• 0 Shopping Cart

Internet Geography

2018 Sulawesi, Indonesia Earthquake and Tsunami Case Study

What caused the Sulawesi, Indonesian earthquake and what were the effects?

On Friday 28th September 2018 a magnitude 7.5 earthquake struck Palu, on the Indonesian island of Sulawesi, just before dusk wreaking havoc and destruction across the city and triggering a deadly tsunami on its coast. The 7.5 magnitude earthquake hit only six miles from the country’s coast.

A map to show the location of Palu

A map to show the location of Palu

The shallow tremor was more powerful than a series of earthquakes that killed hundreds on the Indonesian island of Lombok this July and August.

Palu is located on the Indonesian island of Sulawesi, 1,650 kilometres northeast of Jakarta, at the mouth of the Palu River. It is the capital of the province of Central Sulawesi, situated on a long, narrow bay .

A satellite image to show the location of Palu

A satellite image to show the location of Palu – Source Google Earth

The coastal city of Palu is home to 350,000 people.

Small foreshocks had been happening throughout 28th September in Palu. However, in the early evening, the Palu-Koru fault suddenly slipped, a short distance offshore and only 10km (6 miles) below the surface. This generated the 7.5 magnitude earthquake.

The impact of the earthquake was magnified because of the thick layers of sediment on which the city lies. Whereas bedrock shakes in an earthquake, sediment moves a lot more, behaving like a liquid. Poorly constructed houses cannot withstand movement of this magnitude.

Scientists don’t pay much attention to the Palu-Koru fault line, as far as tsunamis are concerned.  This is because the two plates are moving past each other, not with the vertical thrust required to form a tsunami.

Scientists are still trying to work out what happened to cause the tsunami. It is possible that the earthquake caused an underwater landslide which disturbed the water or there could be inaccuracies in the identification of the type of fault.

Once the wave started moving, Palu, at the end of a narrow 10km-long bay, was a sitting duck.

Tsunamis are no danger when out at sea. But when the waves come closer to land, their base drags on the seabed causing them to rise up.

Primary Effects

The quake destroyed thousands of homes in the city, as well as an eight-storey hotel, hospital and a large department store.

More before/after comparisons from around the #PaluTsunami and #PaluEarthquake captured by @planetlabs . Included rough lat/long. Keep an eye on https://t.co/Kz73HlYmGF as they often post the sat. imagery for responders, relief agencies et al. pic.twitter.com/1Vreovjt9b — Murray Ford (@mfordNZ) October 1, 2018

At least 2256 people have been confirmed dead, with more than 10,679 injured and 1075 missing.  200,000 people were in urgent need of assistance, about a quarter of them children.

The earthquake caused widespread liquefaction , which is when soil and groundwater mix. The ground becomes very soft, similar to quicksand. It causes foundations of buildings and other structures to sink into the ground.

In the case of Palu, buildings not only collapsed but some were moved by the liquefaction. This is why it is better to build on bedrock rather than on top of the soil.

The control tower and runway at Palu’s airport also sustained damage. Commercial flights were cancelled with only humanitarian and search and rescue flights permitted.

Secondary Effects

The earthquake triggered a tsunami reaching 6 metres in height. As the tsunami approached the coast it was reported to be travelling 250mph. The damage was as extensive: the main highway was cut off by a landslide and a large bridge washed away by the tsunami wave, which hit Palu’s Talise beach and the coastal town of Donggala.

Landslides, downed communications networks and collapsed bridges have made it hard for aid workers and rescuers to reach rural areas.

Due to hospitals being damaged, people received medical treatment in the open.

Strong aftershocks hit the island the day after the earthquake.

Immediate (Short Term) Response

A tsunami warning was issued by Indonesia’s geophysics agency (BMKG) when the earthquake was detected. However, the agency lifted the warning 34 minutes after it was first issued. The closest tidal sensor to Palu is around 200km (125 miles) away. The decision to lift the tsunami warning was based on this data.

Search and rescue teams were deployed to the worst-affected areas. Around 700 army and police officers were dispatched to assist in the emergency response.

The military sent cargo planes with aid from Jakarta and other cities. However, this was slow to arrive.

A large number of charities set up appeals to raise funds to support people in the affected area. Buckingham Palace reported that the Queen had made a donation to the Disasters Emergency Committee (DEC) appeal for survivors, which raised £6m in a day when it was launched.

The RAF delivered thousands of shelter kits, solar lanterns and water purifiers to the disaster zone in addition to trucks and power generators to help get them to where they are needed.

At least 70,000 people gathered in evacuation sites across the island.

Long-term Response

Further reading.

Indonesia tsunami: UK charities launch a joint appeal – BBC News

Internet Geography Plus

Premium Resources

Please support internet geography.

If you've found the resources on this page useful please consider making a secure donation via PayPal to support the development of the site. The site is self-funded and your support is really appreciated.

Related Topics

Use the images below to explore related GeoTopics.

Lombok Indonesia Earthquake 2018 Case Study

Topic home, 2010 chile earthquake, share this:.

  • Click to share on Twitter (Opens in new window)
  • Click to share on Facebook (Opens in new window)
  • Click to share on Pinterest (Opens in new window)
  • Click to email a link to a friend (Opens in new window)
  • Click to share on WhatsApp (Opens in new window)
  • Click to print (Opens in new window)

If you've found the resources on this site useful please consider making a secure donation via PayPal to support the development of the site. The site is self-funded and your support is really appreciated.

Search Internet Geography

Top posts and pages.

Home

Latest Blog Entries

Britains Most Desirable Towns

Pin It on Pinterest

  • Click to share
  • Print Friendly

a case study on tsunami

  • Login or signup to the library
  • my Channels
  • All Channels
  • List events
  • List channels
  • List collections

You are here

Tsunami disasters: case studies and reports.

a case study on tsunami

This Collection is part of the 'Tsunami Disaster Channel' containing a number of case studies and reports relevant to tsunami disasters, where we try to find out what we have learnt from the past and how we can best reduce risk in future natural disasters. Current guidance comes from leading global organizations: Foreign - Commonwealth & Development Office (FCDO) ,    Swiss Resource Centre and Consultancies for Development Foundation (SKAT) ,  Office of the UN Secretary General Special Envoy for Tsunami Recovery ,  United Nations Children's Fund (UNICEF) ,  United Nations Environment Programme (UNEP) ,  United Nations International Strategy for Disaster Reduction (UNISDR) .   Please send suggestions for additional content for this Collection to  [email protected] . You might find other helpful collections on tsunami disasters below."

Resources on this Collection

a case study on tsunami

10 Lessons Learned from the South Asia Tsunami of 26th December 2004

a case study on tsunami

Approaches to Equity in Post-Tsunami Assistance - Sri Lanka: A Case Study

a case study on tsunami

Environment and Reconstruction in Aceh: two years after the tsunami

a case study on tsunami

Evolving Strategies For Long-term Rehabilitation On Shelter and Development in the Tsunami Affected Areas of Tamil Nadu

a case study on tsunami

Impact of the tsunami response on local and national capacities: Maldives country report

a case study on tsunami

Indian Ocean Earthquake and Tsunami UNICEF response at six months update

You must be logged in to post a comment

General Review of the Worldwide Tsunami Research

  • Published: 15 April 2023
  • Volume 22 , pages 14–24, ( 2023 )

Cite this article

  • Sixue Cheng 1 , 2 &
  • Haijiang Liu 2  

154 Accesses

Explore all metrics

With the advancement of the global economy, the coastal region has become heavily developed and densely populated and suffers significant damage potential considering various natural disasters, including tsunamis, as indicated by several catastrophic tsunami disasters in the 21st century. This study reviews the up-to-date tsunami research from two different viewpoints: tsunamis caused by different generation mechanisms and tsunami research applying different research approaches. For the first issue, earthquake-induced, landslide-induced, volcano eruption-induced, and meteorological tsunamis are individually reviewed, and the characteristics of each tsunami research are specified. Regarding the second issue, tsunami research using post-tsunami field surveys, numerical simulations, and laboratory experiments are discussed individually. Research outcomes from each approach are then summarized. With the extending and deepening of the understanding of tsunamis and their inherent physical insights, highly effective and precise tsunami early warning systems and countermeasures are expected for the relevant disaster protection and mitigation efforts in the coastal region.

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

Access this article

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

Ai C, Ma Y, Yuan C, Xie Z, Dong G (2021) A three-dimensional non-hydrostatic model for tsunami waves generated by submarine landslides. Appl. Math. Model. 96: 1–19. DOI: https://doi.org/10.1016/j.apm.2021.02.014

Article   MathSciNet   MATH   Google Scholar  

Amores A, Monserrat S, Marcos M, Argüeso D, Villalonga J, Jordà G, Gomis D (2022) Numerical simulation of atmospheric Lamb waves generated by the 2022 Hunga-Tonga volcanic eruption. Geophys. Res. Lett. 49: e2022GL098240. DOI: https://doi.org/10.1029/2022GL098240

Article   Google Scholar  

Borrero JC (2005) Field survey of Northern Sumatra and Banda Aceh, Indonesia after the Tsunami and Earthquake of 26 December 2004. Seismol. Res. Lett. 76(3): 312–320. DOI: https://doi.org/10.1785/gssrl.76.3.312

Bourgeois J (2009) Geologic effects and records of tsunamis. In: E. N. Bernard, A. R. Robinson (eds.) The Sea, 15: 53–91

Google Scholar  

Brunt KM, Okal EA, MacAyeal DR (2011) Antarctic ice-shelf calving triggered by the Honshu (Japan) earthquake and tsunami, March 2011. J. Glaciol. 57(205): 785–788. DOI: https://doi.org/10.3189/002214311798043681

Burwell D, Tolkova E, Chawla A (2007) Diffusion and dispersion characterization of a numerical tsunami model. Ocean Model. 19: 10–30. DOI: https://doi.org/10.1016/j.ocemod.2007.05.003

Carvajal M, Sepúlveda I, Gubler A, Garreaud R (2022) Worldwide signature of the 2022 Tonga Volcanic Tsunami. Geophys. Res. Lett. 49: e2022GL098153. DOI: https://doi.org/10.1029/2022GL098153

Cheng S, Liu H (2020) Are all KdV-type shallow water wave equations the same with uniform solutions? Coastal Eng. J. 62(4): 460–472. DOI: https://doi.org/10.1080/21664250.2020.1796240

Cheng S, Liu H (2023) Weakly nonlinear waves over the bottom disturbed topography: Korteweg-de Vries equation with variable coefficients. Eur. J. Mech. B-Fluid 98: 238–246. DOI: https://doi.org/10.1016/j.euromechflu.2022.12.010

Cheng S, Zeng J, Liu H (2020) A comprehensive review of the worldwide existing tsunami databases. J. Earthq. Tsunami 14(5): 2040003. DOI: https://doi.org/10.1142/S1793431120400035

Choi BH, Pelinovsky E, Kim KO, Lee JS (2003) Simulation of the trans-oceanic tsunami propagation due to the 1883 Krakatau volcanic eruption. Nat. Hazards. Earth. Syst. Sci. 3: 321–332. DOI: https://doi.org/10.5194/nhess-3-321-2003

Deng X, Liu H, Jiang Z, Baldock T (2016) Swash flow properties with bottom resistance based on the method of characteristics. Coastal Eng. 114: 25–34. DOI: https://doi.org/10.1016/j.coastaleng.2016.03.012

Duan W, Zhao B (2013) Simulation on 2D underwater landslide-induced tsunamis. Theor. App. Mech. Lett. 3: 032004. DOI: https://doi.org/10.1063/2.1303204

Elbanna A, Abdelmeguida M, Ma X, Amlanic F, Bhatd HS, Synolakise C, Rosakis AJ (2021) Anatomy of strike-slip fault tsunami genesis. P. Natl Acad. Sci USA (PNAS) 118(19): e2025632118. DOI: https://doi.org/10.1073/pnas.2025632118

Exton M, Yeh H (2022) Effects of an impermeable layer on pore pressure response to tsunami-like inundation. Proc. R. Soc. A 478: 20210605. DOI: https://doi.org/10.1098/rspa.2021.0605

Fritz HM, Borrero JC (2006) Somalia field survey after the December 2004 Indian Ocean Tsunami. Earthquake Spectra 22 (3_suppl.): 219–233. DOI: https://doi.org/10.1193/1.2201972

Fritz HM, Blount CD, Thwin S, Thu MK, Chan N (2009a) Cyclone Nargis storm surge in Myanmar. Nat. Geosci. 2: 448–449. DOI: https://doi.org/10.1038/ngeo558

Fritz HM, Mohammed F, Yoo J (2009b) Lituya Bay landslide impact generated Mega-Tsunami 50th Anniversary. In: Cummins, P. R., Satake, K., Kong, L.S.L. (eds) Tsunami Science Four Years After the 2004 Indian Ocean Tsunami, Pageoph Topical Volumes, Birkhäuser Basel. DOI: https://doi.org/10.1007/978-3-0346-0064-4_9

