Two-dimensional simulations of the tsunami dynamo effect using the finite element method

Similar documents
Three Dimensional Simulations of Tsunami Generation and Propagation

Effect of the Emperor seamounts on trans-oceanic propagation of the 2006 Kuril Island earthquake tsunami

Scaling relations of seismic moment, rupture area, average slip, and asperity size for M~9 subduction-zone earthquakes

Tsunami signals from the 2006 and 2007 Kuril earthquakes detected at a seafloor geomagnetic observatory

NUMERICAL SIMULATIONS FOR TSUNAMI FORECASTING AT PADANG CITY USING OFFSHORE TSUNAMI SENSORS

Slip distributions of the 1944 Tonankai and 1946 Nankai earthquakes including the horizontal movement effect on tsunami generation

LETTER Earth Planets Space, 56, , 2004

Source of the July 2006 West Java tsunami estimated from tide gauge records

Estimation of S-wave scattering coefficient in the mantle from envelope characteristics before and after the ScS arrival

Coulomb stress change for the normal-fault aftershocks triggered near the Japan Trench by the 2011 M w 9.0 Tohoku-Oki earthquake

Japan Agency for Marine-Earth Science and Technology (JAMSTEC), 2-15 Natsushima-cho, Yokosuka, Kanagawa, , JAPAN

Tsunami modeling from the seismic CMT solution considering the dispersive effect: a case of the 2013 Santa Cruz Islands tsunami

Seismic Activity and Crustal Deformation after the 2011 Off the Pacific Coast of Tohoku Earthquake

Tsunami waveform analyses of the 2006 underthrust and 2007 outer-rise Kurile earthquakes

Numerical simulation of seismic cycles at a subduction zone with a laboratory-derived friction law

Did a submarine landslide contribute to the 2011 Tohoku tsunami?

Coseismic slip distribution of the 1946 Nankai earthquake and aseismic slips caused by the earthquake

A PROTOTYPE OF WEB-APPLICATION FOR TSUNAMI DATABASE ALONG SOUTHERN JAVA ISLAND COASTLINE

Tsunami source area of the 2011 off the Pacific coast of Tohoku Earthquake determined from tsunami arrival times at offshore observation stations

Earthquakes and Tsunamis

JCR (2 ), JGR- (1 ) (4 ) 11, EPSL GRL BSSA

Fault Length and Direction of Rupture Propagation for the 1993 Kushiro-Oki Earthquake as Derived from Strong Motion Duration

Geophysical Journal International

REAL-TIME TSUNAMI INUNDATION FORECAST STUDY IN CHIMBOTE CITY, PERU

Seismic signals from tsunamis in the Pacific Ocean

TSUNAMI CHARACTERISTICS OF OUTER-RISE EARTHQUAKES ALONG THE PACIFIC COAST OF NICARAGUA - A CASE STUDY FOR THE 2016 NICARAGUA EVENT-

3D modeling of the cycle of a great Tohoku oki earthquake, considering frictional behavior at low to high slip velocities

RELATION BETWEEN RAYLEIGH WAVES AND UPLIFT OF THE SEABED DUE TO SEISMIC FAULTING

Tsunami Simulation of 2009 Dusky Sound Earthquake in New Zealand

Effects of Surface Geology on Seismic Motion

Groundwater changes related to the 2011 Off the Pacific Coast of Tohoku Earthquake (M9.0)

TSUNAMI HAZARD ASSESSMENT FOR THE CENTRAL COAST OF PERU USING NUMERICAL SIMULATIONS FOR THE 1974, 1966 AND 1746 EARTHQUAKES

High initial time sensitivity of medium range forecasting observed for a stratospheric sudden warming

3-D modelling the electric field due to ocean tidal flow and comparison with observations

SOURCE INVERSION AND INUNDATION MODELING TECHNOLOGIES FOR TSUNAMI HAZARD ASSESSMENT, CASE STUDY: 2001 PERU TSUNAMI

Lessons from the 2004 Sumatra earthquake and the Asian tsunami

revised October 30, 2001 Carlos Mendoza

Atmospheric boundary waves excited by the tsunami generation related to the 2011 great Tohoku Oki earthquake

Contents of this file

Hitoshi Hirose (1), and Kazuro Hirahara (2) Abstract. Introduction

Seismogeodesy for rapid earthquake and tsunami characterization

VALIDATION OF TSUNAMI INUNDATION MODELING FOR THE 2004 SUMATRA-ANDAMAN EARTHQUAKE FOR MAKING HAZARD MAPS IN PENANG AND LANGKAWI, MALAYSIA

