Chapter 7

Storm surge modelling


CHAPTER
COORDINATORS

Fujiang Yu
CHAPTER
AUTHORS

David Byrne, Jianxi Dong, Begoña Pérez Gómez, and Shichao Liu

7.1 General introduction to storm surge

Many natural phenomena can cause the sea to rise and fall, such as wind, air pressure, celestial gravity, earthquakes, etc. The sea level changes caused by different phenomena have different periods. For example, wind waves have a period of several seconds, tsunami waves of few minutes to tens of minutes, and the period of storm surge and astronomical tide is about several hours to several days (Figure 7.1). Among them, the storm surge brings huge economic losses and risks to coastal countries every year (Murty, 1988). In order to reduce the impact of storm surge disasters on coastal residents, understanding and forecasting storm surge have always been an important objective for marine forecasters. This chapter will introduce the main overview and elements of storm surge modelling, to guide technical personnel to engage in related work and give full play to the role of storm surge numerical models in various fields. 

Figure 7.1. Frequencies and periods of the vertical motions of the ocean surface (adapted from Pérez et al., 2013).

7.1.1 Overview of storm surge disaster

7.1.1.1 Disasters and forecasting

Storm surge refers to the phenomenon of abnormal water level rise in a coastal or inland body caused by strong atmospheric disturbances, such as tropical cyclones (typhoons, hurricanes), extratropical cyclones, strong winds from cold fronts, and sudden change in atmospheric pressure.

As a complex coastal dynamic process of major coastal marine disasters, storm surge has received much attention by major affected countries all over the world. Storm surge disasters are mainly caused by the abnormal water level rise and by flooding. The disaster causing factors include not only the storm surge, but also coupling with the effect of astronomical tide and nearshore waves. Storm surge disasters (Figure 7.2, Figure 7.3, and Figure 7.4) not only include the damage to ports, wharfs, dykes, but also include the disasters caused by flooding houses, farmland, and aquaculture facilities.

Figure 7.2. The impact of the storm surge caused by the super typhoon Haiyan on the Philippines, the coastal villages of Tacloban were destroyed (Credits: Photography Marcel Crozet, ILO, 11-2013).
Figure 7.3. People walk among debris next to a ship washed ashore in the aftermath of super typhoon Haiyan in Tacloban, Philippines, 11 November 2013. (Credits: ILO, 11-2013).
Figure 7.4. The impact of the storm surge caused by the super typhoon Haiyan on the Philippines, the coastal villages of Tacloban were inundated with water (Credits: Photography Marcel Crozet, ILO,11-2013).

Areas with severe storm surges are shown on a global map (Figure 7.5). The Gulf Coast of North America and the eastern coast of the United States are affected by storm surges generated by Atlantic hurricanes. In Europe, the North Sea coast is often affected by extratropical cyclones, which bring storm surge disaster. The coast of the Bay of Bengal in the Indian Ocean is threatened by storm surges caused by typhoons in the Indian Ocean. On the western Pacific Ocean coast, China, Japan, and the Philippines are frequently affected by storm surges caused by typhoons, and the north coast of China is also affected by extratropical cyclones.

Figure 7.5. Areas severely affected by storm surge.

In addition to the areas severely affected by the storm surge mentioned above, other countries or regions may also be affected by the storm surge. Areas with low elevation may face the threat of storm surge inundation, and the approaching channel may not meet the navigation requirements due to the drop in water level. For example, 20% of the land in the Netherlands is below mean sea level, and large areas of flooding may be caused without a very severe storm surge. For this reason, they built the famous Storm Surge Barriers (Mooyaart and Jonkman, 2017). In Spain, surges of 60 cm contribute significantly to inundation processes.

Storm surge forecasting is an important means of reducing disasters and losses, and a very necessary link in disaster prevention and mitigation. The methods of storm surge forecasting can be divided into two categories: empirical statistical forecasting and numerical forecasting. With the rapid development of computer technology, numerical models play an increasingly important role in storm surge forecasting. The establishment of a storm surge numerical model will provide strong support for storm surge forecasting. In addition to providing help for disaster prevention and mitigation, the numerical model of storm surge can also be used in offshore engineering design and marine disaster risk assessment of coastal cities.