Fritz HM, Phillips DA, Okayasu A, Shimozono T, Liu H, Mohammed F, Skanavis V, Synolakis CE, Takahashi T (2012) The 2011 Japan tsunami current velocity measurements from survivor videos at Kesennuma Bay using LiDAR. Geophys. Res. Lett. 39: L00G23. DOI: https://doi.org/10.1029/2011GL050686

Fuhrman DR, Madsen PA (2009) Tsunami generation, propagation, and run-up with a high-order Boussinesq model. Coastal Eng. 56(7): 747–758. DOI: https://doi.org/10.1016/j.coastaleng.2009.02.004

Fujii Y, Satake K, Sakai S, Shinohara M, Kanazawa T (2011) Tsunami source of the 2011 off the Pacific coast of Tohoku Earthquake. Earth, Planets Space 63: 55. DOI: https://doi.org/10.5047/eps.2011.06.010

Gao X, Zhao G, Niu X (2022) An approach for quantifying nearshore tsunami height probability and its application to the Pearl River Estuary. Coast. Eng. 175: 104139. DOI: https://doi.org/10.1016/j.coastaleng.2022.104139

Geist EL, Parsons T (2006) Probabilistic analysis of tsunami hazards. Nat. Hazards. 37: 277–314. DOI: https://doi.org/10.1007/s11069-005-4646-z

Geist EL, Lynett PJ (2014) Source processes for the probabilistic assessment of tsunami hazards. Oceanography 27: 86–93. DOI: https://doi.org/10.5670/oceanog.2014.43

Goseberg N, Wurpts A, Schlurmann T (2013) Laboratory-scale generation of tsunami and long waves. Coastal Eng. 79: 57–74. DOI: https://doi.org/10.1016/j.coastaleng.2013.04.006

Goto C, Ogawa Y, Shuto N, Imamura F (1997) Numerical method of tsunami simulation with the leap-frog scheme. IUGG/IOC Time project, Manuals and Guides, UNESCO

Gregg CE, Houghton BF, Paton D, Lachman R, Lachman J, Johnston DM, Wongbusarakum S (2006) Natural warning signs of tsunamis: Human sensory experience and response to the 2004 Great Sumatra Earthquake and Tsunami in Thailand. Earthquake Spectra 22 (3_suppl.): 671–691. DOI: https://doi.org/10.1193/1.2206791

Grilli ST, Tappin DR, Carey S, Watt SFL, Ward SN, Grilli AR, Engwell SL, Zhang C, Kirby JT, Schambach L, Muin M (2019) Modelling of the tsunami from the December 22, 2018 lateral collapse of Anak Krakatau volcano in the Sunda Straits, Indonesia. Sci. Rep. 9: 11946. DOI: https://doi.org/10.1038/s41598-019-48327-6

Gusman AR, Tanioka Y, Sakai S, Tsushima H (2012) Source model of the great 2011 Tohoku earthquake estimated from tsunami waveforms and crustal deformation data. Earth Planet. Sci. Lett. 341–344: 234–242. DOI: https://doi.org/10.1016/j.epsl.2012.06.006

Han P, Yu H, Yu X (2021) A sloshing induced Tsunami: 2018 Palu Bay event. Applied Ocean Res. 117: 102915. DOI: https://doi.org/10.1016/j.apor.2021.102915

Hossen MJ, Cummins PR, Dettmer J, Baba T (2015) Tsunami waveform inversion or sea surface displacement following the 2011 Tohoku earthquake: importance of dispersion and source kinematics. J. Geophys. Res. Solid. Earth. 120: 6452–6473. DOI: https://doi.org/10.1002/2015JB011942

Horspool N, Pranantyo I, Griffin J, Latief H, Natawidjaja DH, Kongko W, Cipta A, Bustaman B, Anugrah SD, Thio HK (2014) A probabilistic tsunami hazard assessment for Indonesia. Nat. Hazards Earth Syst. Sci. 14: 3105–3122. DOI: https://doi.org/10.5194/nhess-14-3105-2014 .

Ide S, Baltay A, Beroza G (2011) Shallow dynamic overshoot and energetic deep rupture in the 2011 M w 9.0 Tohoku-Oki earthquake. Science 332: 1426–1429. DOI: https://doi.org/10.1126/science.1207020

IOC-UNESCO (1998) Post- Tsunami survey field guide. 1st ed., IOC Manuals and Guides, Intergovernmental Oceanographic Commission

Ishimura D, Miyauchi T (2015) Historical and paleo-tsunami deposits during the last 4000 years and their correlations with historical tsunami events in Koyadori on the Sanriku Coast, northeastern Japan. Prog. Earth Planetary Sc. 2: 16. DOI: https://doi.org/10.1186/s40645-015-0047-4

Kataoka R, Winn SD, Touber E (2022) Meteotsunamis in Japan associated with the Tonga Eruption in January 2022. Scientific Online Letters on the Atmosphere 18: 103–106. DOI: https://doi.org/10.2151/sola.2022-019

Kato A, Obara K, Igarashi T, Tsuruoka H, Nakagawa S, Hirata N (2012) Propagation of slow slip leading up to the 2011 M w 9.0 Tohoku-Oki earthquake. Science 335: 705–708. DOI: https://doi.org/10.1126/science.1215141

Kihara N, Niida Y, Takabatake D, Kaida H, Shibayama A, Miyagawa Y (2015) Large-scale experiments on tsunami-induced pressure on a vertical tide wall. Coastal Eng. 99: 46–63. DOI: https://doi.org/10.1016/j.coastaleng.2015.02.009

Kihara N, Arikawa T, Asai T, Hasebe M, Ikeya T, Inoue S, Kaida H, Matsutomi H, Nakano Y, Okuda Y, Okuno S, Ooie T, Shigihara Y, Shoji G, Tateno T, Tsurudome C, Watanabe M (2021) A physical model of tsunami inundation and wave pressures for an idealized coastal industrial site. Coastal Eng. 169: 103970. DOI: https://doi.org/10.1016/j.coastaleng.2021.103970

Kowalik Z, Murty TS (1989) On some future tsunamis in the Pacific Ocean. Nat. Hazards 1(4): 349–369. DOI: https://doi.org/10.1007/BF00134833

Krautwald C, Hafen HV, Niebuhr P, Vogele K, Schurenkamp D, Sieder M, Goseberg N (2022) Large-scale physical modeling of broken solitary waves impacting elevated coastal structures. Coastal Eng. J. 64(1): 169–189. DOI: https://doi.org/10.1080/21664250.2021.2023380

Kubota T, Saito T, Nishida K (2022) Global fast-traveling tsunamis driven by atmospheric Lamb waves on the 2022 Tonga eruption. Science 377: 91–94. DOI: https://doi.org/10.1126/science.abo4364

Latter JN (1981) Tsunamis of volcanic origin: summary of causes with particular references to Krakatoa, 1883. Bull Volcanol 44(3): 467–490. DOI: https://doi.org/10.1007/BF02600578

Lau AYA, Switzer AD, Dominey-Howes D, Aitchison JC, Zong Y (2010) Written records of historical tsunamis in the northeastern South China Sea — challenges associated with developing a new integrated database. Nat. Hazards Earth Syst. Sci. 10: 1793–1806. DOI: https://doi.org/10.5194/nhess-10-1793-2010

Lauber G, Hager WH (1998) Experiments to dambreak wave: Horizontal channel. J. Hydraul. Res. 36(3): 291–307. DOI: https://doi.org/10.1080/00221689809498620

LeVeque RJ, George DL (2008) High-resolution finite volume methods for the shallow water equations with topography and dry-states. Advances in Coastal and Ocean Engineering, Advanced Numerical models for Simulating Tsunami Waves and Runup, 43–73. DOI: https://doi.org/10.1142/9789812790910_0002

Li L, Qiu Q, Li Z, Zhang P (2022) Tsunami hazard assessment in the South China Sea: A review of recent progress and research gaps. Sci. China Earth Sci. 65: 783–809. DOI: https://doi.org/10.1007/s11430-021-9893-8

Li Y, Raichlen F (2001) Solitary wave runup on plane slopes. J. Waterway, Port, Coastal Eng. 127(1): 33–44. DOI: https://doi.org/10.1061/(ASCE)0733-950X(2001)127:1(33)

Liu H, Liu H (2017) Experimental study on the dam-break hydrodynamic characteristics under different conditions. J. Disaster Res. 12(1): 198–207. DOI: https://doi.org/10.20965/jdr.2017.p0198

Liu H, Sakashita T, Sato S (2014) An experimental study on the tsunami boulder movement. Proceedings of 34th International Conference on Coastal Engineering, ICCE2014, Seoul. DOI: https://doi.org/10.9753/icce.v34.currents.16

Liu H, Shimozono T, Takagawa T, Okayasu A, Fritz HM, Sato S, Tajima Y (2013) The 11 March 2011 Tohoku tsunami survey in Rikuzentakata and comparison with historical events. Pure Appl. Geophys. 170(6): 1033–1046. DOI: https://doi.org/10.1007/s00024-012-0496-2

Liu PLF, Woo SB, Cho YS (1998) Computer programs for tsunami propagation and inundation. Technical Report, Cornell University. Liu PLF, Lynett P, Fernando H, Jaffe BE, Fritz H, Higman B, Morton R, Goff J, Synolakis C (2005) Observations by the international tsunami survey team in Sri Lanka. Science 308: 1595. DOI: https://doi.org/10.1126/science.1110730

Lobovsky L, Botia-Vera E, Castellana F, Mas-Soler J, Souto-Iglesias A (2014) Experimental investigation of dynamic pressure loads during dam break. J Fluid Struct. 48: 407–434. DOI: https://doi.org/10.1016/j.jfluidstructs.2014.03.009

Lu S, Liu H, Deng X (2018) An experimental study of the run-up process of breaking bores generated by dam-break under dry and wet bed conditions. J Earthq. Tsunami 12(2): 1840005. DOI: https://doi.org/10.1142/S1793431118400055

MacInnes BT, Gusman AR, LeVeque RJ, Tanioka Y (2013) Comparison of earthquake source models for the 2011 Tohoku event using tsunami simulations and nearfield observations. Bull. Seismol. Soc. Am. 103: 1256–1274. DOI: https://doi.org/10.1785/0120120121

Maeno F, Imamura F (2011) Tsunami generation by a rapid entrance of pyroclastic flow into the sea during the 1883 Krakatau eruption, Indonesia. J. Geophys. Res. 116: B09205. DOI: https://doi.org/10.1029/2011JB008253

McMurthy GM, Watts P, Fryer GJ, Smith JR, Imamura F (2004) Giant landslides, mega-tsunamis and paleo-sea level in the Hawaiian Islands. Mar. Geol. 203: 219–233. DOI: https://doi.org/10.1016/S0025-3227(03)00306-2

Meilianda E, Dohmen-Janssen CM, Maathuisc BHP, Hulscherb SJMH, Mulder JPM (2010) Short-term morphological responses and developments of Banda Aceh coast, Sumatra Island, Indonesia after the tsunami on 26 December 2004. Mar. Geol. 275: 96–109. DOI: https://doi.org/10.1016/j.margeo.2010.04.012

Minoura K, Imamura F, Sugawara D, Kono Y, Iwashita T (2001). The 869 Jogan tsunami deposit and recurrence interval of large-scale tsunami on the Pacific coast of northeast Japan. J. Natural Disaster Sci. 23(2): 83–88

Mori N, Takahashi T, Yasuda T, Yanagisawa H (2011) Survey of 2011 Tohoku earthquake tsunami inundation and run-up. Geophys. Res. Lett. 38: L00G14. DOI: https://doi.org/10.1029/2011GL049210

Mori N, Takahashi T, Esteban M (2012) The 2011 Tohoku Earthquake Tsunami Joint Survey Group, 2012. Nationwide post event survey and analysis of the 2011 Tohoku Earthquake Tsunami. Coastal Eng. J. 54(4): 1250001. DOI: https://doi.org/10.1142/S0578563412500015

Muhari A, Heidarzadeh M, Susmoro H, Nugroho HD, Kriswati E, Supartoyo Wijanarto AB, Imamura F, Arikawa T (2019) The December 2018 Anak Krakatau Volcano Tsunami as inferred from Post-Tsunami field surveys and spectral analysis. Pure Appl. Geophys. 176: 5219–5233. DOI: https://doi.org/10.1007/s00024-019-02358-2

Myers EP, Baptista AM (2001) Analysis of factors influencing simulations of the 1993 Hokkaido Nansei-Oki and 1964 Alaska Tsunamis. Nat. Hazards 23: 1–28. DOI: https://doi.org/10.1023/A:1008150210289

Nagai K, Muhari A, Pakoksung K, Watanabe M, Suppasri A, Arikawa T, Imamura F (2021) Consideration of submarine landslide induced by 2018 Sulawesi earthquake and tsunami within Palu Bay. Coastal Eng. J. 63(4): 446–466. DOI: https://doi.org/10.1080/21664250.2021.1933749

Nalbant SS, Steacy S, Sieh K, Natwidjaja D, McCloskey J (2005) Earthquake risk on the Sunda trench. Nature 435: 756–757. DOI: https://doi.org/10.1038/nature435756a

