Skip to main content

Morphodynamic Modelling in Marine Environments: Model Formulation and Solution Techniques

  • Chapter
  • First Online:
The Mathematics of Marine Modelling

Part of the book series: Mathematics of Planet Earth ((MPE,volume 9))

  • 503 Accesses

Abstract

The bathymetry and geometry of coastal seas, barrier coasts and estuaries strongly influence tides and currents, and the associated transport of sediments. In turn, these transports result in a constantly evolving bathymetry and geometry, thus resulting in a feedback loop between bathymetry and geometry, water motion and sediment transport. To capture this evolution, morphodynamic models are employed. In this chapter, first the conservation laws are derived, resulting in a system of strongly coupled partial differential equations that model the morphodynamic evolution. Subsequently, two different solution strategies, indicated as the initial value and the bifurcation approach, are discussed. In the former approach, the emphasis is on the temporal evolution of bathymetric patterns, whereas the latter approach focuses on the direct identification of asymptotic states of the system under consideration. To exemplify these two approaches, the morphodynamic evolution and asymptotic states of a short, rectangular tidal inlet are considered, showing that these two model approaches result in different and complementary insights.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 119.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 159.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 159.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Similar content being viewed by others

Notes

  1. 1.

    For later reference, long–term time scales are defined in this chapter as time scales much larger than both the forcing and morphodynamic time scales.

  2. 2.

    For a derivation based on spatial averaging of the sub–particle–scale differential equation of solid mass conservation, see Coleman and Nikora (2009).

  3. 3.

    The exchange of sediment between the water and/or sediment layer on the one hand and the air on the other hand is not considered here. These fluxes can either be prescribed at the respective interfaces, or may be modeled by introducing a fourth layer that describes the dynamics of sediment in the air (aeolian transport) and the exchange of sediment over the appropriate interfaces.

  4. 4.

    If only a limited amount of erodible sediment is available, the evolution of a so–called bottom pool, that can be depleted, has to be explicitly taken into account, see Burchard and Baumert (1998), Brouwer et al. (2018), Burchard et al. (2018).

  5. 5.

    The distinction between bedload and suspended load can be made as follows: the bedload is that part of the load which is travelling in the active layer of the bed, supported by intergranular collisions rather than by fluid turbulence (Wilson 1966). Suspended sediment is that part of the transported particles which mainly is supported by the fluid turbulence.

References

  • Adams, C.E., and G.L. Weatherly. 1981. Some effects of suspended sediment stratification on an oceanic bottom boundary layer. Journal of Geophysical Research 86 (C5): 4161–4172.

    Article  Google Scholar 

  • Ardhuin, F., N. Rascle, and K.A. Belibassakis. 2008. Explicit wave-averaged primitive equations using a generalized Lagrangian mean. Ocean Modelling 20 (1): 35–60.

    Article  Google Scholar 

  • Ardhuin, F., N. Rascle, and K.A. Belibassakis. 2017. Erratum: Corrigenda of ‘explicit wave-averaged primitive equations using a generalized Lagrangian mean’. Ocean Modelling 113: 185–186.

    Article  Google Scholar 

  • Ariathurai, C.R. 1974. A finite element model of cohesive sediment transportation. PhD thesis, Ph. D. Dissertation, University California, Davis, 1974.

    Google Scholar 

  • Ashton, A., A.B. Murray, and O. Arnault. 2001. Formation of coastline features by large-scale instabilities induced by high-angle waves. Nature 414: 296–300.

    Article  Google Scholar 

  • Bennis, A.-C., F. Ardhuin, and F. Dumas. 2011. On the coupling of wave and three-dimensional circulation models: Choice of theoretical framework, practical implementation and adiabatic tests. Ocean Modelling 40 (3–4): 260–272.

    Article  Google Scholar 

  • Boelens, T., T. Qi, H.M. Schuttelaars and T. De Mulder. 2021. Morphodynamic equilibria in short tidal basins using a 2DH exploratory model. Journal of Geophysical Research: Earth Surface, 126 (3).

    Google Scholar 

  • Bolla Pittaluga, M., N. Tambroni, A. Canestrelli, R. Slingerland, S. Lanzoni, and G. Seminara. 2015. Where river and tide meet: The morphodynamic equilibrium of alluvial estuaries. Journal of Geophysical Research: Earth Surface 120 (1): 75–94.

    Article  Google Scholar 

  • Booij, N., R.C. Ris and L.H. Holthuijsen. 1999. A third-generation wave model for coastal regions 1. Model description and validation. Journal of Geophysical Research: Oceans, 104 (C4), 7649–7666.

    Google Scholar 

  • Brouwer, R.L., G.P. Schramkowski, Y.M. Dijkstra, and H.M. Schuttelaars. 2018. Time evolution of estuarine turbidity maxima in well-mixed, tidally dominated estuaries: The role of availability- and erosion-limited conditions. Journal of Physical Oceanography 48: 1629–1650.

    Article  Google Scholar 

  • Brouwer, R.L., H.M. Schuttelaars, and P.C. Roos. 2013. Modelling the influence of spatially varying hydrodynamics on the cross-sectional stability of double inlet systems. Ocean Dynamics 63 (11–12): 1263–1278.

    Article  Google Scholar 

  • Brouwer, R.L., J. Van de Kreeke, and H.M. Schuttelaars. 2012. Entrance/exit losses and cross-sectional stability of double inlet systems. Estuarine, Coastal and Shelf Science 107: 69–80.

    Article  Google Scholar 

  • Bruun, P. 1962. Sea level rise as a cause of shore erosion. Journal of the Waterways and Harbors Division 88: 117–130.

    Article  Google Scholar 

  • Burchard, H., and H. Baumert. 1998. The formation of estuarine turbidity maxima due to density effects in the salt wedge. A hydrodynamic process study. Journal of Physical Oceanography 28: 309–321.

