Skip to main content
Log in

Landslide Tsunami Hazard Along the Upper US East Coast: Effects of Slide Deformation, Bottom Friction, and Frequency Dispersion

  • Published:
Pure and Applied Geophysics Aims and scope Submit manuscript

Abstract

Numerical simulations of Submarine Mass Failures (SMFs) are performed along the upper US East Coast to assess the effect of slide deformation on predicted tsunami hazard. Tsunami generation is simulated using the three-dimensional non-hydrostatic model NHWAVE. For rigid slumps, the geometry and law of motion are specified as bottom boundary conditions. Deforming slide motion is modeled using a depth-integrated bottom layer of dense Newtonian fluid, fully coupled to the overlying fluid motion. Once the SMFs are no-longer tsunamigenic, tsunami propagation simulations are performed using the Boussinesq wave model FUNWAVE-TVD, using nested grids of increasingly fine resolution towards shore and employing a one-way coupling methodology. Probable maximum tsunamis are simulated for Currituck SMF proxies sited in four areas of the shelf break slope that have enough sediment accumulation to cause large failures. Deforming slides have a slightly larger initial acceleration, but still generate a smaller tsunami than rigid slumps due to their spreading and thinning out during motion, which gradually makes them less tsunamigenic. Comparing the maximum envelope of surface elevations along a 5 m isobath, consistent with earlier work, the bathymetry of the wide shelf is found to strongly control the spatial distribution of tsunami inundation. Overall, tsunamis caused by rigid slumps are worst case scenarios, providing up to 50% more inundation than for deforming slides having a moderate level of viscosity set in the upper range of debris flows. Tsunamis from both types of SMFs are able to cause water withdrawal to the 5 m isobath or deeper. Bottom friction effects are assessed by performing some of the simulations using two different Manning coefficients, one 50% larger than the other. With increased bottom friction, the largest tsunami inundations at the coast are reduced by up to 15%. Selected simulations are rerun by turning off dispersion in the model, which leads to moderate changes in maximum surface elevations nearshore (− 10 to + 5% changes), but to more significant effects in the far field (− 40 to 80% changes). Onshore, dispersion causes the appearance of short period undular bores that eventually break nearshore without significantly affecting inundation at the coast. However, these bores increase wave-induced maximum flow velocity and impulse forces, the latter by up to 40%, which may affect the design of coastal structures.

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
Fig. 18
Fig. 19
Fig. 20
Fig. 21
Fig. 22

Similar content being viewed by others

References

  • Abadie, S., Harris, J. C., Grilli, S. T., & Fabre, R. (2012). Numerical modeling of tsunami waves generated by the flank collapse of the Cumbre Vieja Volcano (La Palma, Canary Islands): Tsunami source and near field effects. Journal of Geophysical Research, 117, C05030. https://doi.org/10.1029/2011JC007646.

    Article  Google Scholar 

  • Amante, C, & Eakins, B. W. (2009). ETOPO1 1 arc-minute global relief model: Procedures, data sources and analysis. NOAA Technical Memorandum NESDIS NGDC-24. National Geophysical Data Center, NOAA. https://doi.org/10.7289/V5C8276M (March 15, 2017).

  • Booth, J. S., O’Leary, D. W., Popenoe, P., & Danforth, W. W. (1993). US Atlantic Continental Slope Landslides: Their distribution, general attributes, and implications. In: Schwab, W. C., Lee, H. J., & Twichell, D. C. (Eds.) Submarine landslides: Selected studies in the US exclusive economic zone. US Geological Survey Bulletin, 2002, pp. 14–22.

  • ten Brink, U. S., Barkan, R., Andrews, B. D., & Chaytor, J. D. (2009b). Size distributions and failure initiation of submarine and subaerial landslides. Earth and Planetary Science Letters, 287, 31–42.

