithm of Perkins et al. (1997), corrections to the provided values (and nesting interpolation values) are obtained by applying the Orlanski radiation algorithm to the difference between the PE model values and these provided values and using these differences to correct the boundary values.

For the barotropic transport, however, this is only done for the tangential component to the boundary. The correction to the normal component is derived from the correction obtained for the surface elevation, Δη, and the barotropic continuity equation

$$ \frac{\partial \Delta \eta}{\partial t} + \nabla \, \cdot \, \left[ (H+\eta) \Delta {\mathbf U} \right] = 0 \, . $$

1.4 Appendix 1.4: Maintaining the vertically integrated conservation of mass

To see how the free-surface algorithm maintains the vertically integrated conservation of mass, start from Eq. 44, multiply by θ(H + η n), and take the divergence of the result to get

$$ \begin{array}{lll} \nabla &\cdot& \left[ \left(H+\eta^n\right) \theta \mathbf U^{n+1}\right] = \nabla \cdot \left[ \left(H+\eta^n\right) \theta {\hat {\bf U}}^{n+1}\right]\\ &-& \alpha \theta g \tau \nabla \cdot \left[ \left(H+\eta^n\right) \nabla \delta \eta \right] + \theta \nabla \cdot \left(\left.\mathbf{u^\prime}^n\right|_\eta \frac{\delta \eta}{\tau}\right) \, . \end{array} $$

Substitute for the right-hand side of the above equation from Eq. 43 and rearrange to obtain

$$ \label{Eq:FreeSurfaceVertIntCont} \frac{\delta \eta}{\tau} {\kern-.5pt} +{\kern-.5pt} \frac{1}{2} \nabla {\kern-.5pt}\cdot{\kern-.5pt} \left[ \left(H{\kern-.5pt}+{\kern-.5pt}\eta^n\right) \left( \theta \mathbf U^{n+1} {\kern-.5pt}+{\kern-.5pt} \mathbf U^n {\kern-.5pt}+ (1 {\kern-.5pt}-{\kern-.5pt} \theta) \mathbf U^{n-1} \right) \right] {\kern-.5pt}={\kern-.5pt} 0 . $$
(65)

Equation 65 represents the discrete form of the barotropic continuity enforced by the free-surface algorithm. Imbalances in Eq. 65 produce unrealistic vertical velocities via Eqs. 61 and 62.

However, as illustrated by the above derivation, Eq. 65 is only satisfied to the same degree that Eqs. 43 and 44 are satisfied. This places restrictions on the valid avenues for nesting. For example, we can safely replace the coarse domain estimates of \((H+\eta){\hat {\bf U}}^{n+1}\) with averages from the fine domain without disturbing (Eq. 65). Moving this exchange one step later in the algorithm and trying to average \((H+\eta)[\theta{\hat {\bf U}}^{n+1} + \mathbf U^n + (1 - \theta) \mathbf U^{n-1}]\) would violate Eq. 43, in the sense that we would not be able to make the last substitution leading to Eq. 65 and hence we would violate Eq. 65. Similar to the bias argument of Section 4.7, these fields and their errors will persist over many time steps, leading to unsustainable vertical velocities.

Appendix 2: Additional details on running our fully implicit two-way nested free-surface PE model

1.1 Appendix 2.1: Setting up domains

1.1.1 Appendix 2.1.1: Topography

There are two main issues when defining topographies for nested simulations. The first is that the finer resolution grid can support finer topography scales, including sharper gradients. The bathymetry on the finer grid is not an interpolation of the coarser grid bathymetry, but the coarser grid bathymetry is a coarse-control-volume average of the finer grid bathymetry. The refinement in topographic scales can lead to abrupt artificial discontinuities in the topography where the fine and coarse domains meet. This can be exacerbated by conditioning the topography (Haley and Lozano 2001) to control the hydrostatic consistency condition (Haney 1991). For a given value of the hydrostatic consistency factor (roughly proportional to \(\frac{d\mathbf{x} \times \nabla h}{h}\)), the finer resolution domain can support steeper bathymetric features (e.g., shelfbreak). To ensure a smooth transition, we define a band of points around the outer edge of a fine domain (e.g., a band from the boundary to 6 points inside the boundary, see also Penven et al. 2006). In this band, we replace the fine grid topography with a blend of the coarse and fine grid topographies:

