Volume 40 Issue 11
Nov.  2021
Turn off MathJax
Article Contents
Peitao Wang, Zhiyuan Ren, Lining Sun, Jingming Hou, Zongchen Wang, Ye Yuan, Fujiang Yu. Observations and modelling of the travel time delay and leading negative phase of the 16 September 2015 Illapel, Chile tsunami[J]. Acta Oceanologica Sinica, 2021, 40(11): 11-30. doi: 10.1007/s13131-021-1830-2
Citation: Peitao Wang, Zhiyuan Ren, Lining Sun, Jingming Hou, Zongchen Wang, Ye Yuan, Fujiang Yu. Observations and modelling of the travel time delay and leading negative phase of the 16 September 2015 Illapel, Chile tsunami[J]. Acta Oceanologica Sinica, 2021, 40(11): 11-30. doi: 10.1007/s13131-021-1830-2

Observations and modelling of the travel time delay and leading negative phase of the 16 September 2015 Illapel, Chile tsunami

doi: 10.1007/s13131-021-1830-2
Funds:  The National Key Research and Development Program of China under contract Nos 2018YFC1407000 and 2016YFC1401500; the National Natural Science Foundation of China under contract Nos 41806045 and 51579090.
More Information
  • Corresponding author: E-mail: wpt@nmefc.cn
  • Received Date: 2020-11-21
  • Accepted Date: 2021-02-23
  • Available Online: 2021-08-27
  • Publish Date: 2021-11-30
  • The systematic discrepancies in both tsunami arrival time and leading negative phase (LNP) were identified for the recent transoceanic tsunami on 16 September 2015 in Illapel, Chile by examining the wave characteristics from the tsunami records at 21 Deep-ocean Assessment and Reporting of Tsunami (DART) sites and 29 coastal tide gauge stations. The results revealed systematic travel time delay of as much as 22 min (approximately 1.7% of the total travel time) relative to the simulated long waves from the 2015 Chilean tsunami. The delay discrepancy was found to increase with travel time. It was difficult to identify the LNP from the near-shore observation system due to the strong background noise, but the initial negative phase feature became more obvious as the tsunami propagated away from the source area in the deep ocean. We determined that the LNP for the Chilean tsunami had an average duration of 33 min, which was close to the dominant period of the tsunami source. Most of the amplitude ratios to the first elevation phase were approximately 40%, with the largest equivalent to the first positive phase amplitude. We performed numerical analyses by applying the corrected long wave model, which accounted for the effects of seawater density stratification due to compressibility, self-attraction and loading (SAL) of the earth, and wave dispersion compared with observed tsunami waveforms. We attempted to accurately calculate the arrival time and LNP, and to understand how much of a role the physical mechanism played in the discrepancies for the moderate transoceanic tsunami event. The mainly focus of the study is to quantitatively evaluate the contribution of each secondary physical effect to the systematic discrepancies using the corrected shallow water model. Taking all of these effects into consideration, our results demonstrated good agreement between the observed and simulated waveforms. We can conclude that the corrected shallow water model can reduce the tsunami propagation speed and reproduce the LNP, which is observed for tsunamis that have propagated over long distances frequently. The travel time delay between the observed and corrected simulated waveforms is reduced to <8 min and the amplitude discrepancy between them was also markedly diminished. The incorporated effects amounted to approximately 78% of the travel time delay correction, with seawater density stratification, SAL, and Boussinesq