Niu X, Zhou H (2015) Wave pattern induced by a moving atmospheric pressure disturbance. Appl. Ocean Res. 52: 37–42. DOI: https://doi.org/10.1016/j.apor.2015.05.003

Niu X, Chen Y (2020) Energy accumulation during the growth of forced wave induced by a moving atmospheric pressure disturbance. Coastal Eng. J. 62(1): 23–34. DOI: https://doi.org/10.1080/21664250.2019.1682747

Oetjen J, Engel M, Schuttrumpf H (2020) Experiments on tsunami induced boulder transport — A review. Earth-Sci. Rev. 220: 103714. DOI: https://doi.org/10.1016/j.earscirev.2021.103714

Okal EA, Fritz HM, Raad PE, Synolakis C, Al-Shijbi Y, Al-Saifi M (2006) Oman field survey after the December 2004 Indian Ocean Tsunami. Earthquake Spectra 22 (3_suppl.): 203–218. DOI: https://doi.org/10.1193/1.2202647

Onozato M, Nishigaki A, Okoshi K (2016) Polycyclic aromatic hydrocarbons in sediments and bivalves on the Pacific Coast of Japan: Influence of tsunami and fire. PLoS ONE 11(5): e0156447. DOI: https://doi.org/10.1371/journal.pone.0156447

Otsuka S (2022) Visualizing Lamb waves from a volcanic eruption using meteorological satellite Himawari-8. Geophys. Res. Lett. 49: e2022GL098324. DOI: https://doi.org/10.1029/2022GL098324

Ozawa S, Nishimura T, Suito H, Kobayashi T, Tobita M, Imakiire T (2011) Coseismic and postseismic slip of the 2011 magnitude-9 Tohoku-Oki earthquake. Nature 475: 373–376. DOI: https://doi.org/10.1038/nature10227

Paris R, Switzer AD, Belousova M, Belousov A, Ontowirjo B, Whelley PL, Ulvrova M (2014) Volcanic tsunami: a review of source mechanisms, past events and hazards in Southeast Asia (Indonesia, Philippines, Papua New Guinea). Nat. Hazards. 70: 447–470. DOI: https://doi.org/10.1007/s11069-013-0822-8

Pattiaratchi CB, Wijeratne EMS (2015) Are meteotsunamis an underrated hazard? Proc. R. Soc. A 373: 20140377. DOI: https://doi.org/10.1098/rsta.2014.0377

Ramsden JD (1996) Forces on a vertical wall due to long waves, bores, and dry-bed surges. J Waterw. Port, Coast. Ocean Eng. 122(3): 134–141. DOI: https://doi.org/10.1061/(ASCE)0733-950X(1996)122:3(134)

Ren Z, Zhao X, Liu H (2019) Numerical study of the landslide tsunami in the South China Sea using Herschel-Bulkley rheological theory. Phys. Fluids 31: 056601. DOI: https://doi.org/10.1063/1.5087245

Sandanbata O, Watada S, Satake K, Fukao Y, Sugioka H, Ito A, Shiobara H (2018) Ray tracing for dispersive tsunamis and source amplitude estimation based on Green’s Law: Application to the 2015 Volcanic Tsunami Earthquake Near Torishima, South of Japan. Pure. Appl. Geophys. 175: 1371–1385. DOI: https://doi.org/10.1007/s00024-017-1746-0

Satake K, Bourgeois J, Abe K, Abke K, Tsuji Y, Imamura F, Lio Y, Katao H, Noguera E, Estrada F (1993) Tsunami field survey of the 1992 Nicaragua earthquake. EoS 74(13): 145–157. DOI: https://doi.org/10.1029/93EO00271

Satake K, Fujii Y, Harada T, Namegaya Y (2013) Time and space distribution of coseismic slip of the 2011 Tohoku earthquake as inferred from tsunami waveform data. Bull. Seismol. Soc. Am. 103(2B): 1473–1492. DOI: https://doi.org/10.1785/0120120122

Sato M, Ishikawa T, Ujihara N, Yoshida S, Fujita M, Asada A (2011) Displacement above the hypocenter of the 2011 Tohoku-Oki Earthquake. Science 332(6036): 1395. DOI: https://doi.org/10.1126/science.1207401

Sato S (1996) Numerical simulation of 1993 Southwest Hokkaido Earthquake Tsunami around Okushiri Island. J Waterw. Port, Coast. Ocean Eng. 122(5): 209–215. DOI: https://doi.org/10.1061/(ASCE)0733-950X(1996)122:5(209)

Sato S, Liu H, Takewaka S, Nobuoka H, Aoki S (2012) Tsunami damages of Nakoso Coast due to the 2011 Tohoku Earthquake. Proceedings of 33rd International Conference on Coastal Engineering, ICCE2012, Santander, Spain. DOI: https://doi.org/10.9753/icce.v33.currents.2

Schimmels S, Sriram V, Didenkulova I (2016) Tsunami generation in a large scale experimental facility. Coastal Eng. 110: 32–41. DOI: https://doi.org/10.1016/j.coastaleng.2015.12.005

Shen J, Wei L, Wu D, Liu H, Huangfu J (2020) Spatiotemporal characteristics of the dam-break induced surge pressure on a vertical wall. Coastal Eng. J. 62(4): 566–581. DOI: https://doi.org/10.1080/21664250.2020.1828559

Shen J, Liu H (2022) On the structure dynamic response of a coastal structure subject to the dam break induced surge impact pressure. Coastal Eng. J. 64(2): 246–259. DOI: https://doi.org/10.1080/21664250.2021.2006950

Shimozono T, Sato S, Okayasu Y, Tajima Y, Fritz HM, Liu H, Takagawa T (2012) Propagation and inundation characteristics of the 2011 Tohoku tsunami on the central Sanriku Coast. Coastal Eng. J. 54(1): 1250004. DOI: https://doi.org/10.1142/S0578563412500040

Shuto N (1991) Numerical simulation of tsunamis — Its present and near future. Nat. Hazards 4: 171–191. DOI: https://doi.org/10.1007/BF00162786

Simkin T, Fiske RS (1983) Krakatau 1883: the volcanic eruption and its effects. Smithsonian Institution Press, Washington, D.C.

Siripong A (2006) Andaman Seacoast of Thailand field survey after the December 2004 Indian Ocean Tsunami. Earthquake Spectra 22 (3_suppl.): 187–202. DOI: https://doi.org/10.1193/1.2209927

Sugawara D (2021) Numerical modeling of tsunami: advances and future challenges after the 2011 Tohoku earthquake and tsunami. Earth-Sci. Rev. 214: 103498. DOI: https://doi.org/10.1016/j.earscirev.2020.103498

Synolakis CE (1987) The runup of solitary waves. J. Fluid Mech. 185: 523–545. DOI: https://doi.org/10.1017/S002211208700329X

Takagi H, Pratama MB, Kurobe S, Esteban M, Aranguiz R, Ke B (2019) Analysis of generation and arrival time of landslide tsunami to Palu City due to the 2018 Sulawesi earthquake. Landslides 16: 983–991. DOI: https://doi.org/10.1007/s10346-019-01166-y

Tappin DR, Grilli ST, Harris JC, Geller RJ, Masterlark T, Kirby JT, Shi F, Ma M, Thingbaijam KKS, Mai PM (2014) Did a submarine landslide contribute to the 2011 Tohoku tsunami? Mar. Geol. 357: 344–361. DOI: https://doi.org/10.1016/j.margeo.2014.09.043

Terry JP, Winspear N, Goff J, Tan PHH (2017) Past and potential tsunami sources in the South China Sea: A brief synthesis. Earth-Sci. Rev. 167: 47–61. DOI: https://doi.org/10.1016/j.earscirev.2017.02.007

Titov V, Rabinovich AB, Mofjeld HO, Thomson RE, Gonzalez FI (2005) The global reach of the 26 December 2004 Sumatra tsunami. Science 309: 2045–2048. DOI: https://doi.org/10.1126/science.1114576

Titov VV, Gonzalez FI (1997) Implementation and testing of the method of splitting tsunami (MOST) model. NOAA Technical Memorandum ERL PMEL-112, Contribution No. 1927 from NOAA/Pacific Marine Environmental Laboratory

Vilibić I, Rabinovich AB, Anderson EJ (2021) Special issue on the global perspective on meteotsunami science: editorial. Nat. Hazards 106: 1087–1104. DOI: https://doi.org/10.1007/s11069-021-04679-9

Wang G, Liang Q, Shi F, Zheng J (2021) Analytical and numerical investigation of trapped ocean waves along a submerged ridge. J. Fluid Mech. 915: A54. DOI: https://doi.org/10.1017/jfm.2020.1039

Wang X (2009) User manual for COMCOT version 1.7. Edited by Liu PLF, Woo SB, Cho YS, Computer Programs for Tsunami Propagation and Inundation, Cornel University

Ward SN, Day S (2001) Cumbre Vieja Volcano—Potential collapse and tsunami at La Palma, Canary Islands. Geophys. Res. Lett. 28(17): 3397–3400. DOI: https://doi.org/10.1029/2001GL013110

Wu D, Liu H (2022) Effects of the bed roughness and beach slope on the non-breaking solitary wave runup height. Coastal Eng. 174: 104122. DOI: https://doi.org/10.1016/j.coastaleng.2022.104122

Wuthrich D, Pfister M, Nistor I, Schleiss AJ (2018) Experimental study of tsunami-like waves generated with a vertical release technique on dry and wet beds. J Waterw. Port, Coast. Ocean Eng. 144(4): 04018006. DOI: https://doi.org/10.1061/(ASCE)WW.1943-5460.0000447

Xie W, Shimozono T (2022) Water surge impingement onto a vertical wall: Laboratory experiments and stochastic analysis on impact pressure. Ocean Eng. 248: 110422. DOI: https://doi.org/10.1016/j.oceaneng.2021.110422

Yamada M, Ho TC, Mori J, Nishikawa Y, Yamamoto M (2022) Tsunami triggered by the Lamb wave from the 2022 Tonga volcanic eruption and transition in the offshore Japan region. Geophys. Res. Lett. 49(15): 9. DOI: https://doi.org/10.1029/2022GL098752

Yang M, Zheng Y, Liu H (2022) Experimental study of the solitary wave induced groundwater hydrodynamics. Coastal Eng. 177: 104193. DOI: https://doi.org/10.1007/s13344-019-0025-5

Ye L, Kanamori H, Rivera L, Lay T, Zhou Y, Sianipar D, Satake K (2020) The 22 December 2018 tsunami from flank collapse of Anak Krakatau volcano during eruption. Sci. Adv. 6: eaaz1377. DOI: https://doi.org/10.1126/sciadv.aaz1377

Yeh H, Liu P, Briggs M, Synolakis C (1994) Propagation and amplification of tsunamis at coastal boundaries. Nature 372: 353–355. DOI: https://doi.org/10.1038/372353a0

Yeh H, Chadha RK, Francis M, Katada T, Latha G, Peterson C, Raghuraman G, Singh JP (2006) Tsunami runup survey along the Southeast Indian Coast. Earthquake Spectra 22 (3_suppl.): 173–186. DOI: https://doi.org/10.1193/1.2202651

Yeh H, Sato S, Tajima Y (2013) The 11 March 2011 East Japan Earthquake and Tsunami: Tsunami effects on coastal infrastructure and buildings. Pure Appl. Geophys. 170: 1019–1031. DOI: https://doi.org/10.1007/s00024-012-0489-1

Yulianto E, Utari P, Satyawan IA (2020) Communication technology support in disaster-prone areas: Case study of earthquake, tsunami and liquefaction in Palu, Indonesia. Int. J. Disast. Risk Res. 45: 101457. DOI: https://doi.org/10.1016/j.ijdrr.2019.101457

Zeng J, Liu H (2022) An approximate explicit analytical solution for the frictionless swash hydrodynamics with an improved seaward boundary condition. Coastal Eng. 174: 104127. DOI: https://doi.org/10.1016/j.coastaleng.2022.104127

Zhao B, Duan W, Webster WC (2011) Tsunami simulation with Green—Naghdi theory. Coastal Eng. 38: 389–396. DOI: https://doi.org/10.1016/j.oceaneng.2010.11.008

Zhang X, Niu X (2020) Probabilistic tsunami hazard assessment and its application to southeast coast of Hainan Island from Manila Trench. Coastal Eng. 155: 103596. DOI: https://doi.org/10.1016/j.coastaleng.2019.103596

Zhang Y, Liu H (2021) Spatiotemporal variation of wave energy induced by an accelerated moving atmospheric pressure disturbance. Coastal Eng. J. 63(1): 83–91. DOI: https://doi.org/10.1080/21664250.2021.1875690

Zhang Y, Liu H (2022) Generation mechanisms of the water surface elevation induced by a moving atmospheric pressure disturbance. Ocean Eng. 255: 111469. DOI: https://doi.org/10.1007/978-981-15-0291-0_19

Zitellini N, Mendes LA, Cordoba D, Danobeitia J, Nicolich R, Ribeiro GPA, Sartori R, Torelli L, Bartolome R, Bortoluzzi G, Calafato A, Carrilho F, Casoni L, Chierici F, Corela C, Correggiari A, Delia Vedova B, Gracia E, Jornet P, Landuzzi M, Ligi M, Magagnoli A, Marozzi G, Matias L, Penitenti D, Rodriguez P, Rovere M, Terrinha P, Vigliotti L, Ruiz AZ (2001) Source of 1755 Lisbon Earthquake and Tsunami investigated. EoS 82(26): 285–296. DOI: https://doi.org/10.1029/EO082i026p00285-01

Download references

Supported by the National Natural Science Foundation of China under Grant Nos. 52271292, 52071288; the Science and Technology Innovation 2025 Major Project of Ningbo City under Grant No. 2022Z213.