    Article  Google Scholar 

  • Burchard, H., H.M. Schuttelaars, and D.K. Ralston. 2018. Sediment trapping in estuaries. Annual Review of Marine Science 10: 371–395.

    Article  Google Scholar 

  • Coleman, S.E., and V.I. Nikora. 2009. Exner equation: A continuum approximation of a discrete granular system. Water Resources Research 45: W09421.

    Article  Google Scholar 

  • Crisfield, M.A. 1981. A fast incremental/iterative solution procedure that handles “snap-through’’. Computational and Structural 13 (1–3): 55–62.

    Article  MATH  Google Scholar 

  • Dam, G., M. van der Wegen, R.J. Labeur, and D. Roelvink. 2016. Modeling centuries of estuarine morphodynamics in the Western Scheldt estuary. Geophysical Research Letters 43 (8): 3839–3847.

    Article  Google Scholar 

  • Dastgheib, A., J.A. Roelvink, and Z.B. Wang. 2008. Long-term process - based morphological modelling of the Marsdiep Tidal Basin. Marine Geology 256: 90–100.

    Article  Google Scholar 

  • De Vriend, H.J. 1996. Mathematical modelling of meso–tidal barrier island coasts. Part I: Empirical and semi–emperical models. In Advances in coastal and ocean engineering, ed. P.L.-F Liu, 115–149. Singapore: World Scientific.

    Google Scholar 

  • De Vriend, H.J. 2001. Long-term morphological prediction. In River, coastal and estuarine morphodynamics, ed. G. Seminara and P. Blondeaux, 163–190. Berlin: Springer.

    Chapter  Google Scholar 

  • De Vriend, H.J., M. Capobianco, T. Chesher, H.E. de Swart, B. Latteux, and M.J.F. Stive. 1993. Approaches to long-term modelling of coastal morphology: A review. Coastal Engineering 21: 225–269.

    Article  Google Scholar 

  • Dean, R.G. 1990. Equilibrium beach profiles: Characteristics and applications. Journal of Coastal Research 7: 53–84.

    Google Scholar 

  • Deng, X., C. Meerman, T. Boelens, T. De Mulder, P. Salles and H.M. Schuttelaars, H.M. 2021. Morphodynamic equilibria in double-inlet systems: Existence and stability. Journal of Geophysical Research: Earth Surface, 126 (12).

    Google Scholar 

  • Dijkstra, H.A., F.W. Wubs, A.K. Cliffe, E. Doedel, I.F. Dragomirescu, B. Eckhardt, A.Y. Gelfgat, A.L. Hazel, V. Lucarini, A.G. Salinger, E.T. Phipps, S.-U. Juan, H. Schuttelaars, L.S. Tuckerman, and U. Thiele. 2014. Numerical bifurcation methods and their application to fluid dynamics: Analysis beyond simulation. Communications in Computational Physics 15: 1–45.

    Google Scholar 

  • Dijkstra, Y.M., H.M. Schuttelaars, G.P. Schramkowski, and R.L. Brouwer. 2019. Modeling the transition to high sediment concentrations as a response to channel deepening in the Ems River Estuary. Journal of Geophysical Research: Oceans 124: 1578–1594.

    Article  Google Scholar 

  • Dingemans, M.W. (1997). Water wave propagation over uneven bottoms, Vol. 1 and 2. Advanced Series on Ocean Engineering, Vol. 13. World Scientific.

    Google Scholar 

  • Dissanayake, D.M.P.K., J.A. Roelvink, and M. van der Wegen. 2009. Modelled channel patterns in a schematized tidal inlet. Coastal Engineering 56 (11–12): 1069–1083.

    Article  Google Scholar 

  • Dodd, N.P., D. Blondeaux, H.E. de Calvete, A. Swart, S.J.M.H. Falqués, and G. Rózy.ński Hulscher, and G. Vittori. 2003. Understanding coastal morphodynamics using stability methods. Journal of Coastal Research 19 (4): 849–865.

    Google Scholar 

  • Dyer, K.R. 1986. Coastal and estuarine sediment dynamics. Chichester: Wiley.

    Google Scholar 

  • Dyer, K.R., and R.L. Soulsby. 1988. Sand transport on the continental shelf. Annual Review of Fluid Mechanics 20: 295–324.

    Article  Google Scholar 

  • Escoffier, F.F. 1940. The stability of tidal inlets. Shore Beach 8: 114–115.

    Google Scholar 

  • Exner, F.M. 1920. Zur physik der dünen. Akad. Wiss. Wien Math. Naturwiss. Klasse, 129

    Google Scholar 

  • Exner, F.M. 1925. Über die wechselwirkung zwischen wasser und geschiebe in flüssen. Akad. Wiss. Wien Math. Naturwiss. Klasse, 134

    Google Scholar 

  • Fredsøe, J., and R. Deigaard. 1992. Mechanics of coastal sediment transport. Singapore: World Scientific.

    Book  Google Scholar 

  • Ganju, N.K., D.H. Schoellhamer and B.E. Jaffe. 2009. Hindcasting of decadal-timescale estuarine bathymetric change with a tidal-timescale model. Journal of Geophysical Research: Earth Surface, 114 (4).

    Google Scholar 

  • Green, M.O., and G. Coco. 2014. Review of wave-driven sediment resuspension and transport in estuaries. Reviews of Geophysics 52 (1): 77–117.

    Article  Google Scholar 

  • Guckenheimer, J. and P. Holmes. 1983. Nonlinear oscillations, dynamical systems, and bifurcations of vector fields, volume 42 of Applied mathematical sciences. New York: Springer-Verlag.

    Google Scholar 

  • Guo, L., M. van der Wegen, Z.B. Wang, D. Roelvink, and Q. He. 2016. Exploring the impacts of multiple tidal constituents and varying river flow on long-term, large-scale estuarine morphodynamics by means of a 1-d model. Journal of Geophysical Research: Earth Surface 121 (5): 1000–1022.