In recent years, with the rapid economic development of coastal cities and the urgent needs of disaster prevention and mitigation, more and more ocean forecasting centres have started to establish operational storm surge models to provide relevant services for the above activities and purposes (more information in Section 7.2.8).

7.1.1.2 The impact of climate change on storm surge

Coastal cities are directly affected by global warming, sea temperature continues to increase, sea level fluctuates and rises, and natural disasters such as storm surges and huge waves show an increasing trend. Statistics show that there is a significant increase in global super typhoons (or category 4 and 5 hurricanes). In the 1970s, the number of super typhoons accounted for 20% of the total tropical cyclones, while it rose to 35% in the 1990s. Among them, the most evident increase was in the North Pacific, Indian Ocean, and Southwest Indian Ocean, while the increase was the least in the North Atlantic (Webster et al., 2005). Therefore, storm surge disasters caused by typhoons showed an increasing trend, as well as the risk of storm surge disasters in coastal cities. The tide observation data also shows this characteristic. After the storm surge of typhoon Hato (2017) and typhoon Mangkhut (2018) affected coastal cities such as Zhuhai and Shenzhen in China, the return period of coastal tide levels changed significantly. The Hengmen Station, located on the west bank of the Pearl River Estuary (China), has shown that the tide level return period has been reduced from 200 years to 50 years, as well as the original 100 years tide level return period been reduced to 50 years at the Sanzao Station, and the Chiwan Station on the east bank of the Pearl River Estuary.

Sea level rise directly leads to the expansion of storm surge inundation area, increases the mean sea level, and various characteristic tide levels. The increased water depth and enhanced nearshore waves further strengthen the impact of storm surges.

7.1.2 Basic description of storm surge phenomena

Storm surges have periods of several hours to several days, and are usually superimposed on tides, wind waves and swells (with a period of several seconds). Combination of these three factors causes extreme rise of coastal water level and often leads to huge storm surge disasters. However, sometimes the opposite situation can also be encountered: the wind blowing away from the direction of the open coast for a long time causes the water level to drop sharply along the shore and shoals exposed. In this case, the normal navigation is seriously affected, as well as anchoring of ships, especially large oil tankers.

The spatial range of storm surges is generally between tens and thousands of kilometres, and the time scale or period is about several to hundreds hours, which is between a tsunami and the astronomical tide. Since the area affected by storm surges moves with the movement of meteorological forcing, sometimes a storm surge process can affect a coastal area of 1000-2000km, and the impact time can be up to several days. In addition, the period of water level change by a storm surge ranges from some hours to several days, excluding seiches, tsunamis and wind waves.

According to its standard definition, a storm surge is caused by atmospheric disturbance, specifically abnormal alterations in water surface due to strong winds and atmospheric pressure changes. Storm surge can also occur in inland bodies, such as the Great Lakes in the US. In recent years, studies have shown that the nearshore waves breaking can also cause rise of the water level, in the range of tens of centimetres to metres, called wave setup. With the perspective of changes occurred in modern times, the definition of storm surges should be revised as the following: “storm surge refers to strong atmospheric disturbances, such as tropical cyclones (typhoons and hurricanes), extratropical cyclones, strong wind due to cold fronts, and sudden changes in atmospheric pressure inducing abnormal water level rise combined with nearshore wave setup”(Yu et al., 2020). See representation of storm surge components and drivers in Figure 7.6.

Figure 7.6. Storm surge components and drivers.

Meteorological tsunami, or meteotsunami, is caused by strong winds and sudden changes in atmospheric pressure and its period is equivalent to a tsunami. In the Tsunami Glossary (🔗1 ) by the IOC’s ITIC, meteotsunami is defined as following: “Tsunami-like phenomena generated by meteorological or atmospheric disturbances. These waves can be produced by atmospheric gravity waves, pressure jumps, frontal passages, squalls, gales, typhoons, hurricanes and other atmospheric sources. Meteotsunamis have the same temporal and spatial scales as tsunami waves and can similarly devastate coastal areas, especially in bays and inlets with strong amplification and well defined resonant properties (e.g. Ciutadella Inlet, Baleric Islands; Nagasaki Bay, Japan; Longkou Harbour, China; Vela Luka, Stari Grad and Mali Ston Bays, Croatia).”