dispersion contributing approximately 39%, 21%, and 18%, respectively. The simulated results showed that the elastic loading and Boussinesq dispersion not only affected travel time but also changed the simulated waveforms for this event. In contrast, the seawater stratification only reduced the tsunami speed, whereas the earth’s elasticity loading was responsible for LNP due to the depression of the seafloor surrounding additional tsunami loading at far-field stations. This study revealed that the traditional shallow water model has inherent defects in estimating tsunami arrival, and the leading negative phase of a tsunami is a typical recognizable feature of a moderately strong transoceanic tsunami. These results also support previous theory and can help to explain the observed discrepancies.
  • Generally, tsunami travel time is defined as the time required for the first tsunami wave to propagate from its source to a given point.
  • loading
  • [1]
    Abdolali A, Kadri U, Kirby J T. 2019. Effect of water compressibility, sea-floor elasticity, and field gravitational potential on tsunami phase speed. Scientific Reports, 9(1): 16874. doi: 10.1038/s41598-019-52475-0
    [2]
    Abdolali A, Kirby J T. 2017. Role of compressibility on tsunami propagation. Journal of Geophysical Research: Oceans, 122(12): 9780–9794. doi: 10.1002/2017JC013054
    [3]
    Allgeyer S, Cummins P. 2014. Numerical tsunami simulation including elastic loading and seawater density stratification. Geophysical Research Letters, 41(7): 2368–2375. doi: 10.1002/2014GL059348
    [4]
    An Chao, Liu P L F. 2016. Analytical solutions for estimating tsunami propagation speeds. Coastal Engineering, 117: 44–56. doi: 10.1016/j.coastaleng.2016.07.006
    [5]
    Aránguiz R, González G, González J, et al. 2016. The 16 September 2015 Chile tsunami from the post-tsunami survey and numerical modeling perspectives. Pure and Applied Geophysics, 173(2): 333–348. doi: 10.1007/s00024-015-1225-4
    [6]
    Baba T, Allgeyer S, Hossen J, et al. 2017. Accurate numerical simulation of the far-field tsunami caused by the 2011 Tohoku earthquake, including the effects of Boussinesq dispersion, seawater density stratification, elastic loading, and gravitational potential change. Ocean Modelling, 111: 46–54. doi: 10.1016/j.ocemod.2017.01.002
    [7]
    Baba T, Ando K, Matsuoka D, et al. 2016. Large-scale, high-speed tsunami prediction for the great Nankai Trough earthquake on the K computer. The International Journal of High Performance Computing Applications, 30(1): 71–84. doi: 10.1177/1094342015584090
    [8]
    Baba T, Cummins P R, Thio H K, et al. 2009. Validation and joint inversion of teleseismic waveforms for earthquake source models using deep ocean bottom pressure records: A case study of the 2006 Kuril megathrust earthquake. Pure and Applied Geophysics, 166(1-2): 55–76. doi: 10.1007/s00024-008-0438-1
    [9]
    Baba T, Takahashi N, Kaneda Y, et al. 2015. Parallel implementation of dispersive tsunami wave modeling with a nesting algorithm for the 2011 Tohoku tsunami. Pure and Applied Geophysics, 172(12): 3455–3472. doi: 10.1007/s00024-015-1049-2
    [10]
    Barazangi M, Isacks B L. 1976. Spatial distribution of earthquakes and subduction of the Nazca plate beneath South America. Geology, 4(11): 686–692. doi: 10.1130/0091-7613(1976)4<686:SDOEAS>2.0.CO;2
    [11]
    Berger M J, George D L, Leveque R J, et al. 2011. The GeoClaw software for depth-averaged flows with adaptive refinement. Advances in Water Resources, 34(9): 1195–1206. doi: 10.1016/j.advwatres.2011.02.016
    [12]