    Article  Google Scholar 

  • Hepkema, T.M., H.E. De Swart, and H.M. Schuttelaars. 2019. The sensitivity of tidal bar wavelength to channel width. Journal of Geophysical Research: Earth Surface 124: 2417–2436.

    Article  Google Scholar 

  • Hibma, A., H.M. Schuttelaars, and Z.B. Wang. 2003. Comparison of longitudinal equilibrium profiles of estuaries in idealised and process-based models. Ocean Dynamics 53: 252–269.

    Article  Google Scholar 

  • Holthuijsen, L.H., N. Booij, and T.H.C. Herbers. 1989. A prediction model for stationary, short-crested waves in shallow water with ambient currents. Coastal Engineering 13 (1): 23–54.

    Article  Google Scholar 

  • Horemans, D.M.L., Y.M. Dijkstra, H.M. Schuttelaars, P. Meire, and T.J.S. Cox. 2020. Unraveling the essential effects of flocculation on large-scale sediment transport patterns in a tide-dominated estuary. Journal of Physical Oceanography 50 (7): 1957–1981.

    Article  Google Scholar 

  • Hulscher, S.J.M.H., and G.M. van den Brink. 2001. Comparison between predicted and observed sand waves and sand banks in the North sea. Journal of Geophysical Research: Oceans 106 (C5): 9327–9338.

    Article  Google Scholar 

  • Jay, D.A., P.M. Orton, T. Chisholm, D.J. Wilson, and A.M.V. Fain. 2007. Particle trapping in stratified estuaries: Consequences of mass conservation. Estuaries and Coasts 30: 1095–1105.

    Article  Google Scholar 

  • Kandiah, A. 1974. Fundamental aspects of surface erosion of cohesive soils. PhD thesis, University of California, Davis.

    Google Scholar 

  • Keller, H.B. 1977. Numerical solution of bifurcation and nonlinear eigenvalue problems. In Applications of Bifurcation Theory, ed. P. Rabinowitz, 359–384. New York: Academix Press.

    Google Scholar 

  • Komen, G.J., L. Cavaleri, M. Donelan, K. Hasselmann, S. Hasselmann, and P.A.E.M. Janssen. 1994. Dynamics and modelling of ocean waves. Cambridge University Press.

    Google Scholar 

  • Krol, M.S. 1990. The method of averaging in partial differential equations. PhD thesis, University of Utrecht, The Netherlands.

    Google Scholar 

  • Krol, M.S. 1991. On the averaging method in nearly time-periodic advection-diffusion problems. SIAM Journal on Applied Mathematics 51 (6): 1622–1637.

    Article  MathSciNet  MATH  Google Scholar 

  • Kumar, M., H.M. Schuttelaars, and P.C. Roos. 2017. Three-dimensional semi-idealized model for estuarine turbidity maxima in tidally dominated estuaries. Ocean Modelling 113: 1–21.

    Article  Google Scholar 

  • Kuznetsov, Y.A. 2004. Elements of applied bifurcation theory, volume 112 of Applied Mathematical Sciences. New York: Springer

    Google Scholar 

  • Lanzoni, S., and G. Seminara. 2002. Long-term evolution and morphodynamic equilibrium of tidal channels. Journal of Geophysical Research: Oceans 107: 1–13.

    Article  Google Scholar 

  • Latteux, B. 1995. Techniques for long-term morphological simulation under tidal action 126: 129–141.

    Google Scholar 

  • Lesser, G.R., J.A. Roelvink, J.A.T.M. van Kester, and G.S. Stelling. 2004. Development and validation of a three-dimensional morphological model. Coastal Engineering 51: 883–915.

    Article  Google Scholar 

  • Longuet-Higgins, M.S., and R.W. Stewart. 1964. Radiation stresses in water waves; a physical discussion, with applications. Deep Sea Research and Oceanographic Abstracts 11 (4): 529–562.

    Google Scholar 

  • Lorentz, H.A. 1922. Het in rekening brengen van den weerstand bij schommelende vloeistofbewegingen. De Ingenieur, 695.

    Google Scholar 

  • Luijendijk, A.P., M.A. de Schipper and R. Ranasinghe. 2019. Multi-timescale predictions of complex sandy interventions. Journal of Marine Science and Engineering, 7 (3).

    Google Scholar 

  • Marciano, R., Z.B. Wang, A. Hibma, H.J. de Vriend and A. Defina. 2005. Modeling of channel patterns in short tidal basins. Journal of Geophysical Research: Earth Surface, 110 (1).

    Google Scholar 

  • McAnally, W.H., and A.J. Mehta. 2002. Significance of aggregation of fine sediment particles in their deposition. Estuarine, Coastal and Shelf Science 54: 643–653.

    Article  Google Scholar 

  • McWilliams, J.C., J.M. Restrepo, and E.M. Lane. 2004. An asymptotic theory for the interaction of waves and currents in coastal waters. Journal of Fluid Mechanics 511: 135–178.

    Article  MathSciNet  MATH  Google Scholar 

  • Mei, C.C. 1983. The applied dynamics of ocean surface waves. In Advanced series on ocean engineering, vol. 1. Singapore: World Scientific Publishing Co. Pvt. Ltd.

    Google Scholar 

  • Mellor, G. 2003. The three-dimensional current and surface wave equations. Journal of Physical Oceanography 33 (9): 1978–1989.

    Article  MathSciNet  Google Scholar 

  • Mellor, G. 2011. Wave radiation stress. Ocean Dynamics 61 (5): 563–568.

    Article  Google Scholar 

  • Mellor, G. 2015. A combined derivation of the integrated and vertically resolved, coupled wave-current equations. Journal of Physical Oceanography 45 (6): 1453–1463.

    Article  Google Scholar 

  • Meyer-Peter, E. and R. Müller. 1948. Formulas for bedload transport. In World congress of the international association for hydro-environment engineering and research (IAHR).