The water level recorded at coastal or estuarine tide stations usually include a combination of changes caused by astronomical tides, storm surges, tsunamis, and other long waves. Generally, tide gauges filter out sea surface fluctuations caused by short-period waves in the order of seconds. The separation of storm surge phases is obtained by linear subtracting the harmonic analysis forecast astronomical tide from the hourly data (Figure 7.7 and Figure 7.8).

Figure 7.7. Observed water level, astronomical tide, and storm surge (water level subtracting astronomical tide); data from Zhanjiang tide station (China).
Figure 7.8. Observed water level, astronomical tide, and storm surge (water level subtracting astronomical tide); data from Nandu tide station (China).

7.1.3 Physics of storm surge

7.1.3.1 Meteorological forcing

Meteorological forcing is the main driver for storm surges. When a storm passes over the open sea, the low centre pressure of the storm will cause the water level to rise. The height of the surge is related to the barometric pressure drop of the storm, i.e. 1 mbar decrease corresponds approximately to 1 cm increase in sea level (Schalkwijk, 1948; Myers, 1954; Pore, 1964). The raised sea surface will propagate with the movement. At the same time, a kind of free long wave, induced by raised sea surface, could spread outward from the storm centre. This process will typically take place near the coast when interactions with bathymetry changes become relevant. If the pressure disturbance is moving at a speed comparable to the shallow water wave speed, the water level disturbance may be greatly amplified by resonance (Harris, 1957).

Compared with the long wave effect, the wind shear stress is the dominating forcing of storm surges in shallow water of nearshore and estuaries (Miller, 1958; Pore, 1964). With the wind blowing continuously, water accumulates at the coastal line causing the water level to rise. This phenomenon is referred to as "wind set-up" and its magnitude is inversely proportional to water depth. The wind set-up is particularly evident in semi-enclosed seas, such as Bohai Bay in China.

7.1.3.2 The influence of topography and bathymetry

Storm surge is not only influenced by astronomical tide and waves, but also by topography and bathymetry. Due to the shoreline block, storm surge propagates from ocean to nearshore. The surge is generated by water accumulation at the shoreline. The magnitude of the surge is controlled by the shape of the shoreline. In case of onshore direction, semi-enclosed bays or estuaries contribute to intensify storm surge than straight shoreline. That is because the shape of the semi-enclosed bay and estuary hinder water flow out. The water accumulates in the semi-enclosed bay or estuary continuously, resulting in a greater storm surge.

Another factor that can impact storm surge is the variation of bathymetry from the continental shelf to estuaries and coasts. Generally, the water depth of estuaries and coasts is shallower than the continental shelf, and the propagation speed of the storm surge wave is approximately proportional to the square root of the water depth. Therefore, the speed of the wave propagation at estuaries and coasts is slower than at the continental shelf. The storm surge waves converge at estuaries and coasts, causing the water level to increase.

On the other hand, in the process of storm surge wave propagation, the water depth at the crest is greater than at the preceding trough, and the movement of the crest is faster. So, the more waves move inland, the smaller the interval between the crests. This is more pronounced where the continental shelf is longer, for example in the North Sea, and hence larger storm surges will be caused due to the long continental shelf extension.

The propagation speed of storm surge waves is controlled by the water depth: it moves faster at high tide than at low tide. The wind effect is inversely proportional to the total water depth, and the same wind speed will produce a greater surge at low tide than at high tide. Combining the two effects above, surge in an estuary tends to be greater on the rising stage of the tide (Doodson, 1929; Doodson, 1956; Rossiter, 1961).

Extremely accurate topography and bathymetry, especially for shallow water areas, is key to storm surge modelling.