Joint inversion of strong motion, teleseismic, geodetic, and tsunami datasets for the rupture process of the 2011 Tohoku earthquake

Centroid-moment-tensor analysis of the 2011 off the Pacific coast of Tohoku Earthquake and its larger foreshocks and aftershocks

Successive estimation of a tsunami wavefield without earthquake source data: A data assimilation approach toward real-time tsunami forecasting

Tsunami waveform inversion of the 2007 Bengkulu, southern Sumatra, earthquake

A complex rupture image of the 2011 off the Pacific coast of Tohoku Earthquake revealed by the MeSO-net

Differentiating earthquake tsunamis from other sources; how do we tell the difference?

A database of global seismic travel times

TSUNAMI PROPAGATION AND INUNDATION MODELINGS ALONG SOUTH-EAST COAST OF PAPUA NEW GUINEA

Hiroaki; Hino, Ryota; Fujimoto, Hir. Citation GEOPHYSICAL RESEARCH LETTERS (2011) Right American Geophysical

SUPPLEMENTARY INFORMATION

Tsunami waves swept away houses and cars in northern Japan and pushed ships aground.

Numerical Tsunami Propagation of 1703 Kanto Earthquake

Real time Monitoring System for Earthquakes and Tsunamis (DONET)

Magnitude 8.2 NORTHWEST OF IQUIQUE, CHILE

Supplementary Figure 1 Published rupture models of the Tohoku-oki earthquake that included tsunami data as constraints. Each curve is labeled with

Geophysical Research Letters. Supporting Information for

Rapid magnitude determination from peak amplitudes at local stations

DETERMINATION OF SLIP DISTRIBUTION OF THE 28 MARCH 2005 NIAS EARTHQUAKE USING JOINT INVERSION OF TSUNAMI WAVEFORM AND GPS DATA

Tsunami Waveform Inversion based on Oceanographic Radar Data

Earthquake Source. Kazuki Koketsu. Special Session: Great East Japan (Tohoku) Earthquake. Earthquake Research Institute, University of Tokyo

THREE-DIMENSIONAL FINITE DIFFERENCE SIMULATION OF LONG-PERIOD GROUND MOTION IN THE KANTO PLAIN, JAPAN

Trans-Pacific temperature field in the mantle transition region from seismic and electromagnetic tomography

A new technique for the incorporation of seafloor topography in electromagnetic modelling

SCIENCE OF TSUNAMI HAZARDS

Linear potential theory for tsunami generation and propagation

Electromagnetic Induction by Ocean Currents and the Conductivity of the Oceanic Lithosphere

Source rupture process of the 2003 Tokachi-oki earthquake determined by joint inversion of teleseismic body wave and strong ground motion data

Sendai Earthquake NE Japan March 11, Some explanatory slides Bob Stern, Dave Scholl, others updated March

The Japanese University Joint Seismic Observations at the Niigaka-Kobe Tectonic Zone

Coseismic slip model

Source Characteristics of Large Outer Rise Earthquakes in the Pacific Plate

EARTHQUAKE SOURCE PARAMETERS FOR SUBDUCTION ZONE EVENTS CAUSING TSUNAMIS IN AND AROUND THE PHILIPPINES

Supporting Online Material for

Supporting Online Material for

Effects of shallow layer reverberation on measurement of teleseismic P wave travel times for ocean bottom seismograph data

Special Section Marine Control-Source Electromagnetic Methods

Centroid moment-tensor analysis of the 2011 Tohoku earthquake. and its larger foreshocks and aftershocks

University of Bristol - Explore Bristol Research. Publisher's PDF, also known as Version of record

Source Fault Model of the 1771 Yaeyama Tsunami, Southern Ryukyu Islands, Japan, Inferred from Numerical Simulation

GEOPHYSICAL RESEARCH LETTERS, VOL. 31, L19604, doi: /2004gl020366, 2004

2008 Monitoring Research Review: Ground-Based Nuclear Explosion Monitoring Technologies

NUMERICAL SIMULATION AS GUIDANCE IN MAKING TSUNAMI HAZARD MAP FOR LABUAN ISLAND

Earthquakes Chapter 19

Occurrence of quasi-periodic slow-slip off the east coast of the Boso peninsula, Central Japan