$$ \label{Eq:TopoBlending} h_{\rm blend} = \alpha h_{\rm f\/ine} + (1-\alpha) h_{\rm coarse} $$
(66)

where α varies from zero at the boundary to one at the inner edge of the band (e.g., 6 points).

The second issue comes about from the nesting algorithm itself. As mentioned in Section 3, we force the undisturbed vertical grid , \(z^{\rm MSL}_{i,j,k}\), to satisfy the nesting rules of Eqs. 46 and 48. To ensure that the topographies in nested domains satisfy Eq. 46 and 48 and the blending Eq. 66 ,we usually follows these steps:

  1. 1.

    Apply the nesting constraints on the unconditioned topographies. Starting from the smallest domain, average the fine grid topographies on the successively larger topographies according to Eq. 46. Then starting from the coarsest domain, interpolate the topographies to the boundaries of the successively smaller domains according to Eq. 48.

  2. 2.

    Starting from the largest domain, apply the conditioning. After the largest domain is conditioned, apply the blending Eq. 66 to the second largest. Condition that domain and repeat the blending-conditioning cycle with the successively smaller domains.

  3. 3.

    Reapply the nesting constraints on the conditioned topography. Repeat step 1.

1.1.2 Appendix 2.1.2: Land masking

The first constraint for masking occurs at the boundaries of the finer domains. Considering any two nested domains, we want continuity of the masks across the domain boundary. In other words, a coastline that crosses the boundary of the fine domain should not have a jump or jog at the boundary of the fine domain. Enforcing this consistency, along with boundary constraints on the topography (“Appendix 2.1.1”), enforces consistent estimates of the areas of the lateral boundaries of the fine domain as measured in both the coarse and fine grids.

The second constraint is to have a certain degree of consistency in defining land and sea in the interior of the fine domain. This is a less exact statement because the fine domain supports a more detailed resolution of the land/sea boundary than the coarse domain. Because of the superior resolution, we take the view that the land mask in the interior of the fine domain is “more correct” than the coarse domain mask. Since we use collocated grids, this provides us with a simple algorithm for resetting the coarse mask. For each coarse grid point fully supported by fine grid points, we count how many of the supporting fine grid points are land and how many are sea. If at least one half the fine grid points are sea, the coarse grid point is marked as sea; otherwise, it is masked as land.

Our general procedure is to first define the land mask for the largest (coarsest) domain. Then use that mask to define a crude first guess for the mask in the fine domain. We then reset the interior nodes of the fine mask to better resolve the coasts (leaving a narrow band around the exterior untouched to ensure continuity through the boundary). If we have more then two domains, we use the current domain to initialize the mask for the next finest domain and repeat. When we finish the mask in the smallest (finest) domain, we use that mask to reset the mask in the next coarser domain, using the above sea/land counting procedure. We then examine the modified mask in that next coarser domain to eliminate any spurious artifacts that may have been created (e.g., a narrow mouth of a bay may have been closed leaving an isolated “lake” that we do not need to maintain). We repeat with the next coarser domain and so on until we get back to the coarsest domain.

1.2 Appendix 2.2: Initialization

Our common situation is to estimate the best initial synoptic state from temperature and salinity data (in situ, climatologies, satellite, etc.) but little or no direct velocity data. Our initialization scheme for this situation is described next, focusing mainly on the nesting considerations, first briefly for the rigid-lid procedures and then the extensions for initializations with a free surface.