References

Álvarez-Fanjul, E., Perez Gomez, B., Sanchez Arevalo, I.R. (2001). Nivmar: A storm surge forecasting system for Spanish Waters. Scientia Marina, 65(S1), https://doi.org/10.3989/scimar.2001.65s1145

Beckers J. V. L., Sprokkereef, E., and Roscoe, K. L. (2008). Use of Bayesian model averaging to determine uncertainties in river discharge and water level forecasts, Proceeding of “4th International Symposium on Flood Defence: Managing Flood Risk, Reliability and Vulnerability”, Toronto, Ontario, Canada, 6–8 May, 2008.

Bjerknes, V. (1921). On the dynamics of the circular vortex with applications to the atmosphere and atmospheric vortex and wave motions. Quarterly Journal of the Royal Meteorological Society, 48(204), 375-376. https://doi.org/https://doi.org/10.1002/qj.49704820414

Buizza, R., and Palmer, T.N. (1995). The Singular-Vector Structure of the Atmospheric Global Circulation. Journal of the Atmospheric Sciences, 52(9), 1434-1456, https://doi.org/10.1175/1520-0469(1995)052<1434:TSVSOT>2.0.CO;2

Byrne, N. J. (2017). Deterministic models of Southern Hemisphere circulation variability. Ph.D. thesis, University of Reading, 105 pp.

Byrne, D., Horsburgh, K., and Williams, J. (2021). Variational data assimilation of sea surface height into a regional storm surge model: Benefits and limitations. Journal of Operational Oceanography, 1-14, https://doi.org/10.1080/1755876X.2021.1884405

Clementi, E., Aydogdu, A., Goglio, A. C., Pistoia, J., Escudier, R., Drudi, M., Grandi, A., Mariani, A., Lyubartsev, V., Lecci, R., Cretí, S., Coppini, G., Masina, S., and Pinardi, N. (2021). Mediterranean Sea Physical Analysis and Forecast (CMEMS MED-Currents, EAS6 system) (Version 1) set. Copernicus Monitoring Environment Marine Service (CMEMS), https://doi.org/10.25423/CMCC/medsea_analysisforecast_PHY_006_013_EAS6

Daley, R. (1991). Atmospheric Data Analysis. Cambridge University Press: Cambridge, UK

Daniel, P., Haie, B., and Aubail, X. (2009). Operational Forecasting of Tropical Cyclones Storm Surges at Meteo-France. Marine Geodesy, 32(2), 233-242. https://doi.org/10.1080/01490410902869649

de Vries, H. (2009). Probability Forecasts for Water Levels at the Coast of The Netherlands. Marine Geodesy, 32(2), 100-107, https://doi.org/10.1080/01490410902869185

Di Liberto, T., and Colle, B.A. (2011). Verification of a Multimodel Storm Surge Ensemble around New York City and Long Island for the Cool Season. Weather and Forecasting, 26, 922-939.

Doodson, A. T. (1929). Report on Thames Floods. Geophysical Memoirs, Great Britain Meteorological Office, 5(47), 1-26.

Doodson, A. T. (1956). Tides and storm surges in a long uniform gulf. Proceedings of the Royal Society of London. Series A. Mathematical and Physical Sciences, 237(1210), 325-343.

Dube, S. K., Sinha, P. C., Rao, A. D., and Rao, G. S. (1985). Numerical modelling of storm surges in the Arabian Sea. Applied Mathematical Modelling, 9(4), 289-294, https://doi.org/https://doi.org/10.1016/0307-904X(85)90067-8

Dube, S. K., Sinha, P. C., and Roy, G. D. (1984). Numerical simulation of storm surges induced by tropical storms impinging on the Bangladesh coast. Coastal Engineering Proceedings, 1(19), https://doi.org/10.9753/icce.v19.13

Egbert, G. D., Bennett, A. F., and Foreman, M. G. G. (1994). TOPEX/POSEIDON tides estimated using a global inverse model. Journal of Geophysical Research: Oceans, 99(C12), 24821-24852, https://doi.org/10.1029/94JC01894