    Google Scholar 

  • Moghimi, S., K. Klingbeil, U. Gräwe, and H. Burchard. 2013. A direct comparison of a depth-dependent radiation stress formulation and a vortex force formulation within a three-dimensional coastal ocean model. Ocean Modelling 70: 132–144.

    Article  Google Scholar 

  • Munk, W.H., and E.R. Anderson. 1948. Notes on a theory of the thermocline. Journal of Marine Research 7: 276–295.

    Google Scholar 

  • Murray, A.B. 2003. Contrasting the goals, strategies, and predictions associated with simplified numerical models and detailed simulations., vol. 135. Predicition in geomorphologicy. American Geophysical Union, Geophysical Monograph.

    Google Scholar 

  • Murray, A.B. 2007. Reducing model complexity for explanation and prediction. Geomorphology 90: 178–191.

    Article  Google Scholar 

  • Murray, A.B. 2013. Treatise on Geomorphology, volume 2 (Quantitative Modeling of Geomorphology), chapter Which models are good (enough), and when? 50–58. San Diego, CA: Academic Press

    Google Scholar 

  • Murray, A.B., G. Coco and E.B. Goldstein. 2014. Cause and effect in geomorphic systems: Complex systems perspectives. Geomorphology 214: 1–9.

    Google Scholar 

  • Murray, A.B., and C. Paola. 1994. A cellular model of braided rivers. Nature 371: 54–57.

    Article  Google Scholar 

  • Nihoul, J.C.J. 1975. Hydrodynamic models. In Modelling of marine systems, ed. J.C.J. Nihoul, 41–66. Amsterdam: Elsevier Scientific Publishing Company.

    Chapter  Google Scholar 

  • Nnafie, A., T. van Oyen, B. De Maerschalck, M. van der Vegt, and M. van der Wegen. 2018. Estuarine channel evolution in response to closure of secondary basins: An observational and morphodynamic modeling study of the western scheldt estuary. Journal of Geophysical Research: Earth Surface 123 (1): 167–186.

    Article  Google Scholar 

  • O’Brien, M.P. 1931. Estuary tidal prism related to entrance areas. Civil Engineering 1: 738–739.

    Google Scholar 

  • O’Brien, M.P. 1969. Equilibrium flow areas of inlets on sandy coasts. Journal of the Waterways and Harbors Division 95: 43–52.

    Article  Google Scholar 

  • Olabarrieta, M., W.R. Geyer, and N. Kumar. 2014. The role of morphology and wave-current interaction at tidal inlets: An idealized modeling analysis. Journal of Geophysical Research: Oceans 119 (12): 8818–8837.

    Article  Google Scholar 

  • Paola, C., and V.R. Voller. 2005. A generalized Exner equation for sediment mass balance. Journal of Geophysical Research: Earth Surface 110: F04014.

    Article  Google Scholar 

  • Pape, L., and B.G. Ruessink. 2011. Neural-network predictability experiments for nearshore sandbar migration. Continental Shelf Research 31: 1033–1042.

    Article  Google Scholar 

  • Pritchard, D., A.J. Hogg, and W. Roberts. 2002. Morphological modelling of intertidal mudflats: The role of cross-shore tidal currents. Continental Shelf Research 22 (11–13): 1887–1895.

    Article  Google Scholar 

  • Ranasinghe, R., C. Swinkels, A. Luijendijk, D. Roelvink, J. Bosboom, M. Stive, and D. Walstra. 2011. Morphodynamic upscaling with the MORFAC approach: Dependencies and sensitivities. Coastal Engineering 58: 806–811.

    Article  Google Scholar 

  • Reef, K.R.G., P.C. Roos, H.M. Schuttelaars, and S.J.M.H. Hulscher. 2020. Influence of back-barrier basin geometry on multiple tidal inlet systems: The roles of resonance and bottom friction. Journal of Geophysical Research: Earth Surface 125 (3): 2020.

    Google Scholar 

  • Reeve, D.E., H. Karunarathna, S. Pan, J.M. Horrillo-Caraballo, G. Różyński, and R. Ranasinghe. 2016. Data-driven and hybrid coastal morphological prediction methods for mesoscale forecasting. Geomorphology 256: 49–67.

    Article  Google Scholar 

  • Richardson, J.F., and W.N. Zaki. 1954. The sedimentation of a suspension of uniform spheres under conditions of viscous flow. Chemical Engineering Science 8: 65–78.

    Article  Google Scholar 

  • Ris, R.C., N. Booij and L.H. Holthuijsen. 1994. Spectral wave model for the coastal zone. In Proceedings of the 2nd international symposium on ocean wave measurement and analysis, 1993, 630–641.

    Google Scholar 

  • Roelvink, J.A. 2006. Coastal morphodynamic evolution techniques. Coastal Engineering 53: 277–287.

    Article  Google Scholar 

  • Roelvink, J.A., D.-J.R. Walstra, M. van der Wegen and R. Ranasinghe. 2016. Modeling of coastal morphological processes, 611–634. Cham: Springer International Publishing.

    Google Scholar 

  • Roos, P.C., H.M. Schuttelaars, and R.L. Brouwer. 2013. Observations of barrier island length explained using an exploratory morphodynamic model. Geophysical Research Letters 40: 4338–4343.

    Article  Google Scholar 

  • Sanders, J.A., and F. Verhulst. 1985. Averaging methods in nonlinear dynamical systems. New York: Springer.

    Book  MATH  Google Scholar 

  • Sanford, L.P. 2008. Modeling a dynamically varying mixed sediment bed with erosion, deposition, bioturbation, consolidation, and armoring. Computers & Geosciences 34 (10): 1263–1283.

    Article  Google Scholar 

  • Schuttelaars, H.M., and H.E. de Swart. 1996. An idealized long-term morphodynamic model of a tidal embayment. European Journal of Mechanics - B/Fluids 15: 55–80.