1.2.1 Appendix 2.2.1: Rigid lid

Our procedures for rigid lid initializations in nested grids are based on (e.g., Haley et al. 2009). Starting from temperature and salinity data, climatologies, etc., we create 3D estimates of temperature and salinity, often using objective analyses (Carter and Robinson 1987; Agarwal and Lermusiaux 2010). From these 3D temperature and salinity estimates, we construct density (Eq. 6) and the hydrostatic pressure (Eq. 9). We then estimate the total velocity using the rigid lid geostrophic relation

$$ f \hat k \times (\mathbf u - \mathbf u_{\rm ref}) = \frac{1}{\rho_0} \nabla p_h = \frac{g}{\rho_0} \int^z_{Z_{\rm ref}} \nabla \rho \, d\zeta$$

where Z ref is a suitably chosen reference level, which can be a “level of no motion”, u ref is the absolute velocity at that depth, and we have interchanged the horizontal gradient with the vertical integral. When evaluating \(\nabla \rho\) at a particular depth, if any of the ρ values used for the gradient would be below topography, we set \(\nabla \rho\) to zero. To enforce no penetration of land, we find a stream function, ψ, which satisfies \(\nabla^2 \psi = \nabla \times \mathbf u\) with ψ set to be constant along coasts. From this ψ, we recompute the velocity. We decompose this velocity into barotropic and baroclinic parts (Eq. 8). The baroclinic portion is fine as is, but barotropic velocities at this stage generally do not satisfy the nondivergence of transport. To enforce this, we define a transport stream function \(\hat k \times \nabla \Psi = H \mathbf U\) and fit it to our estimated barotropic velocities via the Poisson equation

$$ \nabla \times \left( \frac{\hat k}{H} \times \nabla \Psi \right) = \nabla \times \mathbf U \, . $$

We derive Dirichlet boundary conditions for the above by first noting that the tangential derivative of Ψ to the boundary equals the normal component of transport, H U, through the boundary. We then integrate this relation around the boundary to obtain the Dirichlet values. For domains with islands, we also need to provide constant values for Ψ along the island coasts. We do this in a two-stage process in which we first compute Ψ assuming all the islands are open ocean. We then use that initial guess to derive constant island values that minimize the relative interisland transports using fast marching methods (Agarwal 2009).

Nesting considerations

For nesting the initial temperature, salinity, other tracers, and baroclinic velocity, we can directly enforce some conservation constraints by averaging estimates from finer to coarser grids. For the transport stream function, we go to the additional step of generating the Dirichlet boundary values for the Poisson equation in the fine domain by interpolating the stream function values from the coarse domain. This ensures that the same constant of integration is used for both domains and that the net flows through the fine domain are consistent in both the coarse and fine grids. For island values, if the island is represented in both the coarse and fine domains, the coarse domain value is used. If the island is only in the fine domain, then the procedure of the preceding paragraph is used.

1.2.2 Appendix 2.2.2: Free surface

The starting point for the free-surface initialization scheme is the above rigid-lid initialization. We start by explicitly computing the final, rigid-lid barotropic velocities from

$$ \mathbf U = \frac{\hat k}{H} \times \nabla \Psi \, . $$

We next create an equation for the initial surface elevation. We start from the geostrophic approximation with the full pressure

$$ f \hat k \times \mathbf u = g \nabla \eta + \frac{1}{\rho_0} \nabla p_h \, .$$

Integrating this equation in the vertical from − H to 0 and isolating η results in

$$ \label{Eq:FreeSurfaceGradientIC} g H \nabla \eta = f \hat k \times H \mathbf U - \frac{1}{\rho_0} \int^0_{-H} \nabla p_h dz \, . $$
(67)

Finally, we take the divergence of Eq. 67 to get

$$ \label{Eq:FreeSurfaceDivergenceIC} \nabla \cdot \left( g H \nabla \eta \right ) = \nabla \cdot \left( f H \mathbf U \right) - \frac{1}{\rho_0} \nabla \cdot \left( \int^0_{-H} \nabla p_h dz \right) \, . $$
(68)

To generate Dirichlet boundary values for Eq. 68, we integrate the tangential components of Eq. 67 around the boundary. Because the coastal boundary condition is zero normal derivative, no special action needs to be taken for islands.