Egbert, G. D., and Erofeeva, S. Y. (2002). Efficient Inverse Modeling of Barotropic Ocean Tides. Journal of Atmospheric and Oceanic Technology, 19(2), 183-204, https://doi.org/10.1175/1520-0426(2002)019<0183:Eimobo>2.0.Co;2

Evensem, G. (2004). Sampling strategies and square root analysis schemes for the EnKF. Ocean Dynamics, 54, 539-560, https://doi.org/10.1007/s10236-004-0099-2

Fernández-Montblanc, T., Vousdoukas, M.I., Mentaschi, L., Ciavola, P. (2020). A Pan-European high resolution storm surge hindcast. Environment International, 135, 105367, https://doi.org/10.1016/j.envint.2019.105367

Ferrarin, C., Valentini, A., Vodopivec, M., Karic, D., Massaro, G., Bajo, M., De Pascalis, F., Fadini, A., Ghezzo, M., Menegon, S., Bressan, L., Unguendoli, S., Fettich, A., Jerman, J., Licer, M., Fustar, L., Papa, A.. Carraro, E. (2020). Integrated sea storm management strategy: the 29 October 2018 event in the Adriatic Sea. Natural Hazards and Earth System Sciences, 20, 73-93, https://doi.org/10.5194/nhess-20-73-2020

Flowerdew, J., Horsburgh, K., Wilson, C., and Mylne, K. (2010). Development and evaluation of an ensemble forecasting system for coastal storm surges. Quarterly Journal of the Royal Meteorological Society, 136(651), 1444-1456, https://doi.org/10.1002/qj.648

Flowerdew, J., Mylne, K., Jones, C., and Titley, H. (2013). Extending the forecast range of the UK storm surge ensemble. Quarterly Journal of the Royal Meteorological Society, 139(670), 184-197, https://doi.org/10.1002/qj.1950

Fujita, T. (1952). Pressure distribution within typhoon. Geophysical Magazine, 23, 437-451.

Gandin, L.S. (1963). Objective analysis of meteorological fields. Translated from Russian.

Hamill, T. M. , Snyder, C., and Morss, R.E. (2000). A comparison of probabilistic forecasts from bred, singular vector, and perturbed observation ensembles. Monthly Weather Review, 128(6), 1835-1851, https://doi.org/10.1175/1520-0493(2000)128<1835:ACOPFF>2.0.CO;2

Harris, D.L. (1957). The Effect of a Moving Pressure Disturbance on the Water Level in a Lake. In: “Interaction of Sea and Atmosphere”. Meteorological Monographs, vol 2. American Meteorological Society, Boston, MA. https://doi.org/10.1007/978-1-940033-15-0_4

Hasegawa, H., Kohno, N., and Itoh, M. (2015). Development of Storm Surge Model in Japan Meteorological Agency. Available at http://www.waveworkshop.org/14thWaves/Papers/JCOMM_2015_J4.pdf

Heemink, A.W. (1986). Storm surge prediction using Kalman filtering. Dissertation TU Twente, Hydraulic Engineering Reports. Available at http://resolver.tudelft.nl/uuid:8a7c65f5-f8f5-4033-b796-adaef2eb39a7

Hersbach, H. Bell, B., Berrisford, P., Hirahara, S., Horányi, A., Muñoz-Sabater, J., Nicolas, J., Peubey, C., Radu, R., Schepers, D., Simmons, A., Soci, C., Abdalla, S., Abellan, X., Balsamo, G., Bechtold, P., Biavati, G., Bidlot, J., Bonavita, M., De Chiara, G., Dahlgren, P., Dee, D., Diamantakis, M., Dragani, R., Flemming, J., Forbes, R., Fuentes, M., Geer, A., Haimberger, L., Healy, S., Hogan, R.J., Hólm, E., Janisková, M., Keeley, S., Laloyaux, P., Lopez, P., Lupu, C., Radnoti, G., de Rosnay, P., Rozum, I., Vamborg, F., Villaume, S., Thépaut Hersbach, J.N. (2020). The ERA5 global reanalysis. Quarterly Journal of the Royal Meteorology Society, 146(730), 1999-2049, https://doi.org/10.1002/qj.3803

Holland, G. J. (1980). An Analytic Model of the Wind and Pressure Profiles in Hurricanes. Monthly Weather Review, 108(8), 1212-1218. https://doi.org/10.1175/1520-0493(1980)108<1212:Aamotw>2.0.Co;2

Horsburgh, K., and de Vries, H. (2011). Guide to Storm Surge Forecasting. WMO-No. 1076. ISBN 978-92-63-11076-3

Jarvinen, B. R., and Lawrence, M. B. (1985). An evaluation of the SLOSH storm-surge model. Bulletin of the American Meteorological Society, 66(11), 1408-1411.