Inversion of tsunami data. A. Sladen CNRS, Géoazur 1/35

MODELING OF HIGH-FREQUENCY WAVE RADIATION PROCESS ON THE FAULT PLANE FROM THE ENVELOPE FITTING OF ACCELERATION RECORDS

Near-field tsunami forecasting using offshore tsunami data from the 2011 off the Pacific coast of Tohoku Earthquake

Magnitude 7.1 NEAR THE EAST COAST OF HONSHU, JAPAN

Indian Ocean Tsunami Warning System: Example from the 12 th September 2007 Tsunami

Sumatra earthquake from tsunami tide gauge record inversion

STUDY ON APPROPRIATE MODELING OF TSUNAMIS IN MALAYSIA FOR RISK EVALUATION

Modeling of the 2011 Tohoku-oki. oki Tsunami and it s s impacts to Hawaii

WERA Ocean Radar Capability of Real-Time Tsunami Detection

Masataka; Matsuoka, Toshifumi. Citation EARTH PLANETS AND SPACE (2011), 63(

LETTER Earth Planets Space, 63, , 2011

2, from which K = # " 2 % 4 3 $ ) ( )

JMA Tsunami Warning Services. Takeshi KOIZUMI Senior Coordinator for International Earthquake and Tsunami Information Japan Meteorological Agency

Title. Author(s)Fujii, Yoshiaki; Kodama, Jun-ichi; Fukuda, Daisuke. CitationProceedings ISRM Congress 2015, 2015: 513. Issue Date

Transcription:

GEOPHYSICAL RESEARCH LETTERS, VOL. 40, 4560 4564, doi:10.1002/grl.50823, 2013 Two-dimensional simulations of the tsunami dynamo effect using the finite element method Takuto Minami 1 and Hiroaki Toh 2 Received 29 June 2013; revised 2 August 2013; accepted 5 August 2013; published 4 September 2013. [1] Conductive seawater moving in the geomagnetic main field generates electromotive force in the ocean. This effect is well known as the oceanic dynamo effect. Recently, it has been reported that tsunamis are also associated with the oceanic dynamo effect, and tsunami-induced electromagnetic field variations were actually observed on the seafloor. For instance, our research group succeeded in observing tsunami-induced magnetic variations on the seafloor in the northwest Pacific at the time of the 2011 Tohoku earthquake. In this study, we developed a time domain tsunami dynamo simulation code using the finite element method to explain the tsunami-induced electromagnetic variations observed on the seafloor. Our simulations successfully reproduced the observed seafloor magnetic variations as large as 3 nt. It was also revealed that an initial rise in the horizontal magnetic component prior to the tsunami arrival as large as 1nTwasinducedbythetsunami. Citation: Minami, T., and H. Toh (2013), Two-dimensional simulations of the tsunami dynamo effect using the finite element method, Geophys. Res. Lett., 40, 4560 4564, doi:10.1002/grl.50823. 1. Introduction [2] The oceanic dynamo phenomenon was first investigated mainly to elucidate effects of long-period ocean flows on electromagnetic (EM) field variations. For instance, Sanford [1971] derived useful formulation for the relationship between low-frequency particle motion of seawater and motionally induced EM variations of negligible self-induction. Tyler [2005] then studied oceanic dynamo effects due to tsunamis, hereafter called tsunami dynamo effects, and derived a useful formula for tsunami-induced EM variations including the selfinduction in order to explain the tsunami-induced magnetic signature at low Earth orbit satellite altitudes. Thanks to the advent of seafloor EM observation techniques [e.g., Toh et al., 2006; Kasaya and Goto, 2009], tsunami-induced EM variations started to be observed on the seafloor as well [e.g., Toh et al., 2011; Manoj et al., 2011; Suetsugu et al., 2012]. It is now possible that we further deduce information of tsunami generation as well as propagation from tsunami-induced EM data. However, most of the preceding studies are in frequency Additional supporting information may be found in the online version of this article. 1 Graduate School of Science, Kyoto University, Sakyo, Japan. 2 Data Analysis Center for Geomagnetism and Space Magnetism, Kyoto University, Sakyo, Japan. Corresponding author: T. Minami, Graduate School of Science, Kyoto University, Kitashirakawa oiwake-cho, Sakyo-ku, Kyoto 6068502, Japan. (minami@kugi.kyoto-u.ac.jp) 2013. American Geophysical Union. All Rights Reserved. 0094-8276/13/10.1002/grl.50823 4560 domain and unable to appreciate the effects of actual bathymetry and conductivity structures beneath the seafloor. [3] In this study, we developed a two-dimensional (2-D) time domain tsunami dynamo simulation code using the finite element method (FEM), which can include bathymetry as well as arbitrary conductivity structures beneath the seafloor. In the present paper, the observed data and our simulation method are introduced in detail first. Second, we compare our numerical results with magnetic data observed on the seafloor in the northwest Pacific at the time of the 2011 off-tohoku earthquake tsunami. Finally, the results of the comparison are discussed and summarized in terms of future tsunami early warning. 2. Tsunami Signals Detected in the Northwest Pacific [4] Our SeaFloor ElectroMagnetic Station (SFEMS) installed at the northwest Pacific site (NWP) successfully recorded the tsunami-induced magnetic field at the time of the 2011 earthquake off the Pacific coast of Tohoku. Although the SFEMS has an horizontal electrometer as well as a three-component fluxgate and Overhauser magnetometer [Toh et al., 2006], it could not detect tsunami-induced electric fields due to large electric noises. Figure 1 shows the location of NWP and the epicenter of the 2011 Tohoku earthquake. NWP is about 1500km from the epicenter, and its depth is 5616 m. In Figure 2 (bottom), colored lines with dots show the time series of the vector magnetic components observed at NWP for 3 h after the earthquake. About 100 min after the earthquake, the upward component of the magnetic field, b z, started to vary and has a negative peak as large as approximately 3 nt. On the other hand, the horizontal magnetic component parallel to the tsunami propagation direction, b y has a negative peak as large as approximately 3 nt a few minutes after the negative peak of b z.thesemagnetic field variations can be considered as tsunami signals because (1) variations of horizontal magnetic components appeared only in the predicted tsunami propagation direction, (2) the onset time of the magnetic variations almost coincided with the 3.11 tsunami arrival time predicted by the National Oceanic and Atmospheric Administration (NOAA) (http:// www.ngdc.noaa.gov/hazard/honshu_11mar2011.shtml), and (3) strong magnetic variations due to external sources were not observed at three on-land magnetic observatories operated by Japan Meteorological Agency (JMA) at the estimated time of tsunami arrival (see section 1 of the supporting information.). Toh et al. [2011] pointed out a large sensitivity of the horizontal magnetic field to propagation directions of tsunamis, which enables us to estimate vector tsunami properties by single site observation alone. In the vicinity of NWP, sea level changes were observed by Ocean Bottom Pressure

