simulation of tsunami propagation.pdf

Upload: idlir

Post on 22-Feb-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    1/30

    Coastal Engineering Journal, Vol. 57, No. 4 (2015) 1550016 (30 pages)

    c World Scientific Publishing Company and Japan Society of Civil EngineersDOI:10.1142/S0578563415500163

    Simulation of Tsunami Propagation

    Using Adaptive Cartesian Grids

    Qiuhua Liang,,, Jingming Hou, and Reza Amouzgar,

    State Key Laboratory of Hydrology-Water Resources

    and Hydraulic Engineering, Hohai University,

    Nanjing 210098, P. R. ChinaSchool of Civil Engineering and Geosciences, Newcastle University,

    Newcastle Upon Tyne, NE1 7RU, [email protected]@ncl.ac.uk

    [email protected]

    Received 24 July 2014Accepted 28 July 2015Published 16 September 2015

    This paper presents a 2D model for predicting tsunami propagation on dynamically adap-tive grids. In this model, a finite volume Godunov-type scheme is implemented to solvethe 2D nonlinear shallow water equations on adaptive grids. The simplified adaptive gridachieves automatic adaptation through increasing or reducing the subdivision level of abackground cell according to certain criteria defined by tsunami wave features. The gridsystem is straightforward to implement and no data structure is needed to store grid infor-mation. The present model is validated by applying it to simulate three laboratory-scaletest cases of tsunami propagation over uneven beds and finally used to reproduce the 2011

    Tohoku tsunami in Japan. The model results confirm the models capability in predictingtsunami wave propagation in a reliable and efficient way.

    Keywords: Tsunami Modeling; adaptive grid; shallow water equations; finite volumemethod; Godunov-type scheme.

    Corresponding author.

    1550016-1

    http://dx.doi.org/10.1142/S0578563415500163http://dx.doi.org/10.1142/S0578563415500163
  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    2/30

    Q. Liang, J. Hou & R. Amouzgar

    1. Introduction

    In recent years, tremendous research efforts have been made to develop accurate

    and robust numerical models for tsunamis, for example, the COMCOT (Cornell

    Multi-grid Coupled Tsunami Model), MOST (Method of Splitting Tsunami) model,TUNAMI (Tohoku Universitys Numerical Analysis Model for Investigation), Geo-

    Claw, TSUNAMI3D, THETIS, among others. COMCOT, TUNAMI and GeoClaw

    solve the 2D nonlinear shallow water equations (NSWEs) to represent the tsunami

    propagation and runup, while the TSUNAMI3D and THETIS models solve 3D

    NavierStokes (NS) equations for incompressible fluid flows with free surface and

    interfacial boundaries described based on the concept of the fractional volume of

    fluid (VOF) method. They have been widely used to model tsunami events [Wang

    and Liu,2006;Wei et al.,2008;Oishi et al.,2015;Tang et al.,2012;Imamura,1996;

    George and LeVeque,2008;Arcos and LeVeque,2014;Horrillo et al.,2013;Abadie

    et al.,2012].The wave fronts of a tsunami usually present as bores/surges rapidly moving

    inland, which demands special treatment for accurate simulations. A general way is

    to use fine computational meshes with resolution ranging from several meters to tens

    of meters. But tsunami propagation can take place over an entire ocean and mesh

    refinement over the entire domain may lead to unaffordable computational cost. For

    instance, if 10m 10m cells are adopted for a domain of 100km 100km, 100million computational cells will be produced, which are computationally inhibited

    for most of the existing tsunami models. So it is desirable to intelligently refine

    the grid locally near to the wave fronts instead of the whole domain. This can be

    achieved using adaptive mesh refinement (AMR), which has been an active researchtopic in CFD for over 30 years [Berger and Oliger,1984;Yiu et al.,1996;Lee et al.,

    2011].