    Contreras-López M, Winckler, P, Sepúlveda I, et al. 2016. Field survey of the 2015 Chile tsunami with emphasis on coastal wetland and conservation areas. Pure and Applied Geophysics, 173(2): 349–367. doi: 10.1007/s00024-015-1235-2
    [13]
    DeMets C, Gordon R G, Argus D F, et al. 1994. Effect of recent revisions to the geomagnetic reversal time scale on estimates of current plate motions. Geophysical Research Letters, 21(20): 2191–2194. doi: 10.1029/94GL02118
    [14]
    Eblé M C, Mungov G T, Rabinovich A B. 2015. On the leading negative phase of major 2010-2014 tsunamis. Pure and Applied Geophysics, 172(12): 3493–3508. doi: 10.1007/s00024-015-1127-5
    [15]
    Heidarzadeh M, Murotani S, Satake K, et al. 2016. Source model of the 16 September 2015 Illapel, Chile, Mw8.4 earthquake based on teleseismic and tsunami data. Geophysical Research Letters, 43(2): 643–650. doi: 10.1002/2015GL067297
    [16]
    Heidarzadeh M, Satake K. 2013. Waveform and spectral analyses of the 2011 Japan tsunami records on tide gauge and DART stations across the Pacific Ocean. Pure and Applied Geophysics, 170(6−8): 1275–1293. doi: 10.1007/s00024-012-0558-5
    [17]
    Heidarzadeh M, Satake K. 2014. The El Salvador and Philippines tsunamis of August 2012: insights from sea level data analysis and numerical modeling. Pure and Applied Geophysics, 171(12): 3437–3455. doi: 10.1007/s00024-014-0790-2
    [18]
    Heidarzadeh M, Satake K, Murotani S, et al. 2015. Deep-water characteristics of the trans-Pacific tsunami from the 1 April 2014 Mw8.2 Iquique, Chile earthquake. Pure and Applied Geophysics, 172(3−4): 719–730. doi: 10.1007/s00024-014-0983-8
    [19]
    Heidarzadeh M, Satake K, Takagawa T, et al. 2018. A comparative study of far-field tsunami amplitudes and ocean-wide propagation properties: insight from major trans-Pacific tsunamis of 2010−2015. Geophysical Journal International, 215(1): 22–36. doi: 10.1093/gji/ggy265
    [20]
    Ho T C, Satake K, Watada S. 2017. Improved phase corrections for transoceanic tsunami data in spatial and temporal source estimation: application to the 2011 Tohoku earthquake. Journal of Geophysical Research: Solid Earth, 122(12): 10155–10175. doi: 10.1002/2017JB015070
    [21]
    Inazu D, Saito T. 2013. Simulation of distant tsunami propagation with a radial loading deformation effect. Earth Planets Space, 65(8): 835–842. doi: 10.5047/eps.2013.03.010
    [22]
    Jakeman J D, Nielsen O M, Putten K V, et al. 2010. Towards spatially distributed quantitative assessment of tsunami inundation models. Ocean Dynamics, 60(5): 1115–1138. doi: 10.1007/s10236-010-0312-4
    [23]
    Ji Chen, Wald D J, Helmberger D V. 2002. Source description of the 1999 hector mine, California, earthquake, part I: wavelet domain inversion theory and resolution analysis. Bulletin of the Seismological Society of America, 92(4): 1192–1207. doi: 10.1785/0120000916
    [24]
    Kajiura K. 1963. The leading wave of a tsunami. Bulletin of the Earthquake Research Institute, University of Tokyo, 41(3): 535–571
    [25]
    Kato T, Terada Y, Nishimura H, et al. 2011. Tsunami records due to the 2010 Chile Earthquake observed by GPS buoys established along the Pacific coast of Japan. Earth, Planets and Space, 63(6): e5–e8. doi: 10.5047/eps.2011.05.001
    [26]
    Kirby J T, Shi Fengyan, Tehranirad B, et al. 2013. Dispersive tsunami waves in the ocean: model equations and sensitivity to dispersion and Coriolis effects. Ocean Modelling, 62: 39–55. doi: 10.1016/j.ocemod.2012.11.009
    [27]
    Li Bo, Ghosh A. 2016. Imaging rupture process of the 2015 Mw 8.3 Illapel earthquake using the US seismic array. Pure and Applied Geophysics, 173(7): 2245–2255. doi: 10.1007/s00024-016-1323-y
    [28]