Figure 1. The epicenter of the 2011 Tohoku earthquake that occurred on 11 March 2011 (open red star), its W-phase MT solution by USGS, and the seafloor EM observatory installed in the northwest Pacific (red rectangle). The epicentral distance to NWP (41.1026 N, 159.9518 E) is approximately 1500 km. Two DART stations operated by NOAA (yellow squares) are also depicted. Sea level changes at the DART stations were fitted by our 2-D hydrodynamic simulations in order to calculate the magnetic signals observed at NWP. gauges (OBPs) operated by NOAA at DART21401 and DART21419, as shown in Figure 1. It was found that profiles of the sea level changes at the two DART stations are very similar to the downward magnetic component observed at NWP. In the present paper, we try to explain both the sea level changes at the DART stations and the magnetic variations at NWP by 2-D tsunami dynamo simulations, assuming that tsunamis can be approximated by plane waves. 3. Simulation Method [5] We developed a 2-D FEM tsunami dynamo simulation code in order to explain the sea level changes at DART21401 and DART21419 and the tsunami-induced magnetic variations at NWP simultaneously. We adopted simulations in the time domain, because transient waves such as solitary waves cost a lot to reproduce in the frequency domain. Furthermore, we adopted FEM because unstructured triangular meshes can express actual bathymetry, which may virtually be impossible using rectangular meshes [Minami and Toh, 2012]. Our simulation method consists of two steps. In the first step, oceanic flows associated with tsunami propagation are calculated. Second, the induction equation in terms of the magnetic field is solved numerically for given oceanic flows. In order to obtain consistent results between the two steps, the same numerical mesh was used for both finite element calculations. [6] In the first step, the Laplace equation was solved in terms of the velocity potential, assuming the oceanic flows are irrotational. In the present paper, we focused mainly on offshore tsunami propagation observable at NWP. We therefore adopted linear boundary conditions on the sea surface. As for the seafloor, we assumed negligible bottom friction. The governing equation to be solved and the related boundary conditions are summarized as follows: ΔΦ ¼ 0 (1) Φ þ gη ¼ 0 at sea surface (2) η Φ ¼ 0 z at sea surface (3) Φ ¼ 0 n at sea bottom (4) [7] Here, Φ, g, and η denote the velocity potential, the gravity acceleration, and the sea surface elevation, respectively. Operators, / t, / z, and / n, denote differentiations with respect to time, the vertical coordinate, and the direction normal to the seafloor, respectively. Note that η and z are upward positive throughout this paper. The leapfrog method was adopted to solve equations (2) and (3), while equation (1) is solved by FEM in order to calculate Φ with the given boundary values. [8] In the second step, we calculated the tsunami-induced EM fields using the tsunami flows of the first step as follows: We began with equation (9) in Tyler [2005], the induction equation in terms of the tsunami-induced vertical magnetic component, b z ¼ H ðf z u H ÞþK 2 b z : (5) [9] K =(μσ) 1 is the magnetic diffusivity, where μ and σ are the magnetic permeability and the conductivity, respectively. b z, u H, and F z denote the tsunami-induced vertical magnetic component, the horizontal oceanic velocity, and the vertical component of the geomagnetic main field, respectively. In 2-D configuration with tsunamis propagating only in the y direction, we can set the x component of H and v H to nil. Under this circumstance, equation (5) is reduced to 2 y 2 þ 2 z 2 μσ b z ¼ μσ y F zu y : (6) [10] Equation (6) is applicable wherever the conductivity is homogeneous, including the regions above and beneath the ocean layer. When equation (6) is used in the air, σ = 0 and u y = 0 are substituted to make equation (6) the Laplace 4561