Once an initial value for η is constructed, then, by Eq. 35, the initial depths are recomputed. The tracers (temperature, salinity, etc.) and baroclinic velocity are re-interpolated to these new initial depths. Finally, the barotropic velocities from the rigid-lid calculation are rescaled to preserve the transports:

$$ \mathbf U_{\rm free\ surface} = \frac{H}{H+\eta} \mathbf U_{\rm rigid\ lid} \, . $$

Nesting considerations

These are the same as for the rigid-lid case. The additional detail is that now we also interpolate the coarse grid estimate of η to generate Dirichlet boundary values for solving Eq. 68 in the fine domain.

1.3 Appendix 2.3: Tidal forcing

1.3.1 Appendix 2.3.1: Constructing the tidal forcing

When adding tidal forcing to our simulations, our underlying assumption is that our regional domains are small enough so that the tidal forcing through the lateral boundaries completely dominates the local body force effects. To model these lateral forcings, we employ the linearized barotropic tidal model (Logutov 2008; Logutov and Lermusiaux 2008). We use a shallow water spectral model and generate 2D fields for the amplitude and phase of tidal surface elevation and the barotropic tidal velocity. We dynamically balance these barotropic tidal fields with our best available topographic and coastal data along with the best exterior barotropic tidal fields (e.g., Egbert and Erofeeva 2002). Once we have constructed our tidal fields for the desired modes, we can simply evaluate them for any time.

The above procedures can provide tidal fields on different grids than used by our PEs. For example, Logutov (2008) and Logutov and Lermusiaux (2008) are formulated on a C-grid, instead of the B-grid being used here. In particular, this means that tidal fields interpolated from these grids will not, in general, exactly satisfy the same discrete continuity as in our grid. Our experience shows that satisfying the same discrete continuity leads to more robust solutions. To enforce this constraint, we solve the constrained minimization problem

$$ \begin{array}{rll} {\cal J} & = & \int \left\{ \alpha \eta^*_1 \eta_1 + \theta \beta \mathbf U^*_1 \cdot \mathbf U_1 + \phi \beta \mathbf U^* \cdot \mathbf U \right.\\ & & \left. + \lambda \Re\left[i\omega \eta + \nabla \cdot \left(H \mathbf U \right) \right] +\, \gamma \Im \left[i\omega \eta + \nabla \cdot \left(H \mathbf U \right) \right] \vphantom{\eta^*_1}\right\} d {\cal V} \nonumber \end{array} $$

where η 0, U 0 as the complex tidal surface elevation and barotropic tidal velocity interpolated from the original grid and η 1, U 1 are the additive “correction” complex tidal surface elevation and barotropic tidal velocity that minimize \(\cal J\),

$$ \eta = \eta_0 + \eta_1 \; ; \quad \mathbf U = \mathbf U_0 + \mathbf U_1 $$

α and β are the weights (including nondimensionalizing factors), λ and γ are the Lagrange multipliers, the superscript * indicate complex conjugation, \(\Re\) and \(\Im\) refer to the real and imaginary parts, and θ, φ are penalty parameters to inhibit unreasonably large total velocities. Using the calculus of variations, the above minimization is equivalent to solving the following system of equations

$$ \begin{array}{rll} \omega \eta_1 - \nabla \cdot \left[ \frac{H^2}{(\theta + \phi) \omega \beta} \nabla \left( \alpha \eta_1 \right) \right] & = & - \omega \eta_0 + i \nabla \notag\\&&\cdot \left( \frac{\theta}{\theta + \phi} H \mathbf U_0 \right) \\ \left. \alpha \right|_{\rm open\ boundary} & = & 0 \\ \left.\left( i \omega \eta_1 + \sqrt{g H} \frac{\partial \eta_1}{\partial n}\right) \right|_{\rm open\ boundary} & = & 0 \\ \mathbf U_1 & = & - \frac{\phi}{\theta + \phi} \mathbf U_0 \nonumber\\&&- i \frac{H}{(\theta + \phi) \omega \beta} \nabla \left(\alpha \eta_1\right) \end{array} $$

Note that the radiation boundary condition does not come from the variations but is a useful addition we are free to make after obtaining α = 0 from the variations.