Author information

Authors and affiliations.

Faculty of Mechanical Engineering and Mechanics, Ningbo University, Ningbo, 315211, China

Sixue Cheng

College of Civil Engineering and Architecture, Zhejiang University, Hangzhou, 310058, China

Sixue Cheng & Haijiang Liu

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Haijiang Liu .

Additional information

Article highlights.

• Characteristics of the tsunamis induced by four different mechanisms are individually specified.

• Tsunami research from three different approaches is comprehensively reviewed and summarized.

• From the direct post-tsunami field survey to the complicated numerical and physical models, understanding of tsunami’s sophisticated mechanism is gradually extended.

• Great research efforts with interdisciplinary cooperation is appealed for the tsunamis simultaneously triggered by multiple sources.

Rights and permissions

Reprints and permissions

About this article

Cheng, S., Liu, H. General Review of the Worldwide Tsunami Research. J. Marine. Sci. Appl. 22 , 14–24 (2023). https://doi.org/10.1007/s11804-023-00315-z

Download citation

Received : 15 August 2022

Accepted : 28 November 2022

Published : 15 April 2023

Issue Date : March 2023

DOI : https://doi.org/10.1007/s11804-023-00315-z

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Earthquake induced tsunami
  • Landslide induced tsunami
  • Volcano eruption induced tsunami
  • Meteorological tsunami
  • Post-tsunami field survey
  • Numerical modeling
  • Laboratory experiment
  • Find a journal
  • Publish with us
  • Track your research

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • My Account Login
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 15 April 2021

Early forecasting of tsunami inundation from tsunami and geodetic observation data with convolutional neural networks

  • Fumiyasu Makinoshima   ORCID: orcid.org/0000-0001-9247-4104 1 ,
  • Yusuke Oishi   ORCID: orcid.org/0000-0003-4264-8932 1 ,
  • Takashi Yamazaki 1 ,
  • Takashi Furumura   ORCID: orcid.org/0000-0002-2091-0533 2 &
  • Fumihiko Imamura   ORCID: orcid.org/0000-0001-7628-575X 3  

Nature Communications volume  12 , Article number:  2253 ( 2021 ) Cite this article

9127 Accesses

31 Citations

10 Altmetric

Metrics details

  • Natural hazards

Rapid and accurate hazard forecasting is important for prompt evacuations and reducing casualties during natural disasters. In the decade since the 2011 Tohoku tsunami, various tsunami forecasting methods using real-time data have been proposed. However, rapid and accurate tsunami inundation forecasting in coastal areas remains challenging. Here, we propose a tsunami forecasting approach using convolutional neural networks (CNNs) for early warning. Numerical tsunami forecasting experiments for Tohoku demonstrated excellent performance with average maximum tsunami amplitude and tsunami arrival time forecasting errors of ~0.4 m and ~48 s, respectively, for 1,000 unknown synthetic tsunami scenarios. Our forecasting approach required only 0.004 s on average using a single CPU node. Moreover, the CNN trained on only synthetic tsunami scenarios provided reasonable inundation forecasts using actual observation data from the 2011 event, even with noisy inputs. These results verify the feasibility of AI-enabled tsunami forecasting for providing rapid and accurate early warnings.

Introduction

The rapid forecasting of hazards and dissemination of warnings can increase evacuation lead times and thus are key to saving lives during natural disasters. For tsunami disasters, quick evacuations supported by such warnings can drastically reduce the number of casualties, but inaccurate hazard forecasts and warnings can have the opposite effect. During the 2011 Tohoku tsunami event, the earthquake magnitude and tsunami height were initially underestimated based on a precomputed database; as a result, some residents felt safe based on the initial warning and were unaware of the need for evacuation 1 . Although the warning was updated several times based on additional observations, the updated information could not always reach the residents due to communication disruptions; consequently, many of the coastal residents did not realise the tsunami risk at their locations. The underestimation of tsunami risk increased the number of tsunami-induced casualties, and as a result, Japan experienced the loss of over 18,000 citizens, even with the in-place warning system. The need for accurate and reliable early warnings has been common during past mega-tsunamis; notably, the 2004 Indian Ocean tsunami had a catastrophic regional impact 2 , and the immense loss of nearly 230,000 lives stressed the importance of tsunami early warning systems, resulting in efforts to establish tsunami early warning frameworks in broader regions 3 . Therefore, fast and accurate tsunami forecasting methods based on real-time tsunami observation data are urgently needed, and the early warnings provided can contribute to mitigating casualties in future tsunami events.

To date, tsunami early warning systems have been developed based on past tsunami catastrophes and available technologies 4 . Especially in the decade since the 2011 Tohoku tsunami, dense tsunami observation networks have been implemented 5 , 6 , and various tsunami forecasting methods using real-time observation data, such as real-time tsunami inundation simulations using supercomputers 7 with rapid source estimations 8 , 9 and data assimilation approaches 10 , 11 , have been proposed based on the lessons learned from the 2011 event. However, real-time tsunami inundation forecasting immediately after an earthquake has remained challenging due to the difficulty of rapid estimation of the tsunami source, in which various uncertainties exist 12 , and due to the high computational costs associated with simulating nonlinear tsunami propagation in shallow water.

To overcome the above challenges, we present a tsunami forecasting method using a convolutional neural network (CNN) developed as a deep learning approach in AI research. The present method is capable of directly forecasting tsunami inundation based solely on up-to-date observation data and does not require extensive computational resources, such as those provided by supercomputers. During the past decade, deep learning has achieved great success in image and pattern recognition 13 as well as in broader areas, including physics-based simulations such as structural analysis 14 and computational fluid dynamics 15 ; moreover, tsunami observation networks have been improved.

In this work, we utilise a CNN to process valuable data from dense tsunami and geodetic observation networks and achieve remarkable tsunami forecasting performance. A notable advantage of a CNN is its low computational cost; i.e., the computational cost of CNN inference is much lower than that of nonlinear tsunami propagation simulations. Additionally, the present approach does not require a tsunami source estimation process since our CNN is designed for end-to-end forecasting from observation data to tsunami inundation forecasting. Therefore, the present CNN can immediately and accurately predict the tsunami inundation time series at a single location. To the best of our knowledge, this study is the first attempt at end-to-end tsunami inundation forecasting with a CNN, and the results verify the feasibility of AI-enabled tsunami forecasting for the establishment of early warnings.

CNN tsunami forecasting

Figure  1 shows a schematic view of the proposed tsunami forecasting method based on a CNN. First, 10,000 cases of numerical tsunami propagation and inundation simulations solving a set of partial differential equations were conducted to prepare the data sets for training the CNN. In the simulations, sets of synthetic observation data at observation points and the resulting tsunami inundation waveforms within 2 h were calculated based on randomly generated tsunami scenarios. For the tsunami observations obtained with ocean bottom pressure gauges, simulated tsunami waveforms converted into pressure waveforms at the ocean bottom were used as inputs. Then, we trained a CNN with synthetic tsunami data to directly predict the tsunami inundation waveform at a single onshore location (the green star in Fig.  1 ) solely from the observations. We built a 1-D CNN comprising a total of 15 layers for tsunami forecasting since compact 1-D CNNs are suited for real-time and low-cost applications because of their low computational complexity 16 . The stacked waveforms were fed into the network as inputs, and the corresponding features were extracted via convolutional and pooling-like layers. To consider additional observations as inputs for the network, we simply stacked same-size arrays so that the size of the input channels equalled the considered number of observation points. Since recent studies on tsunami source inversion suggest that inland geodetic observational data are useful for determining the tsunami source 8 , 9 , 17 , we also considered inland geodetic observations of initial ground heights as additional inputs to the CNN. The geodetic observations were fed as vectors with the same length as the offshore waveforms. The onshore tsunami inundation waveform was then forecasted by the fully connected layers using the features extracted from the convolutional layers. In contrast to general time series forecasting using deep learning, in which the subsequent transition of values at future time t  +  h is predicted based on the available observations at time t in the same series 18 , our network predicts a future onshore tsunami inundation waveform based on the available observations at different offshore and onshore points. After the training process, the trained CNN can output a tsunami inundation waveform at a single location on land (the green star in Fig.  1 ) when observation data for an unknown tsunami are given. Preparing multiple networks enables tsunami forecasting for multiple points.

figure 1

The CNN learns the relation between observation data and the resulting tsunami waveform from thousands of numerical simulation results. The tsunami and geodetic observation points are uniformly selected to cover the forecasting areas and are illustrated as red points in the map. After the training process is completed, the trained CNN can predict a time series of inundation tsunami waveform at the onshore forecasting site, which is represented with a green star, solely from observation data for unknown tsunami scenarios.

Data preparation for the CNN

The source fault geometry of the 2011 Tohoku-Oki earthquake was considered based on the parameters presented in Fujii et al. 19 , and 44 sub-faults were used to generate synthetic tsunami data sets for the training of CNN (Fig.  2a ). Here, the slip parameters of each sub-fault were randomly assigned within their defined range for the sub-faults to generate various tsunami scenarios. We conducted tsunami simulations for a total of 12,000 scenarios and obtained observations for inputs (red points in Fig.  2a ) and an onshore tsunami inundation waveform at a single location (the green star in Fig.  2b ) to train the CNN. A total of 49 offshore tsunami observation points and five onshore geodetic observation points were considered as inputs for the CNN based on an actual tsunami observation network 20 and the Global Navigation Satellite System observation network 21 in Japan. These tsunami and geodetic observation points were selected to cover the fault region and the forecasting point. The offshore tsunami observation waveforms were sampled at 1 Hz, and a constant deformed ground height with the same number of samples was considered as the onshore geodetic observations. The forecasting waveform was sampled at 0.5 Hz with a data size of 3600. For the observation data, various time windows of tsunami observations (e.g., 5, 10, 15, 20, 25, and 30 min) were considered as inputs for the CNN to investigate the effect of the observation length on the forecasting accuracy. From the 12,000 simulation results, we used 10,000 cases for training, 1000 cases for validation monitoring, and 1000 for testing. The generated earthquake scenarios have a seismic moment M 0 ranging between 4.03 × 10 22 and 8.21 × 10 22  Nm, which corresponds to a moment magnitude M w ranging from 9.0 to 9.2, assuming a rigidity of 30 GPa (Fig.  2c ). The initial sea-bottom deformation caused by an earthquake was calculated with Okada’s formula 22 , and the tsunami propagation and inundation were simulated using TUNAMI-N2 code (see “Methods” for details). A single tsunami simulation for 2 h to generate data for the CNN required ~3 h using a CPU node with two Intel Xeon Gold 6148 processors with 384 GiB memory.

figure 2

a Fault geometry and observation point as input for convolutional neural network (CNN). Red circles represent observation points for CNN. A total of 49 offshore tsunami observation points and five onshore geodetic observation points were considered to cover the fault region and the forecasting point (the green star). Fault geometry is considered based on Fujii et al. 19 , but its slip amount is randomly assigned within a given range to generate various tsunami scenarios. b A close view of the forecasting site, the Sendai plain where experienced the large extent of tsunami inundation. The green star is the forecasting site for which the CNN forecast a tsunami inundation waveform. c Distribution of the seismic moment of generated earthquake scenarios. A total of 12,000 scenarios were generated and divided into 10,000 training sets, 1000 validation sets and 1000 test sets.

Network configuration and training

The constructed CNN consisted of nine convolutional layers and three pooling-like convolutional layers followed by three fully connected layers (Supplementary Table  1 ). Each convolutional layer was followed by a Leaky ReLU activation function 23 with a negative slope parameter of a  = 0.01. Dropout with a dropout ratio of p  = 0.5 was applied to the outputs of the fully connected layers to prevent overfitting 24 . In the convolutional layers, we set the kernel size, stride and padding to minimise changes in the array size, and dimensionality reduction was performed mainly by the pooling-like layers. For the 5, 15, 25 and 35 min observation cases, the kernel size of the last convolutional layer was set as 3 to prevent the remainder from being generated, but such changes were minimised to evaluate the effect of the observation lead time. Thus, a longer observation period leads to a greater number of learnable parameters in the network.