    Over recent decades, two grid adaption approaches have been widely used to

    carry out the AMR on Cartesian grids. They are respectively known as block adap-

    tion and hierarchical grid adaption [Popinet,2011;Liang,2012]. The former method

    uses a group of coexisting grids of different resolutions to achieve grid adaption

    [Berger and Oliger, 1984; Berger and LeVeque, 1998]. It has been implemented in

    CeoClaw and applied byGeorge and LeVeque[2008], Watanabe et al. [2012],Arcos

    and LeVeque [2014] to model tsunami propagation. A hierarchical grid generally

    employs a quadtree data structure to store grid information such as neighbors and

    subdivision levels. The information is updated by searching the data tree during

    grid adaptation. Hierarchical grids perform AMR on an individual grid cell rather

    than a block of grid cells as used in the block method. Therefore, it gives more

    flexibility to track the flow features and is favored by numerous researchers, e.g. Yiu

    et al. [1996], Rogers et al. [2001], Popinet[2003], Liang and Borthwick[2009], Lee

    et al.[2011].Popinet[2011,2012] used adaptive quadtree grids in his tsunami model.

    Despite its greater flexibility, quadtree grids suffer from large storage requirement

    1550016-2

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    3/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    and additional computational overhead for searching the data tree [Liang, 2012].

    To improve the efficiency of adaptive quadtree grids, Ji et al. [2010] developed a

    new cell-based structured AMR data structure, which requires less storage than the

    conventional quadtree. Liang[2012] devised a simplified block-based adaptive grid

    system which just imposes subdivision levels on coarse background cells. There aretwo main advantages for the adaptive mesh method proposed in Liang[2012] com-

    paring with the aforementioned conventional approaches. Firstly, no complicated

    procedure is required for generating an initial grid apart from simple allocation of

    specific subdivision level to each of the coarse cells on the background grid. Secondly,

    the neighbors of an arbitrary cell are fully determined by simple algebraic relation-

    ships and thus no data structure is demanded. Grid adaption is straightforward to

    achieve by altering the subdivision level of a cell to follow certain flow conditions.

    Such an efficient grid system may be well-suited for large-scale tsunami simulations

    that involve multi-level bathymetric datasets.

    This paper aims to introduce an efficient tsunami model based on the afore-

    mentioned simplified adaptive grid system. The remainder of the paper is organized

    as follows: the governing equations that mathematically describe the propagation

    of the tsunami waves and the numerical scheme to solve them are briefly reviewed

    in Sec. 2; the adaptive Cartesian grid system is introduced in Sec. 3; the model is

    verified against three laboratory tests and applied to a field-scale tsunami event in

    Sec. 4; brief discussions and conclusions are drawn up in Sec. 5.

    2. Governing Equations and Numerical Scheme

    As inJi et al.[2010] andPopinet[2011] the NSWEs derived from the conservation of

    mass and momentum are chosen to simulate tsunamis. In a vector form, the NSWEs

    can be written as[Liang and Borthwick,2009]

    q

    t +

    f

    x+

    g

    y =S, (1)

    q=

    qx

    qy

    , f=

    uh

    u2h+g(2 2zb)/2uvh

    , g=

    vh

    vuh

    v2h+g(2 2zb)/2

    , (2)

    S= Sb+ Sf=

    0

    gzb/xgzb/y

    +

    0

    Cfu

    u2 +v2

    Cfv

    u2 +v2

    , (3)

    where t is the time; x and y represent the Cartesian coordinates; q denotes the

    vector of conserved flow variables consisting of ,qx=uh and qy =vhy, i.e. the free

    surface water level, unit-width discharges in the x- and y-directions, respectively;

    h, u and v are water depth, depth-averaged velocities in the x- and y-directions,

    respectively; f and g are the flux vectors in the two Cartesian directions; Sdenotes

    1550016-3

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    4/30

    Q. Liang, J. Hou & R. Amouzgar

    the source vector including slope Sb and friction source terms Sf; zb represents the

    bed elevation; Cf is the bed roughness coefficient that is generally calculated by

    gn2/h1/3 with n being the Manning coefficient.

    On the adaptive grid system, the SWEs are solved numerically using a two-step

    MUSCL-Hancock scheme originally developed by van Leer [1984], which consistsof a predictor step and a corrector step and has been widely applied by numerous

    researchers (e.g. Zhou et al. [2002]) for shallow water flow modeling. It has been

    incorporated in a cell-centered finite volume Godunov-type scheme for nonuniform

    grid as presented by Liang [2011, 2012]. In the predictor step, intermediate flow

    variables are predicted over half of a time interval. These intermediate variables

    are then utilized in the corrector step to update the results to a new time level.