1.3.2 Appendix 2.3.2: Applying the tidal forcing

We use the barotropic tides both for initialization and for boundary forcing. For the surface elevation, we simply superimpose the tidal surface elevation with the subtidal elevation estimated in “Appendix 2.2.2”. For initialization, this superposition is done over the entire area before the final vertical interpolation of tracers and u . For lateral forcing, this is done at run time in the PE model at the exterior boundaries (and also along two bands inside these boundaries for Perkins et al. (1997) boundary conditions). The resulting boundary values are used to generate Dirichlet boundary conditions for Eq. 43. A similar procedure is used for the barotropic velocities with two notable differences. First, the superposition is performed to preserve transport:

$$ \begin{array}{rll} \left(H + \eta_{\rm superimposed}\right) \mathbf U_{\rm superimposed} &=& \left(H + \eta_{\rm subtidal}\right) \mathbf U_{\rm subtidal}\\ &&+\, H \mathbf U_{\rm tidal} \, . \end{array} $$

Note that the tidal velocity is only multiplied by the undisturbed water depth. This is because we are using a linearized tidal model. The second difference is that the run-time boundary values of the barotropic velocity are used for Eq. 64, not directly applied to the final barotropic velocities.

Nesting considerations

For initialization, the process is as for the unnested case. The superpositions described above are done for the initial conditions of each domain. For the lateral forcing, however, the barotropic tidal fields are only applied at the boundaries of the coarsest domain. The reason being that applying the barotropic tidal fields at the boundary of the coarsest domain can produce the full tidal response in the interior, and hence, the barotropic tidal fields are unnecessary for the nested subdomains.

1.4 Appendix 2.4: Solving the equation for the surface elevation

Equation 43, with Dirichlet boundary conditions, represents an elliptic system of equations for the surface elevation, η. To numerically solve this system, we use a preconditioned conjugate gradient solver for sparse matrices (e.g., SPARSKIT; Saad 2009). A typical convergence test for such an iterative solver is an integrated measure of the reduction in the norm of the residual over all points. Specifically, if r is the residual of the current solver iteration and r 0 is the residual of the initial guess, the convergence test is

$$\Vert r \Vert \le \tau_r \Vert r_0 \Vert + \tau_a$$

where τ r is the relative tolerance and τ a is the absolute tolerance. In practice, we tend to use very small values (10 − 12 and 10 − 25, respectively) to ensure a tight convergence. Using the results of “Appendix 3”, we also supplement this global constraint with the following point-wise constraint:

$$ \begin{array}{rll} \left\vert \frac{\partial \delta\eta^k}{\partial x} - \frac{\partial \delta\eta^{k-1}}{\partial x} \right\vert &\le& \tau^g_r \left\vert \frac{\partial \delta\eta^k}{\partial x} \right\vert + \tau_a \; ; \quad \\ \left\vert \frac{\partial \delta\eta^k}{\partial y} - \frac{\partial \delta\eta^{k-1}}{\partial y} \right\vert &\le& \tau^g_r \left\vert \frac{\partial \delta\eta^k}{\partial y} \right\vert + \tau_a\\ \left\vert \delta\eta^k - \delta\eta^{k-1} \right\vert \le \tau^g_r \left\vert \delta\eta^k \right\vert + \tau_a \end{array} $$

where the superscript k refers to the iteration number and \(\tau^g_r\) is the relative tolerance for the gradient test (typically around 10 − 8). Here we test on both δη and its gradients to ensure the relative convergence of the barotropic velocities (Eq. 44).