We considered the mean squared error (MSE) between the simulated and forecasting waveforms over the forecasting time as the loss function. The network parameters were optimised to minimise the loss function by the Adam optimisation algorithm 25 . We used the default parameters suggested in the original paper that proposed the Adam algorithm ( β 1  = 0.9, β 2  = 0.999 and ε  = 10 −8 ), except for a step size of α  = 10 −4 . The training of the network was performed in the AI Bridging Cloud Infrastructure, a GPU-accelerated supercomputer in which each node has two Intel Xeon Gold 6148 CPUs and four NVIDIA Tesla V100 SXM2 GPUs with 384 GiB memory. The network training and validation in this study were implemented using Pytorch 26 and Horovod 27 . The batch size for each GPU during the training phase was 25, and five computer nodes were used for training. We considered 3000 epochs in the training process and retained the model that yielded the minimum validation loss for the validation data sets during training. The training process was completed within 2 h even for the largest network in this study.

Forecasting performance on synthetic tsunamis

We evaluated the performance of the trained CNN by analysing the forecasting results for 1000 test tsunami scenarios that were not considered during the training process, and we confirmed that the CNN successfully predicted the tsunami inundation waveform at the forecasting site (Fig.  3b ). For the evaluation of the forecasting accuracy, we considered two metrics: the maximum tsunami amplitude, defined as the maximum value of the tsunami amplitude over the forecasting time, and the tsunami arrival time, defined as the time when the tsunami flow depth first exceeds 10% of the maximum flow depth. Even with only 5 min offshore tsunami and inland geodetic observations, the mean absolute errors of the maximum tsunami amplitude and the tsunami arrival time were 0.4 m and 47.7 s, respectively. The average relative errors were 8.1% and 1.2% for the maximum tsunami amplitude and the tsunami arrival time, respectively. For these forecasts, the trained CNN required only 0.004 s on average using a single CPU node with 40 cores; this approach is much faster than conventional simulation-based forecasting approaches and requires fewer computational resources. The trained CNN yielded accurate and rapid forecasts of both the tsunami size and the arrival time for various tsunami scenarios directly from the observations.

figure 3

a Observation points and a snapshot of tsunami propagation at 300 s for a simulated test scenario. The location of the observation points as inputs for the CNN are illustrated as circles. The forecasting site is represented with the green star. b Result of tsunami inundation forecasting with the CNN at the forecasting site with the ground-truth simulation result. c Error distribution of the maximum tsunami amplitude at the forecasting site for 1000 test scenarios.

Effects of the offshore observation length and geodetic data

We investigated the effect of the length of the observation inputs (5, 10, 15, 20, 25 and 30 min) and the importance of geodetic observation inputs by training different CNN models with different inputs (Fig.  4 ). The results show that longer observations led to higher forecasting accuracy. Using the geodetic observation data, the CNN achieved good forecasting performance equivalent to that of a CNN with a long observation period. We confirmed that this performance improvement achieved with comparatively long-term observations and geodetic data corresponds mainly to increased accuracy in the initial ground height estimation in which the CNN with a short observation length and without geodetic data exhibited poor accuracy (Supplementary Fig.  1 ). This result can be explained by the characteristics of tsunami long wavelength. Since the propagation speed of a tsunami is slow in shallow water, it is difficult to obtain the waveforms generated from nearshore faults that cause coastal subsidence with a short observation length; thus, short-term observations alone are not sufficient for estimating the magnitude of subsidence. Inverting the offshore fault slip distribution from onshore geodetic information can lead to non-unique solutions; however, geodetic data offer direct information about subsidence much faster than offshore observations. The proposed CNN integrated different information types from different observations and achieved the presented forecasting performance, even with very short observation periods.

figure 4

The height of the bar represents the mean value for 1000 test scenarios, and the error bar in the figure represents the standard deviation.

Sensitivity of offshore and onshore observations

To understand information processing in CNN tsunami forecasting models, we conducted a sensitivity analysis (e.g., occlusion test 28 , 29 ) of the trained CNN models. In this analysis, we systematically removed inputs from certain observation points and evaluated the resulting amounts of change in the forecasting results to represent the impact on the CNN. The sensitivity analysis was conducted for models with different observation lengths, and the sensitivity of the observation points for forecasting was visualised (Fig.  5 ). The observation points with high sensitivities were located in the specific region of the observation network along the major path of tsunamis towards the forecasting site. High sensitivities were observed mainly over large slip areas since the amount of slip on an offshore fault has a predominant influence on tsunami inundation. In contrast, the information from distant observation points had almost no effect on the forecast, and thus, this information was not important for the CNN. This result indicates that for an accurate forecast, the CNN requires only certain observation points along the path of a tsunami propagating to a forecasting site. As the observation time increases, high sensitivities can also be confirmed at nearshore tsunami observation points. The CNNs using both offshore tsunami and onshore geodetic observations showed high sensitivities at both onshore and offshore points. The increased sensitivities for nearshore tsunami observations with longer observation periods and the higher sensitivities for additional onshore geodetic observations suggest that the CNN effectively integrates available information within limited observation periods to achieve high forecasting accuracy.

figure 5

a CNN trained only with offshore tsunami observation data. b CNN trained with both offshore tsunami and onshore geodetic observation data. The observation points with high sensitivity indicate that the forecasting result changes considerably when the inputs from these observation points are lacking.

Forecasting speed

The computational time for tsunami forecasting with the CNN was measured to investigate the forecasting speed and assess the ability of the method to be employed for the issuance of tsunami warnings. Here, we measured the time required to forecast 1000 test scenarios using a single CPU node with 40 cores. Table  1 reports the average computational time for tsunami forecasting. The computational time increases as the observation time increases since a consistent CNN architecture is adopted for all observation lengths; a large neural network structure and a corresponding increase in the number of parameters are needed for longer observation times. The addition of 5 min observations increased the number of parameters by ~13 million, mainly due to the larger size of the fully connected layer after the convolutional layers. Nevertheless, the computational time for tsunami forecasting with the CNN was only 0.011 s, even for the largest CNN settings in the test (30 min offshore tsunami observations with geodetic observations). The addition of five geodetic input channels slightly increased the number of parameters by 1920 but had a negligible effect on the computational time. Thus, for CNN tsunami forecasting, the use of geodetic data was effective from the perspectives of not only the forecasting accuracy but also the computational time required to consider additional inputs. The tsunami forecasting speed achieved by the CNN is sufficiently fast to provide tsunami warnings, even with limited computational resources.

Application to the 2011 Tohoku tsunami event

We trained the CNNs using the 10,000 synthetic tsunami scenarios with the observation settings at the time the 2011 Tohoku tsunami event occurred and investigated the forecasting performance of the CNN for real events using real-world data (Fig.  6a ). This application used the same network configuration and the same 10,000 tsunami scenarios employed in the previous tests using synthetic data. We used publicly available observation data 30 , 31 during the event as inputs for the CNN (Fig.  6b, c ), i.e., three offshore tsunami observations recorded by GPS buoys (803, 801 and 806) and three onshore GNSS observations (Rifu, Watari and Souma1). For the missing parts in the tsunami waveforms observed by GPS buoys, 1 Hz data were prepared by cubic interpolation. For the geodetic observations, displacements at 5 min after the occurrence of the earthquake were used as inputs. Since complete data were not available at Souma1, we used the latest observation as the input. The forecasting site is shown as the green star in Fig.  6a . Most of the buildings around the forecasting site were totally destroyed by the tsunami; however, Arahama Elementary School (Arahama ES, illustrated as the black cross in Fig.  6a ) located close to the forecasting site provided survey results to verify the inundation forecast. During the 2011 event, the tsunami reached the second floor of the Arahama ES 32 , and a survey immediately after the event reported that tsunami debris were found at a height of 4.62 m above the school basement 33 . The arrival time of the devastating tsunami at this site was also estimated as ~15:55 (JST) based on a stopped clock 34 .

figure 6

The elapsed time is from the earthquake occurrence. a Observation points for the CNN and the forecasting site. The red triangles represent the offshore tsunami observation points (803, 801, 806), and the red circles represent onshore geodetic observation points (Rifu, Watari, Souma1). The red cross mark in the small-scale map represents the epicentre of the 2011 earthquake. The forecasting site is shown in the large-scale map as the green star. Arahama ES (the black cross mark) where tsunami inundation traces for validation are available is located close to the forecasting site. b Tsunami waveforms at 803, 801 and 806 observed during the 2011 event 31 . c Geodetic observations at Rifu, Watari and Souma1 observed during the 2011 event. d Results of tsunami inundation forecasting with different observation periods. Grey triangles represent the observation interval used for forecast. Red waveforms are the CNN forecasting results. The height and position of the black cross mark represents the surveyed tsunami inundation trace at Arahama ES. e Results of offshore tsunami waveform forecasting at the Sendai New Port with different observation periods. Grey triangles represent the observation interval used for forecast. Red waveforms are CNN forecasting results. Grey lines are observed waveforms at the Sendai New Port during the 2011 event. Full observation was not available because the gauge was destroyed by the tsunami.

We trained the CNNs using only synthetic tsunami data with different observation periods and forecasted the tsunami inundation waveform using the actual observation data as inputs for the CNNs (Fig.  6d ). Initially, the CNNs forecasted unphysical waveforms with 20 min or less observations, when almost no tsunami signals were available. After obtaining the first positive peak of the tsunami from 30 min observations, the CNN forecasted an inundation waveform, but the forecasted amplitude was small compared to the actual trace at Arahama ES. With the 35 min observations, in which the entire first positive peak of the tsunami is available, the CNN forecasted a maximum flow depth of 3.88 m at 3974 s after the earthquake. Further 40 min observation revealed the negative peak of the tsunami, enabling the forecast of a larger inundation depth (5.64 m at 3952 s). Although the reported tsunami traces do not directly reflect the actual maximum flow depth, the CNN forecasting can be considered reasonably accurate. However, the forecasted arrival time of the maximum flow depth was ~3 min earlier than the estimated arrival time indicated by the stopped clock at the Arahama ES.

To further verify the CNN forecasting result, we trained the CNN for forecasting a tsunami waveform at the Sendai New Port where a time series of the tsunami until the first peak was recorded by a wave gauge (subsequent observations were not available because the gauge was destroyed by the tsunami). The forecasting results for the Sendai New Port are summarised in Fig. 6e . Similar to the previous inundation forecasting results, the CNN could not provide a reasonable forecast with 20 min or less observation time when sufficient tsunami signals were not available. After obtaining the positive peak of the tsunami using 30 min observation, the CNN could forecast the tsunami having the first peak value of 6.78 m, which is compatible with the observed first peak (6.62 m). The CNNs trained with 35 and 40 min observations forecasted similar peaks (5.77 and 6.42 m, respectively), and the rise of the first wave was more consistent with the observation data. Nevertheless, even with sufficient offshore tsunami observations, an ~3 min difference between the observation and the forecast appeared in the tsunami arrival time, suggesting that the CNN forecast tends to be slightly earlier than the actual tsunami arrival. A possible cause of this earlier arrival tendency is the effect of rupture propagations, i.e., when generating tsunami data sets for the CNN, the effect of rupture delays on each sub-fault was not considered, and instantaneous slip was assumed; however, the tsunami source inversion assuming sub-faults with multiple time windows suggested that the duration of the tsunamigenic slip of the 2011 Mw 9.0 earthquake lasted ~2.5–3 min 35 , and this duration was consistent with the difference in the arrival time. The current CNN forecast provides a slightly earlier arrival tendency and might serve as a cautious warning; however, to forecast the tsunami arrival time more accurately for large earthquakes, it may be necessary to consider the effect of rupture propagation on faults in large earthquakes.

Recent high-sampling-rate tsunami observations, especially by ocean bottom pressure gauges, can capture a wide range of geophysical phenomena in the ocean including ocean currents and seismic waves with much shorter periods (seconds to minutes) than that of tsunamis (minutes to hours) 36 . Consequently, such non-tsunami components can affect tsunami forecasting as noise 37 . To investigate the performance of the CNN under actual observation conditions, we further evaluated the effect of the short-period observation noise on the CNN tsunami forecasting. Noise waveforms were obtained using actual sea-level observation data at the GPS buoys 803, 801 and 806 a day before the 2011 tsunami event (Fig.  7a, b ), and the effect of noisy tsunami input on the forecasting results was evaluated using 40 min observation data (Fig.  7c–e ). For this evaluation, the similarity between the forecasting results with and without noise were evaluated using the formula for calculating the variance reduction 38 ; therefore, the similarity becomes 100% for a perfect match and lower for misfits. The result demonstrated that even with disturbances, the CNN successfully forecasted both the inundation and offshore waveforms with a small difference. The similarity between forecasting result with and without noise was 99.999% for inundation forecasting and 99.997% for offshore waveform forecasting. Additionally, we also examined the effect of much larger noise on the inundation forecasts. Larger noise waveforms were generated by adding white noise that ranged from −1.0 to 1.0 and was amplified by a certain ratio of the maximum observation amplitude. The additional noise test demonstrated that the CNN tsunami forecast maintained a high similarity of 99.7% on average for 1000 noisy inputs, even with a large noise level of 20% (see Supplementary Fig.  2 ).

figure 7

a Sea-level observations from GPS buoys before the 2011 tsunami event. The grey and red lines represent the raw observations and the 15 min running average of the raw data, respectively. The elapsed time is from a day before the earthquake occurrence. b Noise waveforms extracted from the observations. The deviation from the averaged data (the red lines in a ) is extracted as noise. The figure shows the noise waveforms during the same observation period for the actual tsunami observations, but a day before the tsunami. c Noise waveforms for CNN input. The extracted noise waveforms before the event are added to the actual tsunami observations to prepare the noise inputs. d Inundation forecasting results with and without noise. The grey and red lines represent the forecasting results without and with noise, respectively. Because the difference between the two waveforms is small, the result without noise is illustrated with a thicker line. e Offshore tsunami forecasting results with and without noise. The grey and red lines represent the forecasting results without and with noise, respectively. Because the difference between the two waveforms is small, the result without noise is illustrated with a thicker line.