    MATH  Google Scholar 

  • Schuttelaars, H.M., and H.E. de Swart. 1999. Initial formation of channels and shoals in a short tidal embayment. Journal of Fluid Mechanics 386: 15–42.

    Article  MathSciNet  MATH  Google Scholar 

  • Schuttelaars, H.M., and H.E. de Swart. 2000. Multiple morphodynamic equilibria in tidal embayments. Journal of Geophysical Research 105: 24105–24118.

    Article  Google Scholar 

  • Seydel, R. 2010. Practical bifurcation and stability analysis. Interdisciplinary applied mathematics, vol. 5. New York: Springer.

    Google Scholar 

  • Soulsby, R. 1997. Dynamics of marine sands. London: Thomas Telfort.

    Google Scholar 

  • Stive, M.J.F., Z.B. Wang, M. Capobianco, P. Ruol, and M.C. Buijsman. 1998. Morphodynamics of a tidal lagoon and the adjacent coast. In Physics of estuaries and coastal seas, 397–407. Rotterdam: Balkema.

    Google Scholar 

  • SWAN team. 2006. SWAN, technical documentation, http://swanmodel.sourceforge.net/-online_doc-/swantech-/swantech.html. report, Delft University of Technology, Delft, The Netherlands.

  • Ter Brake, M. 2011. Tidal embayments: modelling and understanding their morphodynamics. PhD thesis, Delft University of Technology.

    Google Scholar 

  • Ter Brake, M.C., and H.M. Schuttelaars. 2010. Modeling equilibrium bed profiles of short tidal. Ocean Dynamics 60: 183–204.

    Google Scholar 

  • Ter Brake, M.C., and H.M. Schuttelaars. 2011. Channel and shoal development in a short tidal embayment; an idealized model study. Journal of Fluid Mechanics 677: 503–529.

    Article  MathSciNet  MATH  Google Scholar 

  • Todeschini, I., M. Toffolon, and M. Tubino. 2008. Long-term morphological evolution of funnel-shape tide-dominated estuaries. Journal of Geophysical Research 113 (5): C05005.

    Google Scholar 

  • Tolman, H.L. 1991. A third-generation model for wind waves on slowly varying, unsteady, and inhomogeneous depths and currents. Journal of Physical Oceanography 21: 782–797.

    Article  Google Scholar 

  • Uchiyama, Y., J.C. McWilliams, and A.F. Shchepetkin. 2010. Wave-current interaction in an oceanic circulation model with a vortex-force formalism: Application to the surf zone. Ocean Modelling 34 (1–2): 16–35.

    Article  Google Scholar 

  • Van de Kreeke, J. 1990. Can multiple tidal inlets be stable? Estuarine, Coastal and Shelf Science 30: 261–273.

    Article  Google Scholar 

  • Van de Kreeke, J., R.L. Brouwer, T.J. Zitman, and H.M. Schuttelaars. 2008. The effect of a topographic high on the morphodynamical stability of a two inlet bay system. Coastal Engineering 55: 319–332.

    Article  Google Scholar 

  • Van der Wegen, M., and J.A. Roelvink. 2008. Long-term morphodynamic evolution of a tidal embayment using a two-dimensional, process-based model. Journal of Geophysical Research: Oceans 113 (3): C03016.

    Google Scholar 

  • Van Leeuwen, S.M., and H.E. de Swart. 2001. The effect of advective processes on the morphodynamic stability of short tidal embayments. Physics and Chemistry of the Earth, Part B 26: 735–740.

    Article  Google Scholar 

  • Van Maanen, B., G. Coco, and K.R. Bryan. 2013. Modelling the effects of tidal range and initial bathymetry on the morphological evolution of tidal embayments. Geomorphology 191: 23–34.

    Article  Google Scholar 

  • Van Rijn, L.C. 1993. Principles of sediment transport in rivers, estuaries and coastal seas, 1993. Amsterdam: Aqua Publ.

    Google Scholar 

  • Van Rijn, L.C. 2007. Unified view of sediment transport by currents and waves, I: Initiation of motion, bed roughness, and bed-load transport. Journal of Hydraulic Engineering 133: 649–667.

    Article  Google Scholar 

  • Van Rijn, L.C. 2007. Unified view of sediment transport by currents and waves, II: Suspended transport. Journal of Hydraulic Engineering 133: 668–689.

    Article  Google Scholar 

  • Van Rijn, L.C., J.S. Ribberink, and J. van ver Werf and D.J.R. Walstra. 2013. Coastal sediment dynamics: Recent advances and future research needs. Journal of Hydraulic Research 51: 475–493.

    Google Scholar 

  • Werner, B.T., and T.M. Fink. 1993. Beach cusps as self-organized patterns. Science 260: 968–971.

    Article  Google Scholar 

  • Wijnberg, K.M., and J.H.J. Terwindt. 1995. Extracting decadal morphological behaviour from high-resolution, long-term bathymetric surveys along the Holland coast using eigenfunction analysis. Marine Geology 126: 301–330.

    Article  Google Scholar 

  • Wilson, K.C. 1966. Bed-load transport at high shear stress. Journal of the Hydraulics Division 92: 49–59.

    Article  Google Scholar 

  • Winterwerp, J.C. 1998. A simple model for turbulence induced flocculation of cohesive sediment. Journal of Hydraulic Research 36: 309–326.

    Article  Google Scholar 

  • Winterwerp, J.C. 2001. Stratification effects by cohesive and noncohesive sediment. Journal of Geophysical Research: Oceans 106 (C10): 22559–22574.

    Article  Google Scholar 

  • Winterwerp, J.C., and W.G.M. van Kesteren. 2004. Introduction to the physics of cohesive sediment dynamics in the marine environment, vol. 56. Elsevier.

    Google Scholar 

  • Xia, M., M. Mao, and Q. Niu. 2020. Implementation and comparison of the recent three-dimensional radiation stress theory and vortex-force formalism in an unstructured-grid coastal circulation model. Estuarine, Coastal and Shelf Science 240: 106771.