    Lu W F, Jiang Y W, Lin J. 2013. Modeling propagation of 2011 Honshu tsunami. Engineering Applications of Computational Fluid Mechanics, 7(4): 507–518. doi: 10.1080/19942060.2013.11015489
    [29]
    Okada Y. 1985. Surface deformation due to shear and tensile faults in a half-space. Bulletin of the Seismological Society of America, 75(4): 1135–1154. doi: 10.1785/BSSA0750041135
    [30]
    Poupardin P, Heinrich P, Hébert H, et al. 2018. Traveltime delay relative to the maximum energy of the wave train for dispersive tsunamis propagating across the Pacific Ocean: the case of 2010 and 2015 Chilean Tsunamis. Geophysical Journal International, 214(3): 1538–1555. doi: 10.1093/gji/ggy200
    [31]
    Prastowo T, Cholifah L, Madlazim. 2018. Analysis of travel time delay for large tsunamis across the pacific and Indian Oceans. Science of Tsunami Hazards, 37(4): 195–212
    [32]
    Rabinovich A B, Candella R N, Thomson R E. 2013a. The open ocean energy decay of three recent trans-Pacific tsunamis. Geophysical Research Letters, 40(12): 3157–3162. doi: 10.1002/grl.50625
    [33]
    Rabinovich A B, Thomson R E. 2007. The 26 December 2004 Sumatra tsunami: analysis of tide gauge data from the World Ocean Part 1. Indian Ocean and South Africa. Pure and Applied Geophysics, 164(2): 261–308. doi: 10.1007/s00024-006-0164-5
    [34]
    Rabinovich A B, Thomson R E, Fine I V. 2013b. The 2010 Chilean tsunami off the west coast of Canada and the northwest coast of the United States. Pure and Applied Geophysics, 170(9-10): 1529–1565. doi: 10.1007/s00024-012-0541-1
    [35]
    Rabinovich A B, Titov V V, Moore C W, et al. 2017. The 2004 Sumatra tsunami in the southeastern Pacific Ocean: New global insight from observations and modeling. Journal of Geophysical Research: Oceans, 122(10): 7992–8019. doi: 10.1002/2017JC013078
    [36]
    Rabinovich A B, Woodworth P L, Titov V V. 2011. Deep-sea observations and modeling of the 2004 Sumatra tsunami in Drake Passage. Geophysical Research Letters, 38(16): L16604. doi: 10.1029/2011GL048305
    [37]
    Roberts S G, Nielsen O M, Gray D, et al. 2015. ANUGA User Manual, Release 2.0. Symonston: Geoscience Australia, https://www.researchgate.net/publication/318511561_ANUGA_User_Manual_Release_20 [2015-05-19/2020-10-01]
    [38]
    Roberts S G, Nielsen O M, Jakeman J. 2008. Simulation of tsunami and flash floods. In: Bock H G, Kostina E, Phu H X, et al., eds. Modeling, Simulation and Optimization of Complex Processes. Berlin, Heidelberg: Springer, doi: 10.1007/978-3-540-79409-7_35
    [39]
    Röbke B R, Vött A. 2017. The tsunami phenomenon. Progress in Oceanography, 159: 296–322. doi: 10.1016/j.pocean.2017.09.003
    [40]
    Satake K, Heidarzadeh M. 2017. A review of source models of the 2015 Illapel, Chile earthquake and insights from tsunami data. Pure and Applied Geophysics, 174(1): 1–9. doi: 10.1007/s00024-016-1450-5
    [41]
    Shan Di, Wang Peitao, Ren Zhiyuan, et al. 2017. Application and evaluation of the 16 September 2015 Illapel, Chile Mw8.3 earthquake finite fault rupture model from numerical simulation. Haiyang Xuebao (in Chinese), 39(11): 49–60. doi: 10.3969/j.issn.0253-4193.2017.11.005
    [42]
    Tang Liujuan, Titov V V, Moore C, et al. 2016. Real-time assessment of the 16 September 2015 Chile tsunami and implications for near-Field forecast. Pure and Applied Geophysics, 173(2): 369–387. doi: 10.1007/s00024-015-1226-3
    [43]
    Titov V V, Synolakis C E. 1998. Numerical modeling of tidal wave runup. Journal of Waterway, Port, Coastal, and Ocean Engineering, 124(4): 157–171. doi: 10.1061/(ASCE)0733-950X(1998)124:4(157)
    [44]