In this study, we presented a tsunami inundation forecasting approach based on a CNN trained with a large quantity of synthetic tsunami scenarios and verified the approach against both synthetic tsunamis and the actual tsunami observations from the 2011 Tohoku earthquake. The CNN tsunami forecast in this study is expected to overcome the bottlenecks of previous simulation-based real-time tsunami forecasting approaches for real applications, such as the difficulty of rapid tsunami source estimations immediately after the earthquake and high computational costs for simulating nonlinear tsunami propagations.

Large earthquakes generating tsunamis are infrequent, resulting in a lack of sufficient data to train the CNN; however, the test using actual observations during the 2011 tsunami event demonstrated that the CNN trained on only synthetic tsunamis can provide an accurate tsunami inundation forecast even for a real tsunami if the target tsunami scenario exists within the distribution of the training data sets. Therefore, it is important to prepare as many tsunami scenarios as possible for training the CNN to address the various mechanisms generating tsunamis.

Large tsunamis can be caused by large slip at plate boundaries, which is a common tsunami generation mechanism, as well as different types of mechanisms, such as a steep angle slip of splay faults 39 , outer-rise normal faults 40 and slow slip at a shallow plate boundary (tsunami earthquake) 41 . Additionally, non-seismic sources, such as volcanic eruptions 42 and landslides 43 , can also cause large tsunamis. To address these various tsunami sources by the CNN tsunami forecasting, a wide variety of tsunami scenarios should be included in the training set. A promising solution to address the various types of tsunami is to generate synthetic tsunami scenarios directly assuming a sea surface fluctuation (e.g., using Gaussian distributions having a range of several kilometres or tens of kilometres) rather than considering a sea surface deformation based on fault movements. Local tsunamis caused by volcanic eruptions or landslides can be represented with a few sea surface displacement units 44 , 45 , and the initial tsunami profiles generated by large earthquakes can also be represented by a superposition of a series of sea surface displacement units 46 , 47 , 48 . Simulating a wide variety of tsunami scenarios and training CNNs on large data sets are computationally expensive; however, it is feasible given the recent advances in high-performance computing of tsunami simulations and an efficient training approach as demonstrated in this study. CNN tsunami forecasting trained on various sea surface displacements should have the potential to be applied to a wide variety of tsunamis, including non-seismic tsunamis, for which the issuance of early warnings has been difficult by employing conventional earthquake-triggered approaches.

Tsunami simulation

We used the TUNAMI-N2 code 49 , 50 , which is distributed by the Tsunami Inundation Modelling Exchange project of the International Union of Geodesy and Geophysics and Intergovernmental Oceanographic Commission of United Nations Educational, Scientific and Cultural Organisation 50 , to create the tsunami simulation data for training the CNNs. The TUNAMI-N2 solves the following nonlinear shallow water Eqs. ( 1 )–( 3 ) with a staggered-grid finite-difference method:

where η is the tsunami height, D is the water depth, and M and N are the velocity fluxes in the x and y directions, respectively. g is the gravitational acceleration (=9.81 m/s 2 ), and n is Manning’s roughness coefficient, which we set to 0.025 s/m 1/3 in the simulations in this paper. As employed in general tsunami simulations, nested grid configurations were prepared in which the grid size was decreased by a factor of 3 (1215, 405, 135, 45 and 15 m) to reduce the computational cost; accordingly, tsunamis in coastal areas are evaluated at higher resolutions than those in offshore areas. Bathymetry data projected onto the Japan Plane Rectangular CS X were used for the simulation. The entire calculation domain with nested grids is illustrated in Supplementary Fig.  3 . The finest domain (∆ x  = ∆ y  = 15 m) covers the Sendai Plain, which experienced a large inundation extent and devastating damage during the 2011 tsunami. The time step of the simulation was set to ∆ t  = 0.2 s.

Sensitivity analysis

To evaluate the sensitivity at each station, we set the signal from a station to zero and evaluated the change in the MSE over 1000 test scenarios. An occlusion test was conducted for every observation point, and the sensitivity of an observation point was evaluated based on the relative change in the MSE from the baseline MSE (no occlusion). The sensitivity is calculated by Eq. ( 4 ):

where n is the number of test scenarios, \({\mathrm{MSE}}^\prime\) is the MSE with occlusion, and MSE is the MSE without occlusion (baseline).

Data availability

The tsunami observation data used in this study are available from The Nationwide Ocean Wave information network for Ports and HArbourS, NOWPHAS ( https://www.mlit.go.jp/kowan/nowphas/index_eng.html ). The GNSS observation data processed by Shu and Xu are used and available from ( https://doi.pangaea.de/10.1594/PANGAEA.914110 ). Other relevant data in this study are available from the corresponding author upon reasonable request.

Code availability

The code that supports the findings in this study are available from the corresponding author upon reasonable request.

Koshimura, S. & Shuto, N. Response to the 2011 Great East Japan Earthquake and Tsunami disaster. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 373 , 20140373 (2015).

Article   ADS   Google Scholar  

Titov, V., Rabinovich, A. B., Mofjeld, H. O., Thomson, R. E. & González, F. I. The global reach of the 26 December 2004 Sumatra Tsunami. Science 309 , 2045–2048 (2005).

Article   ADS   CAS   Google Scholar  

Satake, K., Okal, E. A. & Borrero, J. C. Tsunami and its hazard in the Indian and Pacific Oceans: Introduction. Pure appl. geophys. 164 , 249–259 (2007).

Bernard, E. & Titov, V. Evolution of tsunami warning systems and products. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 373 , 20140371 (2015).

Kaneda, Y. The advanced ocean floor real time monitoring system for mega thrust earthquakes and tsunamis-application of DONET and DONET2 data to seismological research and disaster mitigation. OCEANS 2010 MTS/IEEE SEATTLE, 2010, 1–6 (2010).

Mochizuki, M. et al. S-Net project: Performance of a large-scale seafloor observation network for preventing and reducing seismic and tsunami disasters. 2018 OCEANS—MTS/IEEE Kobe Techno-Oceans (OTO), 1–4 (2018).

Oishi, Y., Imamura, F. & Sugawara, D. Near-field tsunami inundation forecast using the parallel TUNAMI-N2 model: Application to the 2011 Tohoku-Oki earthquake combined with source inversions. Geophys. Res. Lett. 42 , 1083–1091 (2015).

Tsushima, H., Hino, R., Ohta, Y., Iinuma, T. & Miura, S. tFISH/RAPiD: Rapid improvement of near-field tsunami forecasting based on offshore tsunami data by incorporating onshore GNSS data. Geophys. Res. Lett. 41 , 3390–3397 (2014).

Kawamoto, S. et al. REGARD: A new GNSS-based real-time finite fault modeling system for GEONET. J. Geophys. Res. Solid Earth 122 , 1324–1349 (2017).

Maeda, T., Obara, K., Shinohara, M., Kanazawa, T. & Uehira, K. Successive estimation of a tsunami wavefield without earthquake source data: A data assimilation approach toward real‐time tsunami forecasting. Geophys. Res. Lett. 42 , 7923–7932 (2015).

Wang, Y., Satake, K., Maeda, T. & Gusman, A. R. Data assimilation with dispersive tsunami model: A test for the Nankai Trough. Earth, Planets Space 70 , 131 (2018).

Mai, P. M. et al. The earthquake-source inversion validation (SIV) project. Seismological Res. Lett. 87 , 690–708 (2016).

Article   Google Scholar  

Krizhevsky, A., Sutskever, I. & Hinton, G. E. ImageNet classification with deep convolutional neural networks. in Proc. 25th Int. Conf. Neural Information Processing Systems 1097–1105 (2012).

Liang, L., Liu, M., Martin, C. & Sun, W. A deep learning approach to estimate stress distribution: A fast and accurate surrogate of finite-element analysis. J. R. Soc. Interface 15 , 20170844 (2018).

Kim, B. et al. Deep Fluids: A generative network for parameterized fluid simulations. Computer Graph. Forum 38 , 59–70 (2019).

Kiranyaz, S. et al. 1D convolutional neural networks and applications: A survey. Mech. Syst. Signal Process.   151 , 107398 (2021).

Romano, F. et al. Clues from joint inversion of tsunami and geodetic data of the 2011 Tohoku-oki earthquake. Sci. Rep. 2 , 385 (2012).

Article   CAS   Google Scholar  

Han, Z., Zhao, J., Leung H., Ma, K. F. & Wang, W. A review of deep learning models for time series prediction. IEEE Sens. J.   21 , 7833–7848 (2021).

Fujii, Y., Satake, K., Sakai, S., Shinohara, M. & Kanazawa, T. Tsunami source of the 2011 off the Pacific coast of Tohoku Earthquake. Earth Planets Space 63 , 55 (2011).

The Headquaters for Earthquake Research Promotion. A list of tidal and tsunami observation stations. https://www.jishin.go.jp/database/observation_database/spots/ (2019).

Geospatial Information Authority of Japan. GNSS earth observation network system. https://www.gsi.go.jp/ENGLISH/index.html (2020).

Okada, Y. Internal deformation due to shear and tensile faults in a half-space. Bull. Seismol. Soc. Am. 82 , 1018–1040 (1992).

Maas, A. L., Hannun, A. Y. & Ng, A. Y. Rectifier nonlinearities improve neural network acoustic models. in  Proc. 30th Int. Conf. Machine Learning  (2013).

Srivastava, N., Hinton, G., Krizhevsky, A., Sutskever, I. & Salakhutdinov, R. Dropout: A simple way to prevent neural networks from overfitting. J. Mach. Learn. Res. 15 , 1929–1958 (2014).

MathSciNet   MATH   Google Scholar  

Kingma, D. P. & Ba, J. Adam: A method for stochastic optimization. Preprint at https://arxiv.org/pdf/1412.6980.pdf  (2014).

Paszke, A. et al. Automatic differentiation in PyTorch. in NIPS Autodiff Workshop  (2017).

Sergeev, A. & Balso, M. D. Horovod: Fast and easy distributed deep learning in TensorFlow. Preprint at https://arxiv.org/pdf/1802.05799.pdf (2018).

Zeiler, M. D. & Fergus R. Visualizing and understanding convolutional networks. in European Conference on Computer Vision , 818–833 (2014).

Nielsen, A. A. K. & Voigt, C. A. Deep learning to predict the lab-of-origin of engineered DNA. Nat. Commun. 9 , 3135 (2018).

Ports and Harbours Breau, Ministry of Land, Infrastructure, Transport and Tourism. The Nationwide Ocean Wave information network for Ports and HArbourS, NOWPHAS. https://www.mlit.go.jp/kowan/nowphas/index_eng.html (2020).

Shu Y. & Xu P. 1-Hz PPP displacements at GEONET stations during the 2011 Tohoku Mw9.0 earthquake. PANGEA (2020).

Suppasri, A. et al. Lessons learned from the 2011 Great East Japan Tsunami: Performance of tsunami countermeasures, coastal buildings, and tsunami evacuation in Japan. Pure Appl. Geophysics. 170 , 993–1018 (2013).

Shibayama, T., Matsumaru, R., Takagi, H., Esteban, M. & Mikami, T. Field survey of the 2011 off the Pacific coast of Tohoku Earthquake Tsunami disaster to the south of Miyagi Prefecture. J. Jpn. Soc. Civ. Eng. Ser. B2 (Coast. Eng.) 67 , 1301–1305 (2011).

Google Scholar  

Muhari, A., Imamura, F., Suppasri, A. & Mas, E. Tsunami arrival time characteristics of the 2011 East Japan Tsunami obtained from eyewitness accounts, evidence and numerical simulation. J. Nat. Disaster Sci. 34 , 91–104 (2012).

Satake, K., Fujii, Y., Harada, T. & Namegaya, Y. Time and space distribution of coseismic slip of the 2011 Tohoku Earthquake as inferred from tsunami waveform data. Bull. Seismol. Soc. Am. 103 , 1473–1492 (2013).

Kubota, T., Saito, T., Chikasada, N. Y. & Suzuki, W. Ultrabroadband seismic and tsunami wave observation of high‐sampling ocean‐bottom pressure gauge covering periods from seconds to hours. Earth Space Sci. 7 , e2020EA001197 (2020).