Since we have discretized our equations on the B-grid, both Eq. 43 and, especially, Eq. 68 possess a well-known checkerboard mode in their null spaces (Deleersnijder and Campin 1995; le Roux et al. 2005; Wubs et al. 2006). For realistic geometries, we found that applying a Shapiro filter (Shapiro 1970) to the solution was sufficient to suppress the noise while maintaining the correct physical features. The one case where this approach failed was in creating an initialization for an idealized flow in a periodic channel. The lack of Dirichlet boundary values in that case and corresponding lack of structure they would have imposed allowed the checkerboard mode to suppress all other structures. To control this error, the matrix in Eq. 68 was augmented with a Laplacian filter (Deleersnijder and Campin 1995; Wickett 1999) to prevent the appearance of this mode. Again, this filter was only needed for the idealized periodic channel flow.

Appendix 3: Review of rigid-lid nesting algorithm

One of our first nesting schemes (see Section 4.1) for the nonlinear free-surface version of the PEs was built by analogy with our two-way nesting scheme (e.g., Spall and Holland 1991; Fox and Maskell 1995; Sloan 1996) for the HOPS rigid-lid formulation of the PEs. We have used this scheme in a variety of dynamical situations (Robinson et al. 2002; Leslie et al. 2008; Haley et al. 2009) and we present it next.

We start with the rigid-lid PEs:

$$ \nabla \cdot \mathbf u + \frac{\partial w}{\partial z} = 0 \, , $$
$$\frac{\partial \mathbf{u^\prime}}{\partial t} + f \hat k \times \mathbf{u^\prime} = \widehat{\cal F} - \overline{\widehat{\cal F}} \, , $$
$$\nabla \times \left( \frac{\hat k}{H} \times \nabla \delta \Psi \right) + \alpha_{\rm RL} \nabla \times \left[ f \, \hat k \times \left( \frac{\hat k}{H} \times \nabla \delta \Psi \right) \right] = \nabla \times \overline{\widehat{\cal F}} \,\, , $$
$$p_h = - \int^0_z \rho g \, d\zeta \, , $$
$$\frac{D T}{D t} = F^T \, , $$
$$\frac{D S}{D t} = F^S \, , $$
$$\rho = \rho (z,T,S) $$

where

$$ \mathbf u = \mathbf{u^\prime} + \mathbf U \; ; \quad \mathbf U = \frac{1}{H} \int^0_{-H} \mathbf u \, dz \; ; \quad H \mathbf U = \hat k \times \nabla \Psi \, , $$
$$ \widehat{\cal F} = -{1 \over \rho_0} \nabla p_h - {\boldsymbol{\Gamma}}(\mathbf u) + \mathbf F \; ; \quad \overline{\widehat{\cal F}} = \frac{1}{H} \int^0_{-H} \widehat{\cal F} \, dz $$
$$ {\boldsymbol \Gamma}(\mathbf u) = \left( \begin{array}{c} \Gamma(u) \\ \Gamma(v)\end{array}\right) \; ; \quad \Gamma(\phi) = u \frac{\partial \phi}{\partial x} + v \frac{\partial \phi}{\partial y} + w \frac{\partial \phi}{\partial z} . $$

Here we have introduced the transport stream function, Ψ, to replace the barotropic velocity as a state variable. Using this notation, we write our rigid-lid nesting algorithm as

  1. 1.

    Solve the rigid-lid PEs simultaneously in each domain for (u n + 1, T n + 1, S n + 1)

  2. 2.