    Article  Google Scholar 

  • ten Brink, U. S., Chaytor, J. D., Geist, E. L., Brothers, D. S., & Andrews, B. D. (2014). Assessment of tsunami hazard to the US Atlantic margin. Marine Geology, 353, 31–54.

    Article  Google Scholar 

  • ten Brink, U. S., Lee, H. J., Geist, E. L., & Twichell, D. (2009a). Assessment of tsunami hazard to the US East Coast using relationships between submarine landslides and earthquakes. Marine Geology, 264, 65–73.

    Article  Google Scholar 

  • ten Brink, U., Twichell, D., Geist, E., Chaytor, J., Locat, J., Lee, H., Buczkowski, B., Barkan, R., Solow, A., Andrews, B., Parsons, T., Lynett, P., Lin, J., & Sansoucy, M. (2008). Evaluation of tsunami sources with the potential to impact the US Atlantic and Gulf coasts. USGS Administrative report to the US Nuclear Regulatory Commission, p. 300.

  • Chaytor, J. D., ten Brink, U. S., Solow, A. R., & Andrews, B. D. (2009). Size distribution of submarine landslides along the US Atlantic margin. Mar Geol, 264(12), 16–27.

    Article  Google Scholar 

  • Chock, G. Y. (2016). Design for tsunami loads and effects in the ASCE 7–16 standard. Journal of Structural Engineering, 142(11), 04016093. https://doi.org/10.1061/(ASCE)ST.1943-541X.0001565.

    Article  Google Scholar 

  • ECMAP. (2017). NTHMP tsunami inundation maps for the US East Coast. https://www.udel.edu/kirby/nthmp.html.

  • Eggeling, T. (2012). Analysis of earthquake triggered submarine landslides at four locations along the US east coast. Masters Thesis. University of Rhode Island.

  • Enet, F., & Grilli, S. T. (2007). Experimental study of tsunami generation by three-dimensional rigid underwater landslides. Journal of Waterway Port Coastal and Ocean Engineering, 133(6), 442–454. https://doi.org/10.1061/(ASCE)0733-950X(2007).

    Article  Google Scholar 

  • Fine, I. V., Rabinovich, A. B., Bornhold, B. D., Thomson, R., & Kulikov, E. A. (2005). The Grand Banks landslide-generated tsunami of November 18, 1929: Preliminary analysis and numerical modeling. Marine Geology, 215, 45–57.

    Article  Google Scholar 

  • Geist, E. L., & Lynett, P. J. (2014). Source processes for the probabilistic assessment of tsunami hazards. Oceanography, 27(2), 86–93.

    Article  Google Scholar 

  • Geist, E., Lynett, P., & Chaytor, J. (2009). Hydrodynamic modeling of tsunamis from the Currituck landslide. Marine Geology, 264, 41–52. https://doi.org/10.1016/j.margeo.2008.09.005.

    Article  Google Scholar 

  • Glimsdal, S., Pedersen, G. K., Harbitz, C. B., & Løvholt, F. (2013). Dispersion of tsunamis: Does it really matter? Natural Hazards and Earth System Sciences, 13, 1507–1526. https://doi.org/10.5194/nhess-13-1507-2013.

    Article  Google Scholar 

  • González, F. I., Geist, E. L., Jaffe, B., Kânoglu, U., Mofjeld, H., Synolakis, C. E., et al. (2009). Probabilistic tsunami hazard assessment at seaside, Oregon, for near- and far-field seismic sources. Journal of Geophysical Research Oceans, 114, C11.

    Google Scholar 

  • Grezio, A., Babeyko, A., Baptista, M. A., Behrens, J., Costa, A., Davies, G., et al. (2017). Probabilistic tsunami hazard analysis: Multiple sources and global applications. Reviews of Geophysics, 55(4), 1158–1198. https://doi.org/10.1002/2017RG000579.

    Article  Google Scholar 

  • Grilli, S. T., & Watts, P. (1999). Modeling of waves generated by a moving submerged body. Applications to underwater landslides. Engineering Analysis with Boundary Elements, 23, 645–656.