Saito, T. & Tsushima, H. Synthesizing ocean bottom pressure records including seismic wave and tsunami contributions: Toward realistic tests of monitoring systems. J. Geophys. Res. Solid Earth 121 , 8175–8195 (2016).

Yamamoto, N. et al. Multi-index method using offshore ocean-bottom pressure data for real-time tsunami forecast. Earth Planets Space 68 , 128 (2016).

Park, J.-O., Tsuru, S., Kodaira, S., Cummins, P. R. & Kaneda, Y. Splay fault branching along the Nankai subduction zone. Science 297 , 1157–1160 (2002).

Kanamori, H. Seismological evidence for a lithospheric normal faulting—the Sanriku earthquake of 1933. Phys. Earth Planet. Inter. 4 , 289–300 (1971).

Kanamori, H. Mechanism of tsunami earthquakes. Phys. Earth Planet. Inter. 6 , 346–359 (1972).

Paris, R. Source mechanisms of volcanic tsunamis. Philos. Trans. Royal Soc. A Math. Phys. Eng. Sci. 373 , 20140380 (2015).

Tappin, D. R., Watts, P., McMurtry, G. M., Lafoy, Y. & Matsumoto, T. The Sissano, Papua New Guinea tsunami of July 1998—offshore evidence on the source mechanism. Mar. Geol. 175 , 1–23 (2001).

Fukao, Y. et al. Mechanism of the 2015 volcanic tsunami earthquake near Torishima, Japan. Sci. Adv. 4 , eaao0219 (2018).

Watts, P., Grilli, S. T., Tappin, D. R. & Fryer, G. J. Tsunami generation by submarine mass failure. II: Predictive equations and case studies. J. Waterw. Port. Coast. Ocean Eng. 131 , 298–310 (2005).

Saito, T., Ito, Y., Inazu, D. & Hino, R. Tsunami source of the 2011 Tohoku‐Oki earthquake, Japan: Inversion analysis based on dispersive tsunami simulations. Geophys. Res. Lett. 38 , L00G19 (2011).

Tsushima, H. et al. Near-field tsunami forecasting using offshore tsunami data from the 2011 off the Pacific coast of Tohoku Earthquake. Earth Planets Space 63 , 56 (2011).

Hossen, M. J., Cummins, P. R., Dettmer, J. & Baba, T. Tsunami waveform inversion for sea surface displacement following the 2011 Tohoku earthquake: Importance of dispersion and source kinematics. J. Geophys. Res. Solid Earth 120 , 6452–6473 (2015).

Imamura, F. Review of tsunami simulation with a finite difference method. in Long-wave Runup Models (eds Yeh, H., Liu, P. & Synolakis, C.) 25–42 (World Scientific Publishing, 1995).

Goto, C., Ogawa, Y., Shuto, N. & Imamura, F. Numerical method of tsunami simulation with the leap-frog scheme. IUGG/IOC TIME Project Intergovernmental Oceanographic Commission of UNESCO, Manuals and Guides 35 , 1–126 (1997).

Download references

Acknowledgements

The computational resources of the AI Bridging Cloud Infrastructure (ABCI) provided by the National Institute of Advanced Industrial Science and Technology (AIST) were used. The topography and bathymetry data used in the tsunami numerical simulation were obtained by integrating the data from the Central Disaster Prevention Council; Tohoku Regional Development Bureau of Ministry of Land, Infrastructure, Transport and Tourism (MLIT); and Geospatial Information Authority of Japan. We used the observation data of the 2011 Tohoku tsunami obtained from the Nationwide Ocean Wave information network for Ports and HArbourS, NOWPHAS. The NOWPHAS tsunami and tidal observation data are observed by Ports and Harbours Breau, MLIT and processed by Port and Airport Research Institute.

Author information

Authors and affiliations.

Fujitsu Laboratories Ltd., Kawasaki, Japan

Fumiyasu Makinoshima, Yusuke Oishi & Takashi Yamazaki

Earthquake Research Institute, The University of Tokyo, Bunkyo-ku, Tokyo, Japan

Takashi Furumura

International Research Institute of Disaster Science (IRIDeS), Tohoku University, Sendai, Japan

  • Fumihiko Imamura

You can also search for this author in PubMed   Google Scholar

Contributions

F.M. and Y.O. designed the research. F.M. and T.Y. prepared the simulation codes. F.I. prepared the data during the 2011 tsunami event. F.M. and T.F. wrote the manuscript. T.F. and F.I. contributed to data interpretation and provided discussions to improve the quality of the paper.

Corresponding author

Correspondence to Fumiyasu Makinoshima .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Additional information

Peer review information Nature Communications thanks Vasily Titov, Nils Thuerey and the other, anonymous, reviewer for their contribution to the peer review of this work. Peer reviewer reports are available.

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

Supplementary information

Supplementary information, peer review file, rights and permissions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/ .

Reprints and permissions

About this article

Cite this article.

Makinoshima, F., Oishi, Y., Yamazaki, T. et al. Early forecasting of tsunami inundation from tsunami and geodetic observation data with convolutional neural networks. Nat Commun 12 , 2253 (2021). https://doi.org/10.1038/s41467-021-22348-0

Download citation

Received : 01 August 2020

Accepted : 03 March 2021

Published : 15 April 2021

DOI : https://doi.org/10.1038/s41467-021-22348-0

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

This article is cited by

Japanese literature organization and spatiotemporal database system creation for natural disaster analysis.

Heritage Science (2024)

Coastal tsunami prediction in Tohoku region, Japan, based on S-net observations using artificial neural network

  • Yuchen Wang
  • Kentaro Imai
  • Kenji Satake

Earth, Planets and Space (2023)

Urban structure reinforces attitudes towards tsunami evacuation

  • Fumiyasu Makinoshima
  • Yusuke Oishi

Scientific Reports (2023)

Automatic post-tsunami loss modeling using deep learning CNN case study: Miyagi and Fukushima Japan tsunami

  • Shaheen Mohammed Saleh Ahmed
  • Hakan Güneyli

Natural Hazards (2023)

Towards the automated evaluation of product packaging in the Food&Beverage sector through data science/machine learning methods

  • Marika Parcesepe
  • Francesca Forgione
  • Salvatore Rampone

Quality & Quantity (2023)

By submitting a comment you agree to abide by our Terms and Community Guidelines . If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

a case study on tsunami

ScienceDaily

The Science of tsunamis

The word "tsunami" brings immediately to mind the havoc that can be wrought by these uniquely powerful waves. The tsunamis we hear about most often are caused by undersea earthquakes, and the waves they generate can travel at speeds of up to 250 miles per hour and reach tens of meters high when they make landfall and break. They can cause massive flooding and rapid widespread devastation in coastal areas, as happened in Southeast Asia in 2004 and in Japan in 2011.

But significant tsunamis can be caused by other events as well. The partial collapse of the volcano Anak Krakatau in Indonesia in 2018 caused a tsunami that killed more than 400 people. Large landslides, which send immense amounts of debris into the sea, also can cause tsunamis. Scientists naturally would like to know how and to what extent they might be able to predict the features of tsunamis under various circumstances.

Most models of tsunamis generated by landslides are based on the idea that the size and power of a tsunami is determined by the thickness, or depth, of the landslide and the speed of the "front" as it meets the water. In a paper titled "Nonlinear regimes of tsunami waves generated by a granular collapse," published online in the Journal of Fluid Mechanics , UC Santa Barbara mechanical engineer Alban Sauret and his colleagues, Wladimir Sarlin, Cyprien Morize and Philippe Gondret at the Fluids, Automation and Thermal Systems (FAST) Laboratory at the University of Paris-Saclay and the French National Centre for Scientific Research (CNRS), shed more light on the subject. (The article also will appear in the journal's July 25 print edition.)

This is the latest in a series of papers the team has published on environmental flows, and on tsunami waves generated by landslides in particular. Earlier this year, they showed that the velocity of a collapse -- i.e., the rate at which the landslide is traveling when it enters the water -- controls the amplitude, or vertical size, of the wave.

In their most recent experiments, the researchers carefully measured the volume of the granular material, which they then released, causing it to collapse as a cliff would, into a long, narrow channel filled with water. They found that while the density and diameter of the grains within a landslide had little effect on the amplitude of the wave, the total volume of the grains and the depth of the liquid played much more crucial roles.

"As the grains enter the water, they act as a piston, the horizontal force of which governs the formation of the wave, including its amplitude relative to the depth of the water," said Sauret. (A remaining challenge is to understand what governs the speed of the piston.) "The experiments also showed that if we know the geometry of the initial column [the material that flows into the water] before it collapses and the depth of the water where it lands, we can predict the amplitude of the wave."

The team can now add this element to the evolving model they have developed to couple the dynamics of the landslide and the generation of the tsunami. A particular challenge is to describe the transition from an initial dry landslide, when the particles are separated by air, to an underwater granular flow, when the water has an important impact on particle motion. As that occurs, the forces acting on the grains change drastically, affecting the velocity at which the front of grains that make up the landslide enters the water.

Currently, there is a large gap in the predictions of tsunamis based on simplified models that consider the field complexity (i.e., the geophysics) but do not capture the physics of the landslide as it enters the water. The researchers are now comparing the data from their model with data collected from real-life case studies to see if they correlate well and if any field elements might influence the results.

  • Natural Disasters
  • Drought Research
  • Environmental Issues
  • Underwater explosion
  • Ocean surface wave
  • Hurricane Katrina
  • Seismic wave
  • Megatsunami
  • 2004 Indian Ocean earthquake

Story Source:

Materials provided by University of California - Santa Barbara . Original written by James Badham. Note: Content may be edited for style and length.

Journal Reference :

  • Wladimir Sarlin, Cyprien Morize, Alban Sauret, Philippe Gondret. Nonlinear regimes of tsunami waves generated by a granular collapse . Journal of Fluid Mechanics , 2021; 919 DOI: 10.1017/jfm.2021.400

Cite This Page :

  • Salt Substitutes for Blood Pressure Control
  • Icy Dwarf Planets With Geothermal Activity
  • Reforestation Threatens Vast Tropical Grasslands
  • New Direction for Sustainable Batteries
  • A 'Quantum Leap' at Room Temperature
  • Gravastar Inside Another Gravastar?
  • Root Microbes: Secret to a Tastier Cup of Tea
  • Ancient Retroviruses and Our Big Brains
  • Diverse Ancient Volcanoes On Mars
  • Brain 'Programmed' to Learn from Those We Like
  • Syllabus (Spring & Summer)
  • Syllabus (Fall)
  • Orientation
  • Meet the Instructors
  • Google Earth Download and Tutorials
  • Capstone Project: Directions and Requirements
  • Library Resources
  • Getting Help

Coastal Processes, Hazards, and Society

Case Study: Sumatra and Thailand and the 2004 Tsunami

Print

The Importance of Tsunami Warning Systems and the challenges of warning communication.

Think back to the video you watched in Module 7 – which included scenes of the 2004 tsunami event in Indonesia. The beginning of the video focused on the Banda Aceh area of Sumatra, where fishing communities and small coastal cities were completely destroyed, and the end of the video featured the Phuket area, where more tourist beaches were affected.

Through your reading and watching the videos, you hopefully gained an idea of what it is like to be caught in a tsunami with no advanced warning, and how frantic the attempts to get out of the way must be. Imagine what it would be like to try to move small children, sick or elderly people out of the way of a tsunami with before the wave strikes and with no time to spare!

In Module 7, the events in Phuket, Thailand, are described, with tourists enjoying their vacation on the beach at Christmas 2004. Many are oblivious to the dangers of the approaching tsunami. What could have been done differently? If this were to happen again, would these communities be better informed and prepared?

In Module 7 we also mentioned that early warning systems are very tricky because of the challenges of getting the message out soon enough after the earthquake and before the tsunami waves arrive at a particular shoreline. For example, the towns on the west coast of Sumatra are so close to the Andaman fault that they had almost no time to react, so a warning may not have worked, regardless of how well it was transmitted. Banda Aceh, on the northern tip of Sumatra, was devastated in 2004 because people did not have time to react, while there is evidence that some small nearby island communities fared better where traditional knowledge of the natural warning signs such as the sudden receding of the tidal waters was employed, and residents were able to flee to higher ground. Meanwhile, the tourist destinations of Phuket and Phi Phi, and nearby locations in Thailand had 2 hours, but the warnings were lacking. Visitors lacked necessary knowledge of nature’s warning signs and how to react, and may not have felt the earthquake, so many lives were lost.

In response to the enormous loss of life in the 2004 Indian Ocean tsunami, the Global Tsunami Warning and Mitigation System was put in place. The Indian Ocean tsunami warning system now integrates the signals from seismographs and DART Buoys and transmits data to 26 national centers. Warnings at the local level are generated in the form of SMS messages, mosque loudspeakers, sirens, and other methods to warn citizens. How well the warnings translate into lives saved due to rapid response and appropriate behaviors by the citizens depends on each step working properly. The failure of one of the steps can lead to disaster. If the citizens do not have the knowledge needed to take effective action, then the process will not work, and lives will be lost.