    Article  Google Scholar 

  • Xu, F., G. Coco, J. Tao, Z. Zhou, C. Zhang, S. Lanzoni, and A. D’Alpaos. 2019. On the morphodynamic equilibrium of a short tidal channel. Journal of Geophysical Research: Earth Surface 124 (2): 639–665.

    Article  Google Scholar 

  • Zimmerman, J.T.F. 1982. On the Lorentz linearization of a quadratically damped forced oscillator. Physics Letters 89A: 123–124.

    Article  Google Scholar 

  • Zimmerman, J.T.F. 1992. On the Lorentz linearization of a nonlinearly damped tidal Helmholtz oscillator. Proceedings of the Koninklijke Nederlandse Akademie van Wetenschappen 95: 127–145.

    Google Scholar 

  • Zyserman, J.A., and J. Fredsøe. 1994. Data analysis of bed concentration of suspended sediment. Journal of Hydraulic Engineering 120: 1021–1042.

    Article  Google Scholar 

Download references

Acknowledgements

We would like to thank EMODnet Bathymetry Consortium (2018): EMODnet Digital Bathymetry (DTM), http://doi.org/10.12770/18ff0d48-b203-4a65-94a9-5fd8b0ec35f6 for making available the bathymetric data used in Fig. 1.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to H. M. Schuttelaars .

Editor information

Editors and Affiliations

Appendix

Appendix

In this appendix, specific parameterisations will be chosen and the depth–averaged morphodynamic system of equations, given in Eqs. (10.16)–(10.20), will be scaled using characteristic dimensions of short tidal inlet systems. Using the non–dimensional equations, an asymptotic expansion of the physical variables is proposed and the leading order system of equations is derived.

As a first step, the bottom shear stresses \((\tau _x,\tau _y)^T\), the erosion function E and the deposition function D have to expressed in terms of relevant physical variables. Concerning the bed shear stresses, observations for turbulent flow conditions and dimension analysis suggest a quadratic dependency on the depth–averaged velocity. However, the gross features of the water motion are well captured by linearizing this quadratic dependency, resulting in a linearized friction law (Lorentz 1922; Zimmerman 1982, 1992), which is the approach taken here:

$$\begin{aligned} (\tau _x,\tau _y)^T = \rho r_\star \hat{\textbf{u}}. \end{aligned}$$
(10.38)

In this expression, \(r_\star \) is a friction parameter with units \(\text {m} \text {s}^{-1}\). The parameter \(r_\star \) is chosen such that the net dissipation of energy (averaged over the tidal cycle and embayment) due to the linearized shear stress (10.38) equals that of the bed shear stress based on the quadratic friction law. This implies that the parameter \(r_\star \) is proportional to the tidal current amplitude U and the bottom roughness.

The parameterisation of the friction terms \(\hat{\textbf{F}}^b\) in the depth–averaged momentum Eqs. (10.17) and (10.18), which read

$$\begin{aligned} \hat{\textbf{F}}^b = \frac{(\tau _x,\tau _y)^T}{\rho (H-h+\zeta )}, \end{aligned}$$

becomes unbounded if the water depth tends to zero, as observed near tidal flats. As our model is designed to give only a global description of the tidal flow in an embayment, the friction terms are regularized by increasing the water depth in the denominator of the friction term by a constant \(h_0\) (see Ter Brake and Schuttelaars 2010 for a detailed discussion, and the influence of this parameter on the morphodynamic equilibria).

The erosion function E and deposition function D are due to the pick–up and deposition of the sediment near the bottom. Motivated by field observations (Dyer and Soulsby 1988) and theoretical considerations, the sediment pick–up term E is taken proportional to some power of the absolute value of the difference between the actual bed shear stress and the critical shear stress for erosion. Taking this power to be equal to one (which provides a fair approximation and is computationally beneficial) and assuming the critical shear stress for erosion to be much smaller than the typical bed shear stress in a tidal inlet system, the sediment pick–up term is parameterised as:

$$\begin{aligned} E = \hat{\alpha } \left( \hat{u}^{2} + \hat{v}^2 \right) , \end{aligned}$$

where \(\hat{\alpha }\) is a constant which depends on the sediment characteristics. For fine sand (grain size \(2 \cdot 10^{-4} \, \text {m}\)) a typical value is \(\hat{\alpha } \sim 10^{-2} \text {kg} \, \text {m}^{-2} \, \text {s}^{-1}\).

The deposition function D is obtained by assuming an approximate balance between settling and vertical diffusion in the three dimensional concentration equation (10.1). Assuming the verical eddy diffusivity to be constant in space and time and that the sediment Peclet number \(w_s H/K_v \gg 1\), the resulting bottom concentration can be expressed in terms of the depth–integrated concentration, resulting in \(D = \gamma C\), with \(\gamma = K_v/w_s^2\).

Using these parameterisations, the morphodynamic system of equations is made non–dimensional by introducing characteristic scales for all physical variables. Focussing on basin–wide phenomena, the tidally averaged basin length L is used as a typical horizontal length scale, and the inverse of the angular frequency \(\sigma \) of the semidiurnal tide as a typical time scale. As a typical water depth the tidally and width-averaged water depth at the open boundary \(\overline{H}\) is used, and for the \(M_2\) amplitude of the free surface elevation the width–averaged water level amplitude at the open boundary \(\overline{A_{M_2}}\) is employed. The velocity scale follows from the continuity equation (10.16) by requiring a balance between the temporal change in the free surface elevation and the convergence of the water transport, resulting in \(U = \sigma \overline{A_{M_2}}L/\overline{H}\). The depth-integrated suspended sediment concentration is obtained by assuming an approximate balance between erosion E and deposition D. Substituting the resulting non–dimensional variables (indicated by an asterisk),