    Article  Google Scholar 

  • Grilli, S. T., & Watts, P. (2005). Tsunami generation by submarine mass failure. I: Modeling, experimental validation, and sensitivity analyses. Journal of Waterway Port Coastal and Ocean Engineering, 131(6), 283–297. https://doi.org/10.1061/(ASCE)0733-950X(2005).

    Article  Google Scholar 

  • Grilli, S. T., Dubosq, S., Pophet, N., Pérignon, Y., Kirby, J. T., & Shi, F. (2010). Numerical simulation and first-order hazard analysis of large co-seismic tsunamis generated in the Puerto Rico trench: Near-field impact on the North shore of Puerto Rico and far-field impact on the US East Coast. Natural Hazards and Earth System Sciences, 10, 2109–2125. https://doi.org/10.5194/nhess-2109-2010.

    Article  Google Scholar 

  • Grilli, S. T., Grilli, A. R., Tehranirad, B., & Kirby, J. T. (2017a). Modeling tsunami sources and their propagation in the Atlantic Ocean for coastal tsunami hazard assessment and inundation mapping along the US East Coast. In Proceedings of coastal structures and solutions to coastal disasters 2015: Tsunamis (pp. 1–12). Boston: American Soc. Civil Eng, 2015.

  • Grilli, S. T., Harris, J. C., Tajalibakhsh, T., Masterlark, T. L., Kyriakopoulos, C., Kirby, J. T., et al. (2013b). Numerical simulation of the 2011 Tohoku tsunami based on a new transient FEM co-seismic source: Comparison to far- and near-field observations. Pure and Applied Geophysics, 170, 1333–1359. https://doi.org/10.1007/s00024-012-0528-y.

    Article  Google Scholar 

  • Grilli, S. T., Harris, J. C., Tappin, D. R., Masterlark, T., Kirby, J. T., Shi, F., & Ma, G. (2013a). Numerical modeling of coastal tsunami dissipation and impact. In P. Lynett, & J. Mc Kee Smith (Eds.), Proceedings of 33rd international coastal engineering conference (ICCE12, Santander, Spain, July, 2012) (p. 12). World Scientific Publishing Co. Pte. Ltd.

  • Grilli, S. T., O’Reilly, C., Harris, J. C., Tajalli Bakhsh, T., Tehranirad, B., Banihashemi, S., et al. (2015). Modeling of SMF tsunami hazard along the upper US East Coast: Detailed impact around Ocean City, MD. Natural Hazards, 76(2), 705–746. https://doi.org/10.1007/s11069-014-1522-8.

    Article  Google Scholar 

  • Grilli, S. T., Shelby, M., Kimmoun, O., Dupont, G., Nicolsky, D., Ma, G., et al. (2017b). Modeling coastal tsunami hazard from submarine mass failures: Effect of slide rheology, experimental validation, and case studies off the US East Coast. Natural Hazards, 86(1), 353–391. https://doi.org/10.1007/s11069-016-2692-3.

    Article  Google Scholar 

  • Grilli, S. T., Taylor, O.-D. S., Baxter, C. D. P., & Maretzki, S. (2009). Probabilistic approach for determining submarine landslide tsunami hazard along the upper East Coast of the United States. Marine Geology, 264(1–2), 74–97.

    Article  Google Scholar 

  • Grilli, S. T., Vogelmann, S., & Watts, P. (2002). Development of a 3D Numerical Wave Tank for modeling tsunami generation by underwater landslides. Engineering Analysis with Boundary Elements, 26(4), 301–313.

    Article  Google Scholar 

  • Grothe, P. R., Taylor, L. A., Eakins, B. W., Warnken, R. R., Carignan, K. S., Lim, E., Caldwell, R.J., & Friday, D. Z. (2010). Digital elevation model of Ocean City, Maryland: Procedures, data sources and analysis. NOAA Technical Memorandum NESDIS NGDC-37. National Geophysical Data Center, NOAA.