Jelesnianski, C. P. (1965). A numerical calculation of storm tides induced by a tropical storm impinging on a continental shelf. Monthly Weather Review, 93(6), 343-358. https://doi.org/10.1175/1520-0493(1993)093<0343:Ancos>2.3.Co;2

Kalman, R. E. (1960). A New Approach to Linear Filtering and Prediction Problems. Journal of Basic Engineering, 82(1), 35-45. https://doi.org/10.1115/1.3662552

Knaff, J.A., De Maria, M., and Molenar, D.A. (2011). An automated, objective, multiple satellite platform tropical cyclone surface wind analysis. Journal of Applied Meteorology and Climatology, 50(10), pp.2149-2166, https://doi.org/10.1175/2011JAMC2673.1

Le Roy, S., Pedreros, R., André, C., Paris, F., Lecacheux, S., Marche, F., and Vinchon, C. (2015). Coastal flooding of urban areas by overtopping: dynamic modelling application to the Johanna storm (2008) in Gâvres (France). Natural Hazards and Earth System Sciences, 15, 2497-2510, https://doi.org/10.5194/nhess-15-2497-2015

Leith, C.E. (1974). Theoretical Skill of Monte Carlo Forecasts. Monthly Weather Review, 102 (6), 409-418, https://doi.org/10.1175/1520-0493(1974)102<0409:TSOMCF>2.0.CO;2

Liu, Q., Jiang, J., Yu, F., Zhang, C., Dong, J., Song, X., and Wang, Y. (2020). Typhoon storm surge ensemble forecast based on GPU technique. Acta Oceanologica Sinica, 39(5), 77-86, https://doi.org/10.1007/s13131-020-1570-8

Loftis, J.D., Wang, H., Forrest, D., Rhee, S., Nguyen, C. (2017). Emerging flood model validation frameworks for street-level inundation modeling with StormSense. SCOPE '17: Proceedings of the 2nd International Workshop on Science of Smart City Operations and Platforms Engineering, April 2017, 13-18, https://doi.org/10.1145/3063386.3063764

Lorenc, A. (1981). A global three-dimensional multivariate statistical interpolation scheme. Monthly Weather Review, 109, 701-721.

Lorenc, A. (1986). Analysis methods for numerical weather prediction. Quarterly Journal of the Royal Meteorology Society, 112, 1177-1194.

Lorenz, E.N. (1965). A study of the predictability of a 28-variable atmospheric model. Massachusetts Institute of Technology, Cambridge, Massachusetts. Tellus, 17(3), 321-333.

Lyard, F.H., Allain, D.J., Cancet, M., Carrere, L., Picot, N. (2021). FES2014 global ocean tide atlas: design and performance. Ocean Science, 17, 615-649, https://doi.org/10.5194/os-17-615-2021

Madsen, K.S., Hoyer, J., Fu, W., Donlon, C. (2015). Blending of satellite and tide gauge sea level observations and its assimilation in a storm surge model of the North Sea and Baltic Sea. Journal of Geophysical Research: Oceans, 20(9), 6405-6418, https://doi.org/10.1002/2015JC011070

Matsumoto, K., Sato, T., Takanezawa, T., and Ooe, M. (2001). GOTIC2: A program for computation of oceanic tidal loading effect. Journal of the Geodetic Society of Japan, 47(1), 243-248, https://doi.org/10.11366/sokuchi1954.47.243