$$\begin{aligned}&(x,y)=L(x^*,y^*); \quad t={\sigma }^{-1}t^*; \quad \zeta =\overline{A_{M_2}}\zeta ^*; \quad (z,h) = \overline{H}(z^*,h^*); \nonumber \\&(u,v)=U(\hat{u}^*,\hat{v}^*); \quad C=\frac{\hat{\alpha } {U}^2}{\gamma } C^*, \quad \end{aligned}$$
(10.39)

into the morphodynamic equations results in the following non–dimensional system of equations:

$$\begin{aligned} \frac{\partial \zeta ^*}{\partial t^*} + \nabla ^* \cdot \left[ (1 - h^* + \varepsilon \zeta ^*) \boldsymbol{\hat{u}}^*\right]&= 0, \end{aligned}$$
(10.40)
$$\begin{aligned} \frac{\partial \hat{u}^*}{\partial t^*} + \varepsilon \boldsymbol{\hat{u}}^* \cdot \nabla ^* \hat{u}^* + f^* \hat{v}^* + \left( \frac{1}{\varLambda ^*}\right) ^2 \frac{\partial \zeta ^*}{\partial x^*}&= -\frac{r^* u^*}{(1-h^*+ \varepsilon \zeta ^*)}, \end{aligned}$$
(10.41)
$$\begin{aligned} \frac{\partial \hat{v}^*}{\partial t^*} + \varepsilon \boldsymbol{\hat{u}}^* \cdot \nabla ^* \hat{v}*^ - f^* \hat{u}^* + \left( \frac{1}{\varLambda ^*} \right) ^2 \frac{\partial \zeta ^*}{\partial y^*}&= -\frac{r^* \hat{v}^*}{(1-h^*+\varepsilon \zeta ^*)}, \end{aligned}$$
(10.42)
$$\begin{aligned} a^* \left\{ \frac{\partial C^*}{\partial t^*} + \varepsilon \nabla ^*\cdot \left( \boldsymbol{\hat{u}}^* C^* \right) - K_h^* \nabla ^{*2} C^* \right\}&= \left[ \left( \hat{u}^*\right) ^2 + \left( \hat{v}^*\right) ^2\right] - C^*, \end{aligned}$$
(10.43)
$$\begin{aligned} \frac{\partial h^*}{\partial t^*} - \lambda ^* \delta _b^*\nabla ^{*2} h^*&= \delta _s^* \left\{ C^* - \left[ \left( \hat{u}^*\right) ^2 + \left( \hat{v}^*\right) ^2\right] \right\} . \end{aligned}$$
(10.44)

For simplicity, the diffusive transport resulting from topographic variations is neglected in the concentration equation. Furthermore, the bedload transport has been significantly simplified (compare with Eq. (10.9)): only the contribution related to the topography is retained. Furthermore, this contribution is assumed to be isotropic in space (with the second–rank tensor \(\boldsymbol{\lambda }\) reduced to a scalar coefficient \(\lambda \)).

The non–dimensional numbers in Eqs. (10.40)–(10.43) are \(\varepsilon = \overline{A_{M_2}}/\overline{H} \equiv U/\sigma L\), which is the ratio of the mean \(M_2\) tidal amplitude at the seaward boundary and the mean water depth at this boundary, \(f^* =f/\sigma \) the non–dimensional Coriolis parameter, \(1/\varLambda ^* = \sigma ^2 L^2/g H\) the square of the product of the frictionless tidal wavenumber and the typical lengthscale L, \(r^*= r_*/\sigma H\) the non–dimensional friction parameter, \(K_h^* = K_h/\sigma L^2\) the non–dimensional horizontal eddy viscosity coefficient, and \(a^*=\sigma /\gamma \) the ratio of the deposition time scale and the tidal time scale \(1/\sigma \).

In the bed evolution equation (10.44), apart from the non–dimensional parameter \(\lambda ^* = H \lambda /L\), two other non–dimensional parameters appear. The first one is denoted by \(\delta _s^* = \hat{\alpha } U^2/\rho _s(1-p) \sigma H\) which is the ratio of the tidal time scale and the typical time scale \(T_{\text {s}}\) at which the bed changes due to spatial gradients in suspended sediment transport. The second non–dimensional parameter reads \(\delta _b^* = \hat{s}/\sigma H L\), with \(\hat{s} = (1-p) \rho _{\text {s}} \hat{q}_{\text {b}} \sqrt{g'd_{\text {s}}^3} (\theta - \theta _c)^{b_{\text {bl}}} \). This parameter \(\delta _b^*\) is the ratio of the tidal time scale and the time scale \(T_{\text {b}}\) of the bed evolution due to spatial gradients in bedload transport. In the systems under study the time scale \(T_{\text {b}}\) is much larger than \(T_{\text {s}}\), hence the bedload transport is much smaller than the suspended load transport. However, bedload effects cannot be neglected because they are necessary to suppress the growth of small–scale bedforms (Schuttelaars and De Swart 1999; Ter Brake and Schuttelaars 2011).

The fact that both morphodynamic time scales \(T_{\text {s}}\) and \(T_{\text {b}}\) are much larger than the hydrodynamic time scale \(\sigma ^{-1}\) implies that the bed level h can be regarded as slowly varying compared to the other physical variables. Consequently the method of time–averaging can be used: the sediment and the bedrock layer thickness may be considered stationary on the comparatively short tidal time scale and its evolution is only determined by spatial gradients in the sediment transports averaged over a tidal cycle. The mathematical foundations of this approach are discussed in Sanders and Verhulst (1985) and Krol (1991).