  • Heidarzadeh, M., & Kijko, Andrzej. (2011). A probabilistic tsunami hazard assessment for the Makran subduction zone at the northwestern Indian Ocean. Natural Hazard, 56(3), 577–593.

    Article  Google Scholar 

  • Hill, J. C., Brothers, D. S., Craig, B. K., ten Brink, Uri S., Chaytor, J. D., & Flores, C. H. (2017). Geologic controls on submarine slope failure along the central US Atlantic margin: Insights from the Currituck Slide Complex. Marine Geology, 385, 114–130. https://doi.org/10.1016/j.margeo.2016.10.007.

    Article  Google Scholar 

  • Horspool, N., Pranantyo, I., Griffin, J., Latief, H., Natawidjaja, D. H., Kongko, W., et al. (2014). A probabilistic tsunami hazard assessment for Indonesia. Natural Hazards and Earth System Sciences, 14(11), 3105–3122.

    Article  Google Scholar 

  • Jakeman, J. D., Nielsen, O. M., van Putten, K., Mleczko, R., Burbidge, D., & Horspool, N. (2010). Towards spatially distributed quantitative assessment of tsunami inundation models. Ocean Dynamics, 60(5), 11151138. https://doi.org/10.1007/s10236-010-0312-4.

    Article  Google Scholar 

  • Kaiser, G., Scheele, L., Kortenhaus, A., Løvholt, F., Römer, H., & Leschka, S. (2011). The influence of land cover roughness on the results of high resolution tsunami inundation modeling. Natural Hazards and Earth System Sciences, 11, 2521–2540. https://doi.org/10.5194/nhess-11-2521-2011.

    Article  Google Scholar 

  • Kirby, J. T., Shi, F., Nicolsky, D., & Misra, S. (2016). The 27 April 1975 Kitimat, British Columbia submarine landslide tsunami: A comparison of modeling approaches. Landslides, 13(6), 1421–1434. https://doi.org/10.1007/s10346-016-0682-x.

    Article  Google Scholar 

  • Kirby, J. T., Shi, F., Tehranirad, B., Harris, J. C., & Grilli, S. T. (2013). Dispersive tsunami waves in the ocean: Model equations and sensitivity to dispersion and Coriolis effects. Ocean Modelling, 62, 39–55. https://doi.org/10.1016/j.ocemod.2012.11.009.

    Article  Google Scholar 

  • Krauss, T, (2011). Probabilistic tsunami hazard assessment for the United States East Coast. Masters Thesis, University of Rhode Island. http://chinacat.coastal.udel.edu/nthmp/krause-ms-uri11.pdf.

  • Løvholt, F., Bondevik, S., Laberg, J. S., Kim, J., & Boylan, N. (2017). Some giant submarine landslides do not produce large tsunamis. Geophysical Research Letters, 44(16), 8463–8472. https://doi.org/10.1002/2017GL074062.

    Article  Google Scholar 

  • Løvholt, F., Pedersen, G., & Gisler, G. (2008). Oceanic propagation of a potential tsunami from the La Palma Island. Journal of Geophysical Research, 113, C09026.

    Article  Google Scholar 

  • Løvholt, F., Pedersen, G., Harbitz, C. B., Glimsdal, S., & Kim, J. (2015). On the characteristics of landslide tsunamis. Philosophical Transactions of the Royal Society A, 373, 20140376. https://doi.org/10.1098/rsta.2014.0376.

    Article  Google Scholar 

  • Løvholt, F., Schulten, I., Mosher, D., Harbitz, C., & Kraste, S. (2018). Modelling the 1929 Grand Banks slump and landslide tsunami. In D. G. Lintern, D. C. Mosher, L. G. Moscardelli, P. T. Bobrowsky, C. Campbell, J. D. Chaytor, J. J. Clague, A. Georgiopoulou, P. Lajeunesse, A. Normandeau, D. J. W. Piper, M. Scherwath, C. Stacey, & D. Turmel (Eds.), Subaqueous mass movements (p. 477). London: Geological Society London Special Publications. https://doi.org/10.1144/SP477.28.