    Replace values in the coarse domain at overlap nodes with the following averages from the fine domain values

    $$ \phi^{n+1}_{i_c,j_c,k} = \frac{1}{\Delta {\cal V}_{i_c,j_c}} \sum\limits^{j_{fc}+r_h}_{j=j_{fc}-r_h} \sum\limits^{i_{fc}+r_h}_{i=i_{fc}-r_h} \phi^{n+1}_{i,j,k} \Delta {\cal V}_{i,j,k} \, , $$
    (69)
    $$ \left(\nabla \times \overline{\widehat{\cal F}}\right)_{i_c,j_c} = \frac{1}{\Delta {\cal A}_{i_c,j_c}} \sum\limits^{j_{fc}+r_h}_{j=j_{fc}-r_h} \sum\limits^{i_{fc}+r_h}_{i=i_{fc}-r_h} \left(\nabla \times \overline{\widehat{\cal F}}\right)_{i,j} \Delta {\cal A}_{i,j}\nonumber\\ $$
    (70)

    where r is the ratio of the resolution of the coarse grid to fine grid,

    $$ \begin{array}{lll} &&{\kern-6pt} \phi = \mathbf{u^\prime}, T, S \; ; \quad \Delta {\cal V}_{i,j,k} = \Delta x_{i,j}\Delta y_{i,j} \Delta z_{i,j,k} \; ; \\ &&{\kern-6pt} \Delta {\cal A}_{i,j} = \Delta x_{i,j}\Delta y_{i,j} \; ; \quad r_h = \lfloor r/2\rfloor \; . \end{array} $$
  3. 3.

    In the coarse domain, solve the rigid-lid barotropic momentum equation for Ψn + 1.

  4. 4.

    Using piece-wise bi-cubic Bessel interpolation, \(\cal B\), replace values in the fine grid boundary with values interpolated from the coarse grid

    $$ \begin{array}{rll} \phi^{n+1}_{i_{fb},j_{fb},k} &=& {\cal B}\left(\phi^{n+1}_{i_c,j_c,k}\right) \, , \\ \mathbf{u^\prime}^{n+1}_{i_{fb},j_{fb},k} \Delta z_{i_{fb},j_{fb},k} &=& {\cal B}\left(\mathbf{u^\prime}^{n+1}_{i_c,j_c,k} \Delta z_{i_c,j_c,k}\right) \, , \end{array} $$

    where

    $$ \phi = T,\,S,\,\Psi \, . $$

    Note that the interpolation of baroclinic velocity is written in terms of transport rather than velocity. This is done to preserve the baroclinicity of u in the fine domain.

  5. 5.

    In the fine domain, solve the barotropic momentum equation for Ψn + 1.

We found that this scheme maintains consistency between the estimates on the coarse and fine grids provided that a sufficiently stringent convergence criterion is used when solving for Ψn + 1. Only using a test on the integral of the residuals did not always maintain the consistency. Such a test is global in nature and can give different convergence results in the coarse and fine domains. Instead, we found the best results occurred when we supplemented the residual tests with the following point-wise test:

$$ \left| \frac{\frac{\partial \Psi^{k+1}}{\partial x} - \frac{\partial \Psi^k}{\partial x}} {\frac{\partial \Psi^k}{\partial x}} \right| \le \epsilon \; ; \quad \left| \frac{\frac{\partial \Psi^{k+1}}{\partial y} - \frac{\partial \Psi^k}{\partial y}} {\frac{\partial \Psi^k}{\partial y}} \right| \le \epsilon $$

where here the superscripts k refer to the iteration count in the solver and ϵ is the relative tolerance. We test the derivative of Ψ instead of Ψ so that the convergence is on the relative change in the transport, a physically more meaningful quantity than Ψ itself. This test ensures that in both domains the iterative solution proceeds at least until the specified tolerance is reached at every point, thereby ensuring at least that level of consistency between the solutions in the coarse and fine domains.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Haley, P.J., Lermusiaux, P.F.J. Multiscale two-way embedding schemes for free-surface primitive equations in the “Multidisciplinary Simulation, Estimation and Assimilation System”. Ocean Dynamics 60, 1497–1537 (2010). https://doi.org/10.1007/s10236-010-0349-4

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10236-010-0349-4

Keywords

Navigation