Figure 2. Comparison of the simulation results with the observed sea level changes and magnetic variations. The top panel shows sea level changes at DART21401, DART21419, and NWP, while the bottom panel shows the vertical and horizontal component of the magnetic field at NWP. In the bottom panel, b y corresponds to the horizontal component parallel to the tsunami propagation direction, while b z is the upward component. In both panels, colored and black lines denote observed and simulated time series, respectively. In the top panel, distances written next to the station names are the epicentral distances of each station. equation for b z. In our Galerkin FEM solver, equation (6) is integrated over each element. As for time evolution, the Crank-Nicolson method is adopted. It follows that equation (6) is discretized to give 1 2 2 y 2 þ 2 z 2 b n nþ1 1 z þ b z μσ b nþ1 n z b z Δt ¼ 1 2 μσ y F z u n nþ1 y þ u y ; (7) where b n z, u n y,andδtdenote b z and u y at the nth time step, and the time interval between two time steps, respectively. All the boundary conditions under which equation (7) is solved are b z = 0 (Dirichlet type) at enough distances from the center grid. The other components of tsunami-induced EM fields, b y and e x, can be calculated using the following relations: b y y ¼ b z (8) z e x y ¼ b z [11] Note here that our simulations corresponded to socalled transverse electric (TE) mode calculations in magneto-telluric (MT) studies. As for transverse magnetic (TM) mode, electromotive forces (emf) seem to be generated mainly by coupling of the horizontal oceanic flow, u y, with the horizontal component of the geomagnetic main field perpendicular to the tsunami propagation direction, F x. However, incompressible oceanic flows are not able to generate motional induction in TM mode [Larsen, 1971]. Emfs due to incompressible flows in TM mode immediately short-circuit, so electrical potential differences cannot exist within the seawater (see section 2 in the supporting information). As a result, in 2-D configuration, tsunami-induced EM (9) variations at NWP can be obtained by calculations in TE mode alone. 4. Comparison of Detected Magnetic Signals With Tsunami Dynamo Simulations [12] Here we reproduce the sea level changes at DART21401 and DART21419 and the vector magnetic signals observed at NWP at the time of the 2011 off-tohoku earthquake tsunami, by using our 2-D FEM simulation code. 4.1. Initial Settings for the 2011 Tohoku Tsunami [13] In order to explain the magnetic signals observed at NWP, we first reproduced the sea level changes at DART21401 and DART21419. We assigned four subfaults with a total length of 180 km and a width of 50 km based on Maeda et al. s [2011] subfault setting. In our simulations, the initial seawater elevation was assumed to be equal to the coseismic vertical displacement of the seafloor which was computed by Okada s [1985] analytical formulas assuming a homogeneous half-space with V p = 5.8 km s -1, V s = 3.2 km s -1, and ρ = 2.6 10-3 kg m -3 [Dziewonski and Anderson, 1981], while the initial velocity field was assumed to be nil everywhere. [14] For the 2-D numerical grid, 200 km wide bathymetry data were averaged and resampled every 5 km along a straight line drawn from the epicenter to NWP. In pursuit of consistent calculation results between sea level changes and tsunamiinduced EM fields, the same numerical mesh was used in both calculations (see section 3 in the supporting information). 4.2. Simulation Results and Discussion [15] Figure 2 shows comparison results between observed and calculated time series of sea level changes and tsunamiinduced magnetic fields. The simulation results were calculated 4562