    Google Scholar 

  • Locat, J., Lee, H., ten Brink, U. S., Twitchell, D., Geist, E., & Sansoucy, M. (2009). Geomorphology, stability and mobility of the Currituck slide. Marine Geology, 264, 28–40.

    Article  Google Scholar 

  • Lynett, P., et al. (2017). Inter-model analysis of tsunami-induced coastal currents. Ocean Modeling, 114, 14–32. https://doi.org/10.1016/j.ocemod.2017.04.003.

    Article  Google Scholar 

  • Ma, G., Shi, F., & Kirby, J. T. (2012). Shock-capturing non-hydrostatic model for fully dispersive surface wave processes. Ocean Modellimg, 43–44, 22–35.

    Article  Google Scholar 

  • Madsen, P. A., Fuhrman, D. R., & Schaffer, H. A. (2008). On the solitary wave paradigm for tsunamis. Journal of Geophysical Research, 113(C12012), 22.

    Google Scholar 

  • Matsuyama, M., Ikeno, M., Sakakiyama, T., & Takeda, T. (2007). A study of tsunami wave fission in an undistorted experiment. Tsunami and Its Hazards in the Indian and Pacific Oceans. Pure and Applied Geophysics, 164, 617–631.

    Article  Google Scholar 

  • McMaster, R. L. (1960). Sediments of Narragansett Bay system and Rhode Island Sound, Rhode Island. Journal of Sedimentary Research, 30(2), 249–274.

    Google Scholar 

  • National Geophysical Data Center. (1999a). US coastal relief model—Northeast Atlantic. National Geophysical Data Center, NOAA. https://doi.org/10.7289/V5MS3QNZ (March 15, 2017)

  • National Geophysical Data Center. (1999b). US coastal relief model—Southeast Atlantic. National Geophysical Data Center, NOAA. https://doi.org/10.7289/V53R0QR5 (March 15, 2017).

  • National Centers for Environmental Information. (2014). Tiled coastal elevation models. NOAA National Centers for Environmental Information. Dataset. https://doi.org/10.7289/V5P55KQC.

  • Piper, D. J. W., Cochonat, P., & Morrison, M. L. (1999). The sequence of events around the epicentre of the 1929 Grand Banks earthquake: Initiation of the debris flows and turbidity current inferred from side scan sonar. Sedimentology, 46, 79–97.

    Article  Google Scholar 

  • Shelby, M., Grilli, S. T., & Grilli, A. R. (2016). Tsunami hazard assessment in the Hudson River Estuary based on dynamic tsunami tide simulations. Pure and Applied Geophysics, 173(12), 3999–4037. https://doi.org/10.1007/s00024-016-1315-y.

    Article  Google Scholar 

  • Shi, F., Kirby, J. T., Harris, J. C., Geiman, J. D., & Grilli, S. T. (2012). A high-order adaptive time-stepping TVD solver for Boussinesq modeling of breaking waves and coastal inundation. Ocean Modelling, 43–44, 36–51. https://doi.org/10.1016/j.ocemod.2011.12.004.

    Article  Google Scholar 

  • SLIDE. (2017). Slope Stability Analysis Model. RockScience Inc. https://www.rocscience.com/rocscience/products/slide.

  • Tappin, D. R., Grilli, S. T., Harris, J. C., Geller, R. J., Masterlark, T., Kirby, J. T., et al. (2014). Did a submarine landslide contribute to the 2011 Tohoku tsunami? Marine Geology, 357, 344–361. https://doi.org/10.1016/j.margeo.2014.09.043.