This results in the following non–dimensional bottom evolution equation:

$$\begin{aligned} h^*_{\tau ^*} = \bigl < C^* - \left[ \left( \hat{u}^*\right) ^2 + \left( \hat{v}^*\right) ^2 \right] \bigl > + \lambda ^* \langle \nabla ^{*2} h^* \rangle , \end{aligned}$$
(10.45)

with \(\langle \cdot \rangle \) denoting averaging over the tidal time scale, \(\tau ^* = \delta _{s}^* t^*\) denotes the slow time coordinate, and \(\lambda ^* = \kappa ^* \delta _b/\delta _s\). Using Eq. (10.43), this expression can be rewritten as

$$\begin{aligned} h^*_{\tau ^*} = - \nabla ^* \cdot \left( \boldsymbol{q}^*_{\text {diff}} + \boldsymbol{q}^*_{\text {adv}} + \boldsymbol{q}^*_{\text {bl}} \right) , \end{aligned}$$
(10.46)

with the sediment transport contributions defined as

$$\begin{aligned} \boldsymbol{q}^*_{\text {diff}}&= -a^* K_h^* \langle \nabla ^* C^* \rangle , \nonumber \\ \boldsymbol{q}^*_{\text {adv}}&= a^* \varepsilon \langle \boldsymbol{\hat{u}}^* C^* \rangle , \\ \boldsymbol{q}^*_{\text {bed}}&= -\lambda ^* \langle \nabla ^* h^* \rangle . \nonumber \end{aligned}$$
(10.47)

Considering the rectangular basin geometry, described in Sect. 10.5, the non–dimensional boundary conditions at the sidewalls are given by

$$\begin{aligned} (1-h^* +\varepsilon \zeta ^* ) \hat{v}^* = 0, \qquad K_h^* \frac{\partial C^*}{\partial y^*} = 0, \qquad \lambda ^* \frac{\partial h^*}{\partial y^*} = 0 \qquad \text {at }y^*=0, y^*=\frac{B}{L}, \end{aligned}$$

where the first condition requires the normal transport of water through the wall to vanish, and the second and third conditions require that both the suspended sediment transport and the bedload transport through the sidewalls vanish.

To obtain boundary conditions at the landward boundaries, define the location where the water depth vanishes as \(X^*(t)\). At this moving boundary, the velocity is given by \(\hat{u}^* = dX^*/dt^*\) and the tidally averaged (diffusive) sediment transport is assumed to vanish. The tidally averaged non–dimensional embayment length is 1, with deviations from this averaged length of \(O(\varepsilon )\). Introducing this expansion in the condition of vanishing water depth at \(x^*=X^*\), \(1-h(X^*)+ \varepsilon \zeta (X^*,t^*) = 1\), using a Taylor expansion, it follows that in leading order \(h^* = 1\) at \(x^* = 1\). After substitution of this condition in the continuity equation, the boundary condition at the end of the embayment can be reformulated as a boundary condition at \(x^* = 1\) and is given by \(u^*_{x^*}\) is finite at \(x^*=1\) (Van Leeuwen and De Swart 2001; Ter Brake and Schuttelaars 2011). Apart from this condition, the following boundary conditions are imposed:

$$\begin{aligned} K_h^* \frac{\partial C^*}{\partial x^*} + \lambda ^* \frac{\partial h^*}{\partial x^*} = 0, \qquad h^*=1 \qquad \text {at }x^* = 1. \end{aligned}$$

At the seaward boundary the water motion is forced by a single tidal constituent, the bed level is kept fixed, meaning that erosion is assumed to balance deposition:

$$\begin{aligned} \zeta ^* = \left( {A_{M_2}}/{\overline{A_{M_2}}}\right) \cos (t^*), \qquad \bigl < C^* - \left[ \left( \hat{u}^*\right) ^2 + \left( \hat{v}^*\right) ^2 \right] \bigl > = 0, \qquad \bar{h}^* = 0 \qquad \text {at }x^* = 0. \end{aligned}$$

The latter expression requires that the bed elevation \(\bar{h}^*\), which is \(h^*\) averaged over the seaward boundary, is equal to zero. For a more detailed discussion of these boundary conditions, see Schuttelaars and De Swart (1999), Ter Brake and Schuttelaars (2011), Boelens et al. (2021).

Equations (10.40)–(10.43) and (10.46) can be solved using an asymptotic expansion in the small parameter \(\varepsilon \), reflecting comparatively small deviations from the O(1) values of primary parameters, see Schuttelaars and De Swart (2000), Ter Brake and Schuttelaars (2011). Assuming that all parameters are O(1), the focus will be on the O(1) equations, neglecting terms of \(O(\varepsilon )\) and smaller. Furthermore, for the system under consideration \(1/\varLambda ^* \gg 1\) which allows for a further simplification of the momentum equations (10.41) and (10.42):

$$\begin{aligned} \frac{\partial \zeta ^*}{\partial x^*} = \frac{\partial \zeta ^*}{\partial y^*} = 0, \end{aligned}$$

which states that the variations in the free surface elevations are spatially uniform. Information on the zeroth–order velocity field is obtained from the \(O((\varLambda ^*)^2)\) momentum balance: elimination of the pressure terms results in a vorticity equation, which reads (neglecting earth rotation effects):

$$\begin{aligned} \frac{\partial ^2 \hat{v}^*}{\partial x^* \, \partial t^*} - \frac{\partial ^2 \hat{u}^*}{\partial y^* \, \partial t^*} = -\frac{\partial }{\partial x^*} \left( \frac{r^* \hat{v}^*}{1-h^*+h_0^*} \right) + \frac{\partial }{\partial y^*} \left( \frac{r^* \hat{u}^*}{1-h^*+h_0^*} \right) . \end{aligned}$$
(10.48)

Converting the system of non–dimensional equations and the associated boundary conditions back to dimensional ones resulting in the equations of Sect. 10.5.3.

Rights and permissions

Reprints and permissions

Copyright information

© 2022 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Schuttelaars, H.M., Zitman, T.J. (2022). Morphodynamic Modelling in Marine Environments: Model Formulation and Solution Techniques. In: Schuttelaars, H., Heemink, A., Deleersnijder, E. (eds) The Mathematics of Marine Modelling. Mathematics of Planet Earth, vol 9. Springer, Cham. https://doi.org/10.1007/978-3-031-09559-7_10

Download citation

Publish with us

Policies and ethics