    The friction source terms Sfare not evaluated within the MUSCL-Hancock scheme

    but independently calculated using a splitting point-implicit method proposed by

    Bussing and Murmant[1998]. In order to prevent the reversing flow direction caused

    by unrealistic friction force, a limiting step introduced by Liang and Marche[2009]

    is taken into account. Since the numerical scheme is overall explicit, the solution

    stability is essentially controlled by the CourantFriedrichsLewy (CFL) condition

    [Courant et al., 1928], which is used to estimate the time step for next iteration

    during a simulation. Open and closed boundaries are implemented, as by Liang and

    Borthwick[2009].

    3. Dynamically Adaptive Cartesian Grid

    The dynamically adaptive Cartesian grid system developed by Liang [2012] isadopted in the current tsunami model to allow effective grid adaption. After the

    initial grid is generated, dynamical grid adaption can be easily carried out by spec-

    ifying the subdivision level on a coarse background cell according to certain criteria

    related to flow conditions. This section reviews the procedure of implementing the

    dynamically adaptive Cartesian grid system.

    3.1. Grid generation and cell identification

    The initial grid is generated by simply following three steps:

    A rectangular computational domain is selected and discretized using a coarseuniform Cartesian grid which is called background grid. For example, the domain

    in Fig.1(a) is divided into a 2 3 background grid; Each cell on the background grid is checked and refined by giving a specific subdi-

    vision level according to certain criterion. For instance, the subdivision levels on

    background cells (i, j) and (i, j +1) shown in Fig.1(b) are respectively subdivided

    by respectively specifying as level one and two;

    To ensure grid quality, the final mesh should satisfy a 2:1 rule, i.e. no cell is allowedto have a neighbor that is more than two times bigger or smaller. As shown in

    1550016-4

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    5/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    (a) (b) (c)

    (d)

    Fig. 1. (Color online) Structured but nonuniform grid generation: (a) background grid, (b) irregulargrid, (c) regularized grid, (d) cell indexes of regularized grid. Red-bold numbers in this figurerepresent the number of subdivision level.

    Fig. 1(b), cell (i, j+ 1) does not meet the 2:1 rule since its western neighbor is

    four times bigger. So cell (i, j+ 1) should be regularized and subdivided into levelone [Fig.1(c)].

    As illustrated in Fig. 1(d), each cell on such a nonuniform grid can be identified by

    four indexes (i, j, is, js), where is and js are numbered form 1 to Ms. Ms denotes

    the number of sub-cells in the x- or y-direction at background cell (i, j);Ms = 2lev

    with lev representing the subdivision level of (i, j). The size of a background cell

    is x y, and subsequently the dimensions of the sub-cells can be computed byxs= x/2

    lev and ys= y/2lev.

    On such a nonuniform grid, it is not necessary to define a data structure to store

    neighbor information as the neighbors of an arbitrary cell can be entirely determined

    by simple algebraic relationships. For those cells with the same subdivision levels,

    the neighbors can be easily obtained as on uniform grids. For cells with neighbors of

    different subdivision levels, simple algebraic relationships can be derived to specify

    their neighbors. For example, considering the sub-cells of (i, j) in Fig. 1(d), the

    western neighbor of (i, j, 1, 1) and (i, j, 1, 2) has a lower subdivision level and is

    denoted by (i 1, j, isN, jsN) where the subscript N represents neighbor. Theneighbor sub-cell indices isN and jsNcan be obtained by

    isN = 2levN , jsN= Ceiling (js/2), (4)

    1550016-5

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    6/30

    Q. Liang, J. Hou & R. Amouzgar

    where function Ceiling means the smallest integer not less thanjs/2. If the neighbors

    are in higher subdivision levels, e.g. the two eastern neighbors of (i, j, 2, 1) can be

    denoted by (i+ 1, j, isN, jsN), where

    isN= 1, jsN=js 2 1 and js 2. (5)The northern and southern neighbors can be identified in a similar way. Comparing

    to other existing adaptive grid systems that need special data structure to store

    the neighbor information, the storage requirement for the current adaptive grids is

    minimized and only two matrices are required to store the subdivision level of every

    background cell and the cell IDs of all sub-cells for flow calculation. Therefore,

    cell manipulation during grid adaptation becomes very flexible and efficient [Liang,

    2012].