    Article  Google Scholar 

  • Tehranirad, B., Harris, J. C., Grilli, A. R., Grilli, S. T., Abadie, S., Kirby, J. T., et al. (2015). Far-field tsunami hazard in the north Atlantic basin from large scale flank collapses of the Cumbre Vieja volcano, La Palma. Pure and Applied Geophysics, 172(12), 3589–3616. https://doi.org/10.1007/s00024-015-1135-5.

    Article  Google Scholar 

  • Tehranirad, B., Kirby, J. T., Grilli, S. T., Grilli, A. R, & SHi, F. (2017). Continental shelf bathymetry controls the spatial distribution of tsunami hazard for the US East Coast. Geophysical Research Letters (in preparation).

  • Towns, J., Cockerill, T., Dahan, M., Foster, I., Gaither, K., Grimshaw, A., et al. (2014). XSEDE: Accelerating scientific discovery. Computing in Science and Engineering, 16(5), 62–74. https://doi.org/10.1109/MCSE.2014.80.

    Article  Google Scholar 

  • Twichell, D. C., McClennen, C. E., & Butman, B. (1981). Morphology and processes associated with the accumulation of the fine-grained sediment deposit on the southern New England shelf. Journal of Sedimentary Research, 51(1), 269–280.

    Google Scholar 

  • USGS. (2002). US Geological Survey “2002 interactive deaggregations”. https://geohazards.usgs.gov/deaggint/2002/. Accessed 2012.

  • Watts, P. & Grilli, S. T. (2003). Underwater landslide shape, motion, deformation, and tsunami generation. In: Proceedings of 13th ISOPE, Honolulu (pp. 364–371). http://personal.egr.uri.edu/grilli/3055p364.pdf.

  • Watts, P., Grilli, S. T., Tappin, D. R., & Fryer, G. (2005). Tsunami generation by submarine mass failure, II: Predictive equations and case studies. Journal of Waterway Port Coastal and Ocean Engineering, 131(6), 298–310.

    Article  Google Scholar 

  • Yavari-Ramshe, S., & Ataie-Ashtiani, B. (2016). Numerical simulation of subaerial and submarine landslide generated tsunami waves—recent advances and future challenges. Landslides, 13(6), 1325–1368. https://doi.org/10.1007/s10346-016-0734-2.

    Article  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Tsunami Hazard Mitigation Program (NTHMP), NOAA, through Grants NA-15-NWS4670029 and NA-16-NWS4670034 to the University of Delaware (with subaward to the University of Rhode Island). Additional support at the University of Rhode Island and the University of Delaware came from Grants CMMI-15-35568 and CMMI-15-37568 from the Engineering for Natural Hazards Program, National Science Foundation, respectively. Numerical simulations reported in this work used HPC resources, as part of the Extreme Science and Engineering Discovery Environment (XSEDE) (project BCS-170006), which is supported by the National Science Foundation (NSF) Grant number ACI-1548562. FUNWAVE-TVD is open source software, available at http://github.com/fengyanshi/FUNWAVE-TVD/. NHWAVE is open source software, available at http://github.com/jimkirby/nhwave/. Finally, the authors acknowledge anonymous reviewers for their thorough and constructive reviews of this work.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Stephan T. Grilli.

Electronic supplementary material

Appendix: SMF Geometry and Slump Law of Motion

Appendix: SMF Geometry and Slump Law of Motion

For rigid slumps, kinematics is specified based on the analytical laws developed by Grilli and Watts (1999, 2005); Watts et al. (2005). Additionally, as in Enet and Grilli (2007), the SMF geometry is idealized as having a “Quasi-Gaussian” shape (below seafloor) of elevation \(\zeta (x,y)\), whose steepness is controlled by a shape parameter \(\varepsilon\) (here \(\varepsilon = 0.717\)), and elliptical footprint of downslope length b, width w, and maximum thickness T defined as (Fig. 23)