for a homogeneous seawater conductivity of 4 S/m and a homogeneous resistivity of 100 Ω m beneath the ocean. In both the hydrodynamic and tsunami dynamo EM simulations, the same time step of Δt =5 s was adopted. We adopted an initial sea level distribution calculated for slip lengths of each subfault of 1.1, 0.4, 5.3, and 6.6 m from the coast to trench side. The sea level changes at the two DART sites were well reproduced by our 2-D hydrodynamic simulations, especially for the tsunami arrival times and shapes of the first waves. As for the EM components, the observed magnetic field was also well reproduced. The presence of a negative peak in the vertical magnetic component as large as 3 nt, and an initial rise and the subsequent larger negative peak in the horizontal component as large as 1 nt and 3 nt, respectively, were confirmed. [16] Figure 3 shows a side view of predicted EM field distribution at the time of the 2011 Tohoku tsunami. The figure features a snapshot of an instant at approximately 104 min after the earthquake occurrence. It is evident that an initial rise in b y, which can be observed only on the seafloor, had already passed over NWP prior to the first arrival of the tsunami. Figure 4 illustrates how the initial rise in seafloor b y is generated. The initial rise is caused by the induced electric field, e x, which slightly precedes the inducing v B field and points the direction opposite to v B. Although the induced e x is weaker than the inducing v B, difference in phase between them makes e x significant in front of v B, Figure 4. Generation mechanism of an initial rise in seafloor b y prior to the peak of b z. The emf itself (blue circles with crosses inside) is driven by the coupling of horizontal tsunami flows with the vertical component of the geomagnetic main field. Tsunami-induced b z and e x slightly precede the sea level change in deep oceans due to self-induction effects of the tsunami itself. Because e x is opposite to v B, the slight phase lead of e x to v B causes the initial rise in seafloor b y preceding the peak of the sea level change. Difference in phase between the peak of b z and the secondary peak of b y becomes approximately T/4 due to the current composed of e x and v B. which causes an initial rise in seafloor b y prior to b z variations. It was also found by our additional simulation that the induced EM field variations around the first arrival look very similar to those induced by a solitary wave (see section 4 in the supporting information). It therefore is probable that the initial rise in seafloor b y and the subsequent main peaks in b y and b z result from the mechanism described in the solitary first wave. This small rise in seafloor b y needs further investigation because it may enable us to detect tsunami arrivals by seafloor magnetic observations before actual arrivals of tsunami peaks. [17] Although our simulation results successfully reproduced the observed sea level changes and the tsunami-induced magnetic field simultaneously, it is difficult to constrain the source model of the 2011 Tohoku tsunami. Tsunami phenomena and their source models are always threedimensional (3-D) in practice, because realistic fault geometry, bathymetry, and the electrical conductivity structure in the Earth are always 3-D. It was found difficult in our 2-D simulations to reproduce later tsunami phases in both the hydrodynamic and EM fields, which implies that we should include the contribution of 3-D bathymetry and coastlines. Figure 3. Side view of the calculated tsunami-induced EM fields in TE mode when the 3.11 tsunami passed over the NWP site. The red and blue colors denote positive and negative values in each component, respectively. In the lowest panel, the sum of the inducing v B field and the induced electric field, e x, is shown. The vertical dashed line indicates the location of NWP. 5. Summary [18] At the time of the gigantic tsunamis associated with the 2011 Tohoku earthquake, SFEMS observed vector magnetic signals as large as 3 nt on the northwest Pacific seafloor approximately 1500 km away from the epicenter. The horizontal magnetic variations appeared only in the anticipated tsunami propagation direction, which implies that even the 4563