    3.2. Adaption indicator

    The first step to achieve dynamic grid adaptation is to define an adaptation criterion.

    In this work, the adaptation indicator is defined according to water level gradient,

    which is a key flow feature indicating the complexity of tsunami hydrodynamics.

    At cell ic on the aforementioned nonuniform grid, the adaptation indicator can be

    expressed by a dimensionless factor

    ic= GicPq(G)

    , (6)

    where Gic is the averaged gradient of the water level at cell ic and

    Gic =

    x

    2ic

    +

    y

    2ic

    , (7)

    G is a vector containing the water surface gradients for all of the sub-cells inside

    the computational domain, Pq(G) returns the qth percentile of G and q = 1sawhere sa is a user specified coefficient indicating the sensitivity of the adaptation

    procedure. For instance, sa = 20% means that 20% of the flow cells are subject to

    grid refinement. For coarsening, a critical value of coar must also be prescribed.

    Grid coarsening will be carried out if ic is less than coar for all of the sub-cells.

    Moreover, the cells defining the wet-dry fronts (i.e. cells have both wet and dry

    neighbors) are also refined to accurately capture the moving wet-dry fronts.

    3.3. Dynamic grid adaption

    For dynamic grid adaptation, if one or more sub-cells of a background cell (i, j)

    have values of the adaption indicator ic greater than 1 and the subdivision level

    of (i, j) is less than the specified maximum value, (i, j) will be marked for further

    subdivision and its subdivision level will be increased to lev + 1. Additionally, any

    1550016-6

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    7/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    cell defining the wet-dry interfaces will be marked for subdivision to the maximum

    grid refinement level levm to capture the wet-dry interface. If the values of ic of

    all sub-cells in a background cell (i, j) are lower than coarand its subdivision level

    is higher than zero, the cell will be coarsened by changing its subdivision level to

    lev1. The levels of all newly created cells are then checked to guarantee the 2:1 ruleas explained in Sec. 3.1 to ensure the regularity of the grid. The above procedure of

    grid refinement and coarsening are performed at every time step during a simulation.

    3.4. Mapping newly created cells

    After the grid refinement and coarsening, the flow information will be mapped to

    the newly created cells using a simple linear interpolation scheme, as reported by

    Liang [2011]. But the values of bed elevation (bathymetry) and friction parameter

    are interpolated directly from the original dataset or multiple datasets in order to

    represent precisely the domain. In order to reduce the demand of computer memory,the files storing bathymetric data and values of friction parameter are stored offline

    in the local drive and read in during grid adaptation as suggested byPopinet[2011].

    The flow information in a coarsened cell can be calculated by simply averaging

    that of the four child cells. In newly refined cells, the values of bed elevation and

    friction parameter are interpolated linearly from the original datasets. In addition,

    the approach suggested byLiang et al.[2015] is employed to resolve the contradiction

    between the C-property and mass conservation during grid adaption [Liang and

    Borthwick,2009;Popinet,2011]. Once the new grid and the flow information in the

    newly created cells are obtained, the NSWEs are ready to be solved numerically to

    update the flow variables to the next time step by a Godunov-type finite volume

    method as described in Sec. 2.

    4. Model Applications

    The present tsunami model is validated against four test cases in this section, includ-

    ing three laboratory-scale and a real tsunami cases, and the results are compared

    with measured water levels and maximum runup. The Courant number of 0.5 is

    used for all of the simulations.

    4.1. Solitary wave runup on a simple beach

    The laboratory experiments of a solitary wave on a simple beach were carried out by

    Synolakis[1986] and the results can be found in Synolakis et al. [2007]. As sketched

    in Fig.2(a), a solitary wave is generated in a channel at x= X1 to propagate over

    a simple beach with a slope of 1:19.85. The toe of the slope is at X0 = 9.925m

    in this work. X1 = X0 +L/2 and the half wave length L/2 = (1/)arccosh

    20

    with =

    3(H/D)/4 where H and D denote the amplitude of the incipient wave

    and the still water depth, respectively [Synolakis et al., 2007]. The incoming wave

    1550016-7

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    8/30

    Q. Liang, J. Hou & R. Amouzgar

    D

    H

    X

    Y

    x=X0 x=X1

    1:19.85

    zb

    (a)

    (b)

    (c)

    Fig. 2. Solitary wave runup on a simple beach: (a) definition sketch for the case of a solitary waveon a simple beach[Synolakiset al.,2007], (b) adaptive grid for the wave with H= 0.3 m at:t = 15,(c) t = 25.

    is specified at the right-hand side boundary [Nikolos and Delis,2009] as

    (t) =Hsech2

    3H

    4D3C(t T0)

    , (8)

    u(t) = C

    D+, v(t) = 0, (9)

    whereT0= 0.0 is the time when the wave crest reaches the domain;C=

    g(D+H)

    is the wave celerity. Two cases with wave amplitudes ofH= 0.3 m and 0.0185 m are

    considered, which represent breaking and nonbreaking waves, respectively. The stillwater depth is specified to be 1.0 m. A Manning coefficient of 0.01 s/m1/3 is used

    for the channel of stainless steel bed [Synolakis et al., 2007]. The background cell

    resolution is 0.2 m and it will be refined up to 0.1 m according to the grid adaption

    criterion withsa= 0.3. Grid coarsening will be undertaken for icless than 0.7. The

    grid evolution for the case with the higher wave amplitude is sketched in Figs. 2(b)

    and2(c) at t = 15 and 25 where t =t

    g/D [Synolakis,1986].

    The computed and measured water levels are plotted in Figs. 3 and 4, overall

    good agreements with the measured wave profiles are observed for both cases where

    1550016-8

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    9/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    (a) (b)

    (c) (d)

    Fig. 3. Solitary wave runup on a simple beach: comparison between the computed and measuredwater levels for the wave with amplitude of 0.3 m at (a) t = 15, (b) t = 20, (c) t = 25, (d)t = 30.

    H = 0.3 m and 0.0185 m. This demonstrates that the current adaptive grid-based

    model is able to reproduce the wave runup process. However, due to the limita-

    tion of the shallow water equations in representing 3D fluid patterns and breaking

    waves, certain discrepancies between the prediction and measurement are detected,

    especially in the area adjacent to the wave fronts. The simulation results may beimproved by including the Boussinesq approximation in shallow water flow model

    as shown, for example, byKazolea et al.[2012]. As shown in Figs.3 and 4, the com-

    puted water surfaces are almost identical between the simulations on adaptive grid

    and uniform grid of 0.1 m resolution, suggesting that the adaptive grid system can

    capture the key wave features and produce accurate numerical solutions. Tables 1

    and 2 further summarize the performance of the adaptive grid model, in term of

    accuracy and efficiency. The computed root mean square errors (RMSE) for the

    water level along a section in the x-direction are very similar for the simulations

    1550016-9

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    10/30

    Q. Liang, J. Hou & R. Amouzgar

    (a) (b)

    (c) (d)

    Fig. 4. Solitary wave runup on a simple beach: comparison between the computed and measuredwater levels for the wave with amplitude of 0.3 m at (a) t = 40, (b) t = 50, (c) t = 60,(d) t = 70.

    Table 1. Solitary wave runup on a simple beach: RMSE () and relative computational cost forthe wave with H= 0.0185 m.

    t 30 40 50 60 70 Relative CPU time

    RMSE() on adaptive grid [mm] 1.72 2.68 2.40 1.40 0.74 1.00RMSE() on uniform grid [mm] 1.71 2.77 2.42 1.43 0.76 2.12

    Table 2. Solitary wave runup on a simple beach: RMSE () and relative computational cost forthe wave with H= 0.3 m.

    t 15 20 25 30 Relative CPU time

    RMSE() on adaptive grid [mm] 5.30 3.11 4.30 1.66 1.00RMSE() on uniform grid [mm] 5.35 3.19 4.34 1.68 2.27

    1550016-10

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    11/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    on adaptive and uniform grids, revealing that the adaptive grid-based model can

    achieve similar simulation accuracy to the high-resolution uniform grid. The RMSE

    for the adaptive grid is marginally lower than that obtained for the uniform grid.

    The propagation of the solitary wave is successfully reproduced by both models.

    But the adaptive grid model is computationally more efficient and requires less than50% of computing time on the uniform grid.

    4.2. Solitary wave runup over a conical island

    A series of experiments to study the runup of tsunami waves on a conical island were

    conducted in the US Army Engineer Waterways Experiment Station[Briggs et al.,

    1995], and have been widely used to validate SWE models [Titov and Synolakis,

    1995, 1998; Bradford and Sanders, 2002; Lynett et al., 2002; Hubbard and Dodd,

    2002; Choi et al.

    , 2007; Nikolos and Delis, 2009; Zijlema et al.

    , 2011; Hou et al.

    ,2013]. As in [Hou et al.,2013], a 25.92 m27.6 m frictionless computational domainis chosen, in which a conical island with a base diameter of 7.2m, top diameter of

    2.2 m and height of 0.625 m, is located at the domain center (Fig.5). Solitary planar

    waves are generated by a directional wave-maker at the left boundary and the rest

    of the boundaries are closed. The inflow conditions are specified by Eqs. (8) and(9).

    Three different waves are considered with H/D = 0.045, 0.091 and 0.181 which are

    termed as case A, B and C, respectively, and D is specified to be 0.32 m for all cases.

    The simulations are run on a coarse background grid with 130 138 cells with aresolution of 0.2 m, which can be refined up to 0.05 m in certain parts according to

    the wave patterns. The sensitivity parametersa and the grid coarsening parametercoar are set to be 0.15 and 0.8, respectively.

    The 3D view of wave propagation over the island for case A at t = 28.2, 33.2

    and 36.2s is plotted in Fig. 6.The incipient waves propagate along the domain and

    hit the island after about 30.5 s. Part of the wave is refracted around the island as

    it propagates toward the lee side. Two trapped waves are created at each side of

    the island and then collide at the lee side, generating a runup around t = 36.2 s.

    After running down from the island, the waves propagate back toward the island.

    The computed wave patterns for the case B and C are very similar to case A,

    with slight difference in arriving time and magnitude. The corresponding adapted

    grids are plotted in Figs. 6(b), 6(d) and 6(f), confirming that the grid adaptionworks well in tracking wave front by refining and coarsening the relevant cells. To

    further demonstrate the performance of the adaptive grid model, the measured time

    histories of water level at certain wave gauges as sketched in Fig. 5are compared

    with the numerical predictions in Figs. 79 for the case A, B and C, respectively.

    The computed results agree satisfactorily with the measured data for all of the

    three cases. The leading wave-heights and the arrival times are well-predicted at

    most gauges, suggesting the model is capable of reproducing the key features of the

    propagating waves with H/D ranging from 0.045 to 0.181. But it should be noted

    1550016-11

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    12/30

    Q. Liang, J. Hou & R. Amouzgar

    Fig. 5. Solitary wave runup over a conical island: model set up and locations of selected water-levelgauges[Hou et al.,2013].

    that the steepness of the wave fronts and the secondary depression wave following

    the leading wave are not exactly reproduced. Such discrepancies may be due to the

    inherent inability of a 2D SWE model in representing fully 3D flow features [Choi

    et al., 2007]. Similar results are also predicted by other researchers using different

    SWE models, for example [Liu et al.,1995;Lynett et al.,2002;Hubbard and Dodd,

    2002;Nikolos and Delis,2009;Hou et al.,2013].The computed maximum runup on the island is recorded and compared with the

    measurements in Fig. 10. The variable denotes the angle of the section as shown

    in the layout plot (Fig. 5). For all of the three cases, the computed highest and

    the second highest runups are detected at the front and the lee sides, respectively.

    The computed profiles of the maximum runup have similar trends to the measured

    ones, despite certain level of deviation. The computed values are found to be slightly

    higher than the measured data at the front side of the island, perhaps because the

    friction is not taken into account as suggested by other researchers, for example by

    1550016-12

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    13/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    (a) (b)

    (c) (d)

    (e) (f)

    Fig. 6. Solitary wave runup over a conical island: case A: (a) wave propagation at t = 28.2s, (b)adaptive grid at t= 28.2 s, (c) wave propagation at t = 33.2 s, (d) adaptive grid at t = 33.2 s, (e)wave propagation at t = 36.2 s, (f) adaptive grid at t= 36.2 s.

    1550016-13

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    14/30

    Q. Liang, J. Hou & R. Amouzgar

    20 25 30 35 400.04

    0.02

    0

    0.02

    0.04

    t(s)

    (

    m)

    measuredcomputed

    20 25 30 35 400.04

    0.02

    0

    0.02

    0.04

    t(s)

    (

    m)

    measuredcomputed

    (a) (b)

    20 25 30 35 400.04

    0.02

    0

    0.02

    0.04

    t(s)

    (

    m)

    measuredcomputed

    20 25 30 35 400.04

    0.02

    0

    0.02

    0.04

    t(s)

    (

    m)

    measuredcomputed

    (c) (d)

    Fig. 7. Solitary wave runup a conical island: case A: comparison between the computed and mea-sured water levels at (a) WG3, (b) WG9, (c) WG16, (d) WG22.

    Nikolos and Delis [2009]. While the height of the real waves become steeper as a

    result of approaching the shorelines, vertical motions are expected to be more and

    more significant, which are not represented in the 2D depth averaging assumption.

    The computed results by the adaptive grid model are also compared with those

    obtained on the uniform grid with a cell length of 0.1 m. As illustrated in Fig. 8,the time histories of computed water surface on the two different grids agree well

    with each other at all of the gauges, demonstrating that the AMR can produce

    results with equivalent accuracy without compromising computational efficiency.

    The RMSEs computed against measured data for all of the three cases at the five

    gauge points are shown in Tables 35, further confirming the accurate simulation

    results produced by the adaptive grid model. However, the uniform grid model

    requires 1.75, 1.87 and 1.69 times of the runtime, compared with the adaptive grid

    code for case A, B and C, respectively.

    1550016-14

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    15/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    20 25 30 35 400.04

    0.02

    0

    0.02

    0.04

    0.06

    t(s)

    (

    m)

    measured

    computed on adaptive gridcomputed on uniform grid

    20 25 30 35 400.04

    0.02

    0

    0.02

    0.04

    0.06

    t(s)

    (

    m)

    measured

    computed on adaptive gridcomputed on uniform grid

    (a) (b)

    20 25 30 35 40

    0.04

    0.02

    0

    0.02

    0.04

    0.06

    t(s)

    (

    m)

    measured

    computed on adaptive gridcomputed on uniform grid

    20 25 30 35 40

    0.04

    0.02

    0

    0.02

    0.04

    0.06

    t(s)

    (

    m)

    measured

    computed on adaptive gridcomputed on uniform grid

    (c) (d)

    Fig. 8. Solitary wave runup a conical island: case B: comparison between the computed and mea-sured water levels at (a) WG3, (b) WG9, (c) WG16, (d) WG22.

    4.3. Tsunami wave approaching Monai valley

    Monai valley is located in Okushiri island, Japan, and was attacked by a severe

    tsunami happened in the Japan Sea in 1993. This event was reproduced in labora-

    tory using a 1:400 scaled physical model built in the Research Institute for ElectricPower Industry (CRIEPI) in Abiko, Japan [Liu et al., 2008]. The test involves a

    5.488m 3.402 m computational domain over the topography as shown in Fig.11.A 22.5 s incident tsunami wave enters the domain from the left boundary and other

    domain boundaries are set to be closed. During the simulation, the domain is covered

    by a background grid of 98 61 cells with a resolution of 0.056 m. Fine bathymetrydata of 0.014 m resolution is also available for the entire domain, which allows the

    background grid to refine up to 2 levels. The sensitivity parameter sa and the

    grid coarsening parameter coar are specified as 0.24 and 0.8. Uniform Manning

    1550016-15

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    16/30

    Q. Liang, J. Hou & R. Amouzgar

    20 25 30 35 400.06

    0.04

    0.02

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    t(s)

    (

    m)

    measured

    computed

    20 25 30 35 400.06

    0.04

    0.02

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    t(s)

    (

    m)

    measured

    computed

    (a) (b)

    20 25 30 35 400.06

    0.04

    0.02

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    t(s)

    (

    m)

    measuredcomputed

    20 25 30 35 400.06

    0.04

    0.02

    0

    0.02

    0.04

    0.06

    0.08

    0.1

    t(s)

    (

    m)

    measuredcomputed

    (c) (d)

    Fig. 9. Solitary wave runup a conical island: case C: comparison between the computed and mea-sured water levels at (a) WG3, (b) WG9, (c) WG16, (d) WG22.

    coefficient of 0.001m1/3s is imposed on the whole domain as suggested by other

    researchers [Popinet,2011;Funke et al.,2011]. The model is run for 30 s.

    The numerical results at different output times (t = 14, 17 and 20s) are plot-

    ted in Fig. 12 in a 3D view, including the adapted grids. Grid adaptation clearly

    evolves according to the dynamics of the tsunami wave. The adaptive grids effectivelycapture the complex tsunami wave patterns and the moving shoreline by generat-

    ing refined mesh at those regions with steep water surface gradients and near the

    wet-dry front. Figure13 demonstrates the comparison between the measured and

    predicted time histories of water level at three gauges G5, G6 and G7, which are

    located at [4.521m, 1.196 m], [4.521m, 1.696 m] and [4.521m, 2.196 m], respectively.

    In the first 10 s, because of the initial water disturbances in the realistic wave tank,

    the computed water levels do not agree well with the measurements. The results

    are consistent with predictions reported by other researchers [Zhang and Baptista,

    1550016-16

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    17/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    0

    1

    2

    3

    4

    5

    0 60 120 180 240 300 360

    maximumr

    unu

    p(cm)

    ()

    measured computed

    (a)

    0

    3

    6

    9

    12

    0 60 120 180 240 300 360

    maximumr

    unup(cm)

    ()

    measured computed

    (b)

    0

    5

    10

    15

    20

    25

    0 60 120 180 240 300 360

    maximumr

    unup(cm)

    measured

    computed

    ()

    (c)

    Fig. 10. Solitary wave runup a conical island: computed and measured maximum runup for wavecase (a) A, (b) B and (c) C.

    1550016-17

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    18/30

    Q. Liang, J. Hou & R. Amouzgar

    Table 3. Solitary wave runup over a conical island: RMSE () and relative computational cost forcase A.

    Gauges WG3 WG6 WG9 WG16 WG22 Relative CPU time

    RMSE () on adaptive grid [mm] 0.792 1.073 0.9234 1.931 0.097 1.00RMSE () on uniform grid [mm] 0.792 1.074 0.9129 1.923 0.1083 1.75

    Table 4. Solitary wave runup over a conical island: RMSE () and relative computational costfor case B.

    Gauges WG3 WG6 WG9 WG16 WG22 Relative CPU time

    RMSE() on adaptive grid [mm] 1.701 2.313 2.915 0.234 3.309 1.00RMSE () on uniform grid [mm] 1.704 2.313 2.920 0.267 3.325 1.87

    Table 5. Solitary wave runup over a conical island: RMSE () and relative computational costfor case C.

    Gauges WG3 WG6 WG9 WG16 WG22 Relative CPU time

    RMSE() on adaptive grid [mm] 1.540 3.117 4.234 1.961 1.936 1.00RMSE () on uniform grid [mm] 1.532 3.103 4.104 1.994 2.030 1.69

    Fig. 11. Tsunami wave approaching Monai valley: domain and bathymetry for the laboratory exper-iment (unit : m).

    2008; Clain and Clauzon,2010;Nicolsky et al., 2011]. Nevertheless, the amplitude

    and the phase of the first wave are satisfactorily reproduced at the gauges after

    10 s. As shown in Fig.14,the maximum runup height is computed to be 0.072 m in

    the Monai valley, which is equivalent to 28.8 