$$\begin{aligned} \zeta (x,y) = \frac{T}{1-\varepsilon } \text {max}\{0, \text {sec}h(k_{\text {b}} \xi )\,\text {sec}h(k_{\text {w}} \chi ) - \varepsilon \}, \end{aligned}$$
(1)

where (\(\xi , \chi\)) are the local downslope and spanwise horizontal coordinates, rotated in the direction of SMF motion \(\theta\), and \(k_{\text {b}}=2C/b\), \(k_{\text {w}}=2C/w\), with \(C = a{\text{cos}} h(1/\varepsilon )\). With this geometry and parameters, the SMF volume is given by

$$\begin{aligned} V_{\text {s}} = b w T \frac{I_2}{C^2}\bigg (\frac{\frac{I_1}{I_2} - \varepsilon }{1-\varepsilon }\bigg )\quad \mathrm{~with~} \quad I_{1;2} = \int _0^C f(\mu )\,\text {d}\mu ;g(\mu )\,\text {d}\mu \end{aligned}$$
(2)

and

$$\begin{aligned} f(\mu ) = \text {sec}h\mu \,{a \tan }({\text{sin}}h{g(\mu )})\mathrm{~,~}g(\mu )=a{\text{cos}} h\bigg (\frac{\text {sec}h\mu }{\varepsilon }\bigg ). \end{aligned}$$
(3)

[Note that Eqs. 2 and 3 have been corrected and are different from those reported in earlier papers (e.g., Enet and Grilli 2007; Grilli et al. 2015), which resulted from a mistake in the volume calculation.] For the specified \(\varepsilon\), we find, \(C=0.8616\), \(I_1=0.4804\), \(I_2=0.5672\), and \(V_{\text {s}} = 0.3508\, b w T\).

Earlier modeling work (Locat et al. 2009) indicates that, during its tsunamigenic period of motion, the Currituck SMF achieved a relatively small maximum displacement (runout) \(S_\mathbf{f } < b\) in its main direction of motion down the slope, over an unknown time of motion \(t_\mathbf{f }\). The combination of rigid block SMF and small displacement parallel to the slope supports modeling the SMF kinematics as a rigid slump or a deforming slide with moderate deformation achieving a similar runout over the same time. In either cases, one can assume a constant basal friction (i.e., slide-to-substrate friction) and negligible hydrodynamic drag (Grilli and Watts 2005). This type of rigid-body motion kinematics was investigated in earlier work (see above-listed references), leading for the slump to a pendulum-like center of mass motion S(t) parallel to the local mean slope of angle \(\alpha\). Here, this simple law of motion for rigid slumps is used, which reads

$$\begin{aligned} S(t) = S_0 \bigg (1-\cos \frac{t}{t_0}\bigg )\quad \mathrm{~for~}\quad 0 \le t \le t_\mathbf{f } \end{aligned}$$
(4)

with \(S_0=S_\mathbf{f }/2\) and \(t_0 = t_\mathbf{f }/\pi\), and \(S=S_\mathbf{f }\) for \(t>t_\mathbf{f }\) (assuming SMF triggering occurs at \(t=0\)).

Fig. 23
figure 23

Geometric parameterization of a SMF initially centered at (\(x_0, y_0\)) moving in direction \(\xi\), with an azimuth angle \(\theta\) from North and center of mass motion S(t) measured parallel to the mean local slope of angle \(\alpha\); (xy) denote the longitudinal and latitudinal horizontal directions, respectively

At \(t=0\), the SMF elevation is specified below the current seafloor bathymetry \(h_0(x,y)\). Given the SMF initial center of mass location (\(x_0, y_0\)) in global axes (xy) (i.e., coordinates of the center of the elliptical footprint) and azimuth angle of SMF motion \(\theta\), the coordinate transformation to the local SMF slope-parallel coordinate system (\(\xi , \chi\)) (Fig. 23) is defined as

$$\begin{aligned} \xi= & (x-x_o)\cos \theta - (y -y_0) {\text{sin}} \theta - S(t) \cos \alpha \nonumber \\ \chi= & (x-x_o){\text{sin}} \theta + (y -y_0) \cos \theta \end{aligned}$$
(5)

with S(t) given by Eq. (4).