single site EM observation on the seafloor enables us to extract vector properties of tsunamis. [19] We developed a 2-D FEM time domain tsunami dynamo simulation code in order to reproduce the sea level changes at DART21401 and DART21419 and magnetic tsunami signals at NWP simultaneously. By adoption of FEM, our simulations can include realistic bathymetry and arbitrary conductivity structures beneath the seafloor. [20] Our 2-D tsunami dynamo simulation well reproduced the profiles of the sea level changes observed at the two DART sites and the vector tsunami-induced magnetic field observed at NWP. It was found that the significant peaks both in the vertical and horizontal components as large as 3 nt are very likely to be induced by the 2011 Tohoku tsunami. In addition, our simulation also revealed that an initial rise in the horizontal magnetic field as large as 1 nt was also induced by the 2011 Tohoku tsunami about 5 min prior to their passage. The generation of the initial rise in b y observed at NWP can be understood by considering a simple solitary wave case. It therefore is possible that tsunami arrivals can be detected before its actual arrival, using the initial rise/retreat in the horizontal magnetic component. More detailed analysis of the initial rise may lead to significant improvement of the existing global tsunami early warning systems. [21] Acknowledgments. R/V Kairei and ROV Kaiko7000II of the Japan Agency for Marine-Earth Science and Technology are greatly acknowledged for their skillful help at the time of sea experiments. T.M. expresses his sincere thanks to Y. Hamano, N. Tada, H. Sugioka, and A. Ito for their warm support during the operation of the seafloor geomagnetic observatory. Earthquake Research Institute, University Tokyo, is acknowledged because this work was also supported by its cooperative research program. This work is also supported by Grants in Aid for Scientific Research from MEXT, Japan (21654061, 22340124, 23654158, and 24340108). [22] The Editor thanks Robert Tyler and an anonymous reviewer for their assistance in evaluating this paper. References Dziewonski, A. M., and D. L. Anderson (1981), Preliminary reference Earth model, Phys. Earth Planet. Inter., 25, 297 356, doi:10.1016/0031-9201 (81)90046-7. Kasaya, T., and T. Goto (2009), A small ocean bottom electromagnetometer and ocean bottom electrometer system with an arm-folding mechanism (Technical Report), Explor. Geophys., 40, 41 48, doi:10.1071/eg08118. Larsen, J. C. (1971), The electromagnetic field of long and intermediate water waves, J. Mar. Res., 29, 28 45. Maeda, T., T. Furumura, S. Sakai, and M. Shinohara (2011), Significant tsunami observed at ocean-bottom pressure gauges during the 2011 off the Pacific coast of Tohoku Earthquake, Earth Planets Space, 63, 803 808, doi:10.5047/eps.2011.06.005. Manoj, C., S. Maus, and A. Chulliat (2011), Observation of magnetic fields generated by tsunamis, EOS Trans, AGU, 92(2), 13 14, doi:10.1029/ 2011EO020002. Minami, T., and H. Toh (2012), An improved forward modeling method for two-dimensional electromagnetic induction problems with bathymetry, Earth Planets Space, 64, e9 e12, doi:10.5047/eps.2012.04.012. Okada, Y. (1985), Surface deformation due to shear and tensile faults in a half-space, Bull. Seismol. Soc. Am., 75(4), 1135 1154. Sanford, T. B. (1971), Motionally induced electric and magnetic fields in the sea, J. Geophys. Res., 76(15), 3476 3492. Suetsugu, D., et al. (2012), TIARES Project Tomographic investigation by seafloor array experiment for the Society hotspot, Earth Planets Space, 64, i iv, doi:10.5047/eps.2011.11.002. Toh, H., Y. Hamano, and M. Ichiki (2006), Long-term seafloor geomagnetic station in the northwest Pacific: A possible candidate for a seafloor geomagnetic observatory, Earth Planets Space, 58, 697 705. Toh, H., K. Satake, Y. Hamano, Y. Fujii, and T. Goto (2011), Tsunami signals from the 2006 and 2007 Kuril earthquakes detected at a seafloor geomagnetic observatory, J. Geophys. Res., 116, B02104, doi:10.1029/ 2010JB007873. Tyler, R. H. (2005), A simple formula for estimating the magnetic fields generated by tsunami flow, Geophys. Res. Lett., 32, L09608, doi:10.1029/ 2005GL022429. 4564