    Tsai V C, Ampuero J P, Kanamori H, et al. 2013. Estimating the effect of earth elasticity and variable water density on tsunami speeds. Geophysical Research Letters, 40(3): 492–496. doi: 10.1002/grl.50147
    [45]
    Vigny C, Rudloff A, Ruegg J C, et al. 2009. Upper plate deformation measured by GPS in the Coquimbo Gap, Chile. Physics of the Earth and Planetary Interiors, 175(1−2): 86–95. doi: 10.1016/j.pepi.2008.02.013
    [46]
    Wang Dailin. 2015. An ocean depth-correction method for reducing model errors in tsunami travel time: application to the 2010 Chile and 2011 Tohoku tsunamis. Science of Tsunami Hazards, 34(1): 1–22
    [47]
    Wang Dailin, Becker N C, Walsh D, et al. 2012. Real-time forecasting of the April 11, 2012 Sumatra tsunami. Geophysical Research Letters, 39(19): L19601. doi: 10.1029/2012GL053081
    [48]
    Wang Xiaoming, Liu P L F. 2011. An explicit finite difference model for simulating weakly nonlinear and weakly dispersive waves over slowly varying water depth. Coastal Engineering, 58(2): 173–183. doi: 10.1016/j.coastaleng.2010.09.008
    [49]
    Wang Peitao, Yu Fujiang, Yuan Ye, et al. 2016. Effects of finite fault rupture models of submarine earthquakes on numerical forecasting of near-field tsunami. Chinese Journal Of Geophysics (in Chinese), 59(3): 1030–1045. doi: 10.6038/cjg20160324
    [50]
    Watada S. 2013. Tsunami speed variations in density-stratified compressible global oceans. Geophysical Research Letters, 40(15): 4001–4006. doi: 10.1002/grl.50785
    [51]
    Watada S, Kusumoto S, Satake K. 2014. Traveltime delay and initial phase reversal of distant tsunamis coupled with the self-gravitating elastic earth. Journal of Geophysical Research: Solid Earth, 119(5): 4287–4310. doi: 10.1002/2013JB010841
    [52]
    Wei Yong, Bernard E N, Tang Liujuan, et al. 2008. Real-time experimental forecast of the Peruvian tsunami of August 2007 for U.S. Coastlines. Geophysical Research Letters, 35(4): L04609. doi: 10.1029/2007GL032250
    [53]
    Wessel P, Smith W H F. 1998. New, improved version of generic mapping tools released. Eos, Transactions American Geophysical Union, 79(47): 579. doi: 10.1029/98EO00426
    [54]
    Yamazaki Y, Cheung K F, Kowalik Z. 2011. Depth-integrated, non-hydrostatic model with grid-nesting for tsunami generation, propagation, and run-up. International Journal for Numerical Methods in Fluids, 67(12): 2081–2107. doi: 10.1002/fld.2485
    [55]
    Yamazaki Y, Cheung K F, Lay T. 2013. Modeling of the 2011 Tohoku near-field tsunami from finite-fault inversion of seismic waves. Bulletin of the Seismological Society of America, 103(2B): 1444–1455. doi: 10.1785/0120120103
    [56]
    Ye Lingling, Lay T, Kanamori H, et al. 2016. Rapidly estimated seismic source parameters for the 16 September 2015 Illapel, Chile Mw8.3 earthquake. Pure and Applied Geophysics, 173(2): 321–332. doi: 10.1007/s00024-015-1202-y
    [57]
    Zaytsev O, Rabinovich A B, Thomson R E. 2016. A comparative analysis of coastal and open-ocean records of the great Chilean tsunamis of 2010, 2014 and 2015 off the coast of Mexico. Pure and Applied Geophysics, 173(12): 4139–4178. doi: 10.1007/s00024-016-1407-8
    [58]
    Zaytsev O, Rabinovich A B, Thomson R E. 2017. The 2011 Tohoku tsunami on the coast of Mexico: A case study. Pure and Applied Geophysics, 174(8): 2961–2986. doi: 10.1007/s00024-017-1593-z
  • 加载中

Catalog

    通讯作者: 陈斌, bchen63@163.com
    • 1. 

      沈阳化工大学材料科学与工程学院 沈阳 110142

    1. 本站搜索
    2. 百度学术搜索
    3. 万方数据库搜索
    4. CNKI搜索

    Figures(14)  / Tables(3)

    Article Metrics

    Article views (929) PDF downloads(30) Cited by()
    Proportional views
    Related

    /

    DownLoad:  Full-Size Img  PowerPoint
    Return
    Return