Matsumoto, K., Takanezawa, T., and Ooe, M. (2000). Ocean Tide Models Developed by Assimilating TOPEX/POSEIDON Altimeter Data into Hydrodynamical Model: A Global Model and a Regional Model around Japan. Journal of Oceanography, 56(5), 567-581, https://doi.org/10.1023/A:1011157212596

Mel, R., and Lionello, L. (2014a). Storm Surge Ensemble Prediction for the City of Venice. Weather and Forecasting, 29(4), 1044-1057, https://doi.org/10.1175/WAF-D-13-00117.1

Mel, R., and Lionello, P. (2014b). Verification of an ensemble prediction system for storm surge forecast in the Adriatic Sea. Ocean Dynamics, 64(12), 1803-1814, https://doi.org/10.1007/s10236-014-0782-x

Mel R. and Lionello, P. (2016). Probabilistic Dressing of a Storm Surge Prediction in the Adriatic Sea. Advances in Meteorology, 3764519, 8, https://doi.org/10.1155/2016/3764519

Miller, A. R. (1958). The Effects of Winds on Water Levels on the New England Coast 1. Limnology and Oceanography, 3(1), 1-14.

Minato, S. (1998). Storm surge simulation using POM and a revisitation of dynamics of sea surface elevation short-term variation. Oceanographic Literature Review, 7(45), 1103.

Molteni, F., R. Buizza, T. N. Palmer, and T. Petroliagis (1996). The ECMWF Ensemble Prediction System: Methodology and validation. Quarterly Journal of the Royal Meteorological Society, 122, 73-119, https://doi.org/10.1002/qj.49712252905

Mooyaart, L., and Jonkman, S. N. (2017). Overview and design considerations of storm surge barriers. Journal of Waterway, Port, Coastal, and Ocean Engineering, 143(4), 06017001.

Muis, S., Verlaan, M., Winsemius, H.C., Aerts, J.C.J.H., Ward, P.J. (2016). A global reanalysis of storm surges and extreme sea levels. Nature Communications, 7, 11969, https://doi.org/10.1038/ncomms11969

Muis, S., Irazoqui Apecechea, M., Dullaart, J., de Lima Rego, J., Madsen, K.S., Su, J., Yan, K., and Verlaan M. (2020). A High-Resolution Global Dataset of Extreme Sea Levels, Tides, and Storm Surges, Including Future Projections. Frontiers in Marine Science, 7(263), https://doi.org/10.3389/fmars.2020.00263

Murty, T. (1988). List of major natural disasters, 1960-1987. Natural Hazards, 1(3), 303-304.

Myers, V.A. (1954). Characteristics of United States hurricanes pertinent to levee design for Lake Okeechobee, Florida. US Government Printing Office.

Palmer, T., and Williams, P. (2010). Stochastic Physics and Climate Modelling. Cambridge University Press.

Pérez, B., Álvarez Fanjul, E., Pérez, S., de Alfonso, M., Vela, J. (2013). Use of tide gauge data in operational oceanography and sea level hazard warning systems, Journal of Operational Oceanography, 6(2), 1-18, https://doi.org/10.1080/1755876X.2013.11020147

Pérez, B.,Brouwer, R., Beckers, J., Paradis, D., Balseiro, C., Lyons, K., Cure, M., Sotillo, M.G., Hackett, B., Verlaan, M., Fanjul, E.A. (2012). ENSURF: Multi-model sea level forecast implementation and validation results for the IBIROOS and Western Mediterranean regions. Ocean Science, 8(2), 211-226, https://doi.org/10.5194/os-8-211-2012

Pérez-Gómez, B., Pérez González, I., Sotillo, M. G., and Álvarez-Fanjul, E. (2019). Retos de los sistemas de observación y predicción en los riesgos asociados al nivel del mar. Ribagua, 6, 1-5, https://doi.org/10.1080/23863781.2019.1595212

Pérez González, I., Pérez-Gómez, B., Sotillo, M. G., and Álvarez-Fanjul, E. (2017). Towards a new sea level forecast system in Puertos del Estado. In Proceedings of the 8th EuroGOOS Conference, Bergen.