In 2012 another earthquake occurred near Banda Aceh in the Indian Ocean, so the newly implemented warning systems were put to the test. In this case, no tsunami was generated by the earthquake, but unfortunately, the weaknesses in the system were revealed. Despite the efforts expended to increase levels of tsunami preparedness since 2004, including new tsunami evacuation shelters and education programs, chaos ensued. Hearing the tsunami warning, people panicked and tried to flee by car, resulting in gridlock on the roads. It was clear that better guidance from the local government was needed, including clear evacuation route signage and regular drills. For more detail on this topic, read the National Geographic article Will Indonesia Be Ready for the Next Tsunami? Clearly, more work is still needed and ongoing to address these weaknesses.

Rubble and debris amidst sand, mud, and standing water.

Learning Check Point

We will spend a few minutes also revisiting the accounts of historic tsunami events – in particular, the 1960 event and its effects in Chile and Hilo, Hawaii, and the important messages about how to survive a tsunami. Please re-read some of the accounts of survival during tsunami events in Heed Natural Warnings .

EET Logo

Case Study: Tsunami in Seaside, Oregon

Oregon coast's vulnerability.

tsunami map

The northwestern coast of Oregon is susceptible to both local and far-field tsunamis. The Cascadia Subduction Zone, where the eastward moving Juan de Fuca plate meets the westward moving North American Plate, is just off the Pacific Northwest coast of the United States and Canada. It is a 750-kilometer long fault zone. This area is very active tectonically, and therefore has the potential to produce large earthquakes and possibly, subsequent tsunamis. This subduction zone is thought to have last ruptured in 1700.

Additionally, far-field earthquakes throughout the Pacific are also capable of spawning tsunamis that could eventually reach the Oregon coast. Historical records show that since 1812, about 28 tsunamis with wave heights greater than one meter have reached the U.S. west coast. The March 1964 "Good Friday" earthquake created the most devastating of these tsunamis. The epicenter of this earthquake was near Anchorage, Alaska. The tsunami that followed this earthquake reached coasts all along the western U.S. within six hours. Cannon Beach, a small coastal community in northwestern Oregon was inundated during this event. Just north of Cannon Beach lay the communities of Seaside and Astoria; these communities were also impacted by the 1964 tsunami wave run-up.

The tsunami run-up is defined as the height of the water onshore observed above a referenced sea level. It is this run-up that has the potential to create massive destruction upon it reaching the shore. Seaside, Oregon is particularly vulnerable to tsunami run-up damage for two reasons: first, it was built at a level very close to sea level and second, the proximity of Tillamook Head (a large rock outcrop) has the potential to concentrate the energy of the waves. Additionally, in the case of a locally-generated tsunami, residents of Seaside would only have minutes to evacuate!

Tsunami Warnings and Evacuation Plans

View this seven-minute online video, Tsunami , produced by NOAA for Science on a Sphere, to learn more about the importance of preparing local communities for a possible tsunami evacuation. The video begins with a description of the devastating 2004 tsunami in the Indian Ocean and concludes with a hypothetical tsunami scenario along the U.S. Pacific Northwest coast.

Tsunami warnings can be divided into two groups: natural or "sensed" warnings and those disseminated via an official agency such as NOAA, the National Weather Service, or your local law or emergency management officials.

Local tsunamis usually leave no time for official warnings. Residents in a tsunami evacuation zone must rely on their senses — if they feel or hear an earthquake, or see or hear unusual wave activity, they should seek higher ground on foot immediately, without waiting for official warning.

Tele-tsunamis are often detected with enough time to warn residents of the potential of an approaching tsunami. Tsunami warning centers rely on an array of sensors including: earthquake information, tide gauges, and tsunami detection buoys. If scientists compile data sufficient to indicate a tsunami threat, official agencies will issue a tsunami warning.

In this activity, students will use a data extraction tool, GeoBrain DEM Explorer, to extract desired Shuttle Radar Topography Mission (SRTM) data for a narrow coastal region near Cannon Beach, Seaside, and Astoria, Oregon. Next, they will then import their newly-defined layers into My World GIS and analyze the data to select contours showing possible tsunami run-up heights of 5, 15 and 20 meters to determine which areas of the coastal region are at risk for each of the three run-up heights. In addition, students will determine if existing transportation infrastructure (by highway and/or foot) is adequate for escape routes from the Seaside middle and high schools.

« Previous Page       Next Page »

  • Tsunami Run-up Prediction for Seaside, Oregon with My World GIS
  • Teaching Notes
  • Step-by-Step Instructions
  • Tools and Data
  • Going Further

SERC

  • About this Site
  • Accessibility

Citing and Terms of Use

Material on this page is offered under a Creative Commons license unless otherwise noted below.

Show terms of use for text on this page »

Show terms of use for media on this page »

tsunami map

  • None found in this page
  • Last Modified: October 24, 2023
  • Short URL: https://serc.carleton.edu/111358 What's This?
  • Privacy Policy & Terms of Use
  • Terms & Conditions

Safey Emergency System

Case Study – Tsunami in Japan

a case study on tsunami

On Friday afternoon the 11th of March 2011, a devastating earthquake occurred outside the eastern coast of Japan.

It was the most powerful earthquake ever recorded to have hit Japan , and the fourth most powerful earthquake in the world since modern record-keeping began in 1900. The 9.0 magnitude earthquake triggered powerful tsunami waves that reached heights of up to 40.5 meters in Miyako in Tohoku’s Iwate prefecture, and in the Sendai area the wave traveled up to 10 km inland.

The earthquake moved Honshu (the main island of Japan) 2.4 m east, and generated sound waves detected by the low-orbiting GOCE satellite. SAFEY Emergency System, at this time had a beta version of the SAFEY system and provided alerts to three travelers in Tokyo metropolitan area. They received the red alert of the tsunami warning at 14:49 JST, 47 minutes before the wave hit hit the shoreline. 

Recent Posts

  • Coronavirus Tracker Launched Today
  • Coronavirus
  • October 2019 Preview
  • August 2019 in Review
  • Modi’s Kashmir gambit stokes fears of more violence

IMAGES

  1. (PDF) Case Study on Japan Earthquake and Tsunami

    a case study on tsunami

  2. (PDF) Tsunami Case Studies

    a case study on tsunami

  3. a case study on tsunami

    a case study on tsunami

  4. a case study on tsunami

    a case study on tsunami

  5. 2004 tsunami case study

    a case study on tsunami

  6. Case Study of Indian Ocean Tsunami

    a case study on tsunami

VIDEO

  1. #4 Tsunami Survival Strategy

COMMENTS

  1. 2018 Sulawesi, Indonesia Earthquake and Tsunami Case Study

    2018 Sulawesi, Indonesia Earthquake and Tsunami Case Study Overview On Friday 28th September 2018 a magnitude 7.5 earthquake struck Palu, on the Indonesian island of Sulawesi, just before dusk wreaking havoc and destruction across the city and triggering a deadly tsunami on its coast.

  2. Earthquakes and tsunami

    GCSE WJEC Earthquakes and tsunami - WJEC Case study: Japan tsunami 2011 (HIC) Earthquakes are caused by the release of built-up pressure at plate boundaries. They can destroy buildings and...

  3. The near-field tsunami generated by the 15 January 2022 ...

    Article Open access Published: 07 September 2022 The near-field tsunami generated by the 15 January 2022 eruption of the Hunga Tonga-Hunga Ha'apai volcano and its impact on Tongatapu, Tonga...

  4. Tsunami Disasters: Case Studies and Reports

    This Collection is part of the 'Tsunami Disaster Channel' containing a number of case studies and reports relevant to tsunami disasters, where we try to find out what we have learnt from the past and how we can best reduce risk in future natural disasters.

  5. Indian Ocean tsunami of 2004

    Key People: Michelle Howard See all related content → Top Questions Indian Ocean tsunami of 2004, tsunami that hit the coasts of several countries of South and Southeast Asia in December 2004. The tsunami and its aftermath were responsible for immense destruction and loss on the rim of the Indian Ocean.

  6. Tsunami and Earthquake Research

    Field study of the effects of the December 2004 and March 2005 earthquakes and tsunamis - April 2005. December 26, 2004, Sumatra-Andaman Islands. Tsunami generation from the 2004 M=9.1 Sumatra-Andaman earthquake. Initial findings on tsunami sand deposits, damage, and inundation in Sumatra - January 2005.

  7. Tsunami Case Studies

    Tsunamis are caused by geological processes, such as earthquakes, landslides, or volcanic eruptions, that displace large volumes of ocean water. Large-magnitude, subduction zone earthquakes, where two plates in the ocean push into each other, are the most common source of the recent large tsunamis.

  8. General Review of the Worldwide Tsunami Research

    With the advancement of the global economy, the coastal region has become heavily developed and densely populated and suffers significant damage potential considering various natural disasters, including tsunamis, as indicated by several catastrophic tsunami disasters in the 21st century. This study reviews the up-to-date tsunami research from two different viewpoints: tsunamis caused by ...

  9. Giant tsunami monitoring, early warning and hazard assessment

    Tsunami evacuation models have been used for case studies worldwide and can be used for evidence-driven resource allocation 154, to understand the impact of earthquake-induced debris on evacuation ...

  10. Early forecasting of tsunami inundation from tsunami and ...

    In this study, we presented a tsunami inundation forecasting approach based on a CNN trained with a large quantity of synthetic tsunami scenarios and verified the approach against both synthetic ...

  11. Response to the 2011 Great East Japan Earthquake and Tsunami disaster

    1. Introduction On 11 March, 2011 a devastating tsunami triggered by a Mw 9.0 earthquake struck the northern Pacific coast of Japan, and completely destroyed many coastal communities, particularly in Iwate, Miyagi and Fukushima prefectures.

  12. PDF TSUNAMI: Japan Tsunami of 2011

    2 Preparedness Amassadors ase Studies Most people in the towns along the coast of Japan felt safe from a tsunami.After a tsunami wave hit Japan's coast in 1960 from an earthquake that happened in Chile, the town officials created levees.A levee is a high sea wall built to hold back large sea waves. The levees that were built were 10-16 feet tall, which would have been tall enough to ...

  13. Tsunami Case Studies

    The case studies are grouped according to the tsunamigenic source: earthquake (2004 Indian (Sumatra-Andaman) earthquake, 2011 Tōhoku earthquake, 1964 Alaska earthquake, 1960 Valdivia earthquake, 1946 Aleutian Island earthquake, 1908 Messina-Reggio earthquake, 1755 Great Lisbon earthquake), landslide (Storegga Slides 30,000 and 7,200-7,000 YBP, P...

  14. (PDF) Tsunami Case Studies

    The case studies are grouped according to the tsunamigenic source: A view from high ground of one of the first large (>5 m) waves during the 1960 Chilean tsunami at Onagawa, Japan. After...

  15. The Science of tsunamis

    The word 'tsunami' brings immediately to mind the havoc that can be wrought by these uniquely powerful waves. The tsunamis we hear about most often are caused by undersea earthquakes, and the ...

  16. Environmental hazards Case study: Indian Ocean Tsunami 2004

    National 5 Environmental hazards Case study: Indian Ocean Tsunami 2004 Understanding why natural hazards occur can help countries to manage or prevent their consequences. Case studies...

  17. Case Study: Sumatra and Thailand and the 2004 Tsunami

    Case Study: Sumatra and Thailand and the 2004 Tsunami The Importance of Tsunami Warning Systems and the challenges of warning communication. Think back to the video you watched in Module 7 - which included scenes of the 2004 tsunami event in Indonesia.

  18. A case study of tsunami detection system and ocean wave imaging

    Tsunami detection and investigation of its early warning is the very important issue nowadays, which supports our existing system more precise. • This paper proposes a case study of the mathematical models of the ocean wave imaging schemes and the Tsunami detection system model for the Japan's region where Tsunamis hits on March 11, 2011. •

  19. Case Study: Indian Ocean Tsunami 2004

    Case Study: Indian Ocean Tsunami 2004 Just under a decade ago one of the largest earthquakes ever recorded struck off the coast of Indonesia, triggering a deadly tsunami.

  20. Case Study

    Historical records show that since 1812, about 28 tsunamis with wave heights greater than one meter have reached the U.S. west coast. The March 1964 "Good Friday" earthquake created the most devastating of these tsunamis. The epicenter of this earthquake was near Anchorage, Alaska. The tsunami that followed this earthquake reached coasts all ...

  21. Case Study

    Case Study - Tsunami in Japan On Friday afternoon the 11th of March 2011, a devastating earthquake occurred outside the eastern coast of Japan. It was the most powerful earthquake ever recorded to have hit Japan, and the fourth most powerful earthquake in the world since modern record-keeping began in 1900.

  22. Tohoku Earthquake and Tsunami Japan 2011

    Case study examining the causes of the Tohoku 2011 Earthquake that hit Japan in 2011 and measured 8.9 on the Richter scale. It was the largest earthquake to ...