Then, assuming \(\sin \alpha \simeq 0\) for small slopes, the instantaneous seafloor depth above the SMF is given by

$$\begin{aligned} h(x,y,t) = h_0(x,y) + \zeta \{\xi (x,y,t), \chi (x,y,t)\} - \zeta \{\xi (x,y,0), \chi (x,y,0)\} \end{aligned}$$
(6)

with \(\Delta h = h - h_0\). The seafloor motion described by Eq. (6) is similar to a translation parallel to the average slope of part of the seabed, over the actual bathymetry. The vertical seafloor velocity (used in NHWAVE as a bottom boundary condition) is computed as

$$\begin{aligned} \frac{\text {d} h}{\text {d} t}(x,y,t) = \frac{\text {d} \zeta }{\text {d} t}\{\xi (x,y,S(t)), \chi (x,y,t)\} \end{aligned}$$
(7)

which can be easily derived from Eqs. (1) to (6) as

$$\begin{aligned} \frac{\text {d} h}{\text {d} t}(x,y,t) = k_{\text {b}} \cos \alpha \bigg (\zeta + \frac{\varepsilon T}{1-\varepsilon }\bigg )\, U {\text{tan}} h(k_{\text {b}} \xi )\quad \mathrm{~with~}\quad U(t) = \frac{\text {d} S}{\text {d} t} = U_{\max } \sin \frac{t}{t_0} \end{aligned}$$
(8)

the slump velocity obtained from Eq. (4), with \(U_{\max }=S_0/t_0\) the maximum velocity. Similarly, the slump acceleration is found as

$$\begin{aligned} A(t)=\frac{\text {d}^2 S}{\text {d} t^2} = A_0 \cos \frac{t}{t_0}\quad \mathrm{~with~}\quad A_0 = \frac{S_0}{t_0^2} \end{aligned}$$
(9)

the initial acceleration.

For rigid slumps, hydrodynamic drag can be neglected due to low velocity and small amplitude of motion, and inertia includes both the SMF mass \(M_{\text {s}} = \rho _{\text {s}} V_{\text {s}}\), with \(\rho _{\text {s}}\) denoting the sediment bulk density, and the specific density being defined as \(\gamma = \rho _{\text {s}}/\rho _{\text {w}}\), with \(\rho _{\text {w}}\) the water density, and an added mass \(\Delta M_{\text {s}} = C_{\text {M}} \rho _{\text {w}} V_{\text {s}}\), defined by way of an added mass coefficient \(C_{\text {M}}\). Assuming a constant basal friction, a nearly circular rupture surface of radius R, and a small angular displacement \(\Delta \Phi\), Grilli and Watts (2005) derived the characteristic distance and time of motion for rigid slumps as

$$\begin{aligned} S_0 = \frac{R \Delta \Phi }{2} \quad\mathrm{~and~}\quad t_0 = \sqrt{\frac{R}{g} \frac{\gamma +C_{\text {M}}}{\gamma -1}}\quad\mathrm{~with~}\quad R\simeq \frac{b^2}{8T} \end{aligned}$$
(10)

with g denoting the gravitational acceleration. Equation (8), proposed by Watts et al. (2005), is a semi-empirical relationship to estimate the radius of slump motion as a function of slump downslope length and maximum thickness.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Schambach, L., Grilli, S.T., Kirby, J.T. et al. Landslide Tsunami Hazard Along the Upper US East Coast: Effects of Slide Deformation, Bottom Friction, and Frequency Dispersion. Pure Appl. Geophys. 176, 3059–3098 (2019). https://doi.org/10.1007/s00024-018-1978-7

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00024-018-1978-7

Keywords

Navigation