Pore, N. A. (1964). The relation of wind and pressure to extratropical storm surges at Atlantic City. Journal of Applied Meteorology and Climatology, 3(2), 155-163.

Ringler, T., Petersen, M., Higdon, R. L., Jacobsen, D., Jones, P. W., and Maltrud, M. (2013). A multi-resolution approach to global ocean modeling. Ocean Modelling, 69, 211-232, https://doi.org/10.1016/j.ocemod.2013.04.010

Rossiter, J. R. (1961). Interaction between tide and surge in the Thames. Geophysical Journal International, 6(1), 29-53.

Schalkwijk, W.-F. (1948). A contribution to the study of storm surges on the Dutch coast. Ciel et Terre, 64,188.

Siek, M., and Solomatine, D. (2011). Real-time Data Assimilation for Chaotic Storm Surge Model Using NARX Neural Network. Journal of Coastal Research,1189-1194, https://www.jstor.org/stable/26482362

Sotillo, M.G., Cailleau, S., Lorente, P., Levier, B., Aznar, R., Reffray, G., Amo-Baladron, A., Chanut, J., Benkiran, M., Alvarez-Fanjul, E. (2015). The MyOcean IBI Ocean Forecast and Reanalysis Systems: operational products and roadmap to the future Copernicus Service. Journal of Operational Oceanography, 8(1), https://doi.org/10.1080/1755876X.2015.1014663

Takahashi, K. (1939). Distribution of pressure and wind in a typhoon. Journal of the Meteorological Society of Japan, 17(2), 417-421.

Ueno, T. (1981). Numerical computations of the storm surges in tosa bay. Journal of the Oceanographical Society of Japan, 37(2), 61-73, https://doi.org/10.1007/BF02072559

Verlaan, M., Zijderveld, A., de Vries, H., and Kroos, J. (2005). Operational storm surge forecasting in the Netherlands: developments in the last decade. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 363(1831), 1441-1453, https://doi.org/doi:10.1098/rsta.2005.1578

Wang, P., Yu, F., Liu, Q., and Dong, J. (2010). Research and application of fine ensemble numerical forecasting technology for typhoon storm surge in Fujian coastal. Marine Forecasts (in Chinese), 27(05), 7-15.

Webster, P. J., Holland, G. J., Curry, J. A., and Chang, H.-R. (2005). Changes in Tropical Cyclone Number, Duration, and Intensity in a Warming Environment. Science, 309(5742), 1844-1846, https://doi.org/10.1126/science.1116448

Weisberg, R. H., and Zheng, L. (2008). Hurricane storm surge simulations comparing three-dimensional with two-dimensional formulations based on an Ivan-like storm over the Tampa Bay, Florida region. Journal of Geophysical Research: Oceans, 113(C12), https://doi.org/10.1029/2008JC005115

WMO. (2012). Guidelines on Ensemble Prediction Systems and Forecasting.

Ye, F., Zhang, Y. J., Yu, H., Sun, W., Moghimi, S., Myers, E., Nunez, K., Zhang, R., Wang, H. V., and Roland, A. (2020). Simulating storm surge and compound flooding events with a creek-to ocean model: Importance of baroclinic effects. Ocean Modelling, 145, 101526, https://doi.org/10.1016/j.ocemod.2019.101526

Yu, F., Fu, C., Guo, H., and Liu, Q. (2020). Modern technologies and Application in storm surge forecasting. Science Press (in Chinese).

Zheng, L., Weisberg, R. H., Huang, Y., Luettich, R. A., Westerink, J. J., Kerr, P. C., Donahue, A. S., Crane, G., and Akli, L. (2013). Implications from the comparisons between two- and three-dimensional model simulations of the Hurricane Ike storm surge. Journal of Geophysical Research: Oceans, 118(7), 3350-3369, https://doi.org/10.1002/jgrc.20248

Zijl, F., Sumihar, J., Verlaan, M. (2015). Application of data assimilation for improved operational water level forecasting on the northwest European shelf and North Sea. Ocean Dynamics, 65, 1699-1716, https://doi.org/10.1007/s10236-015-0898-7

To start contributing, sharing knowledge and editing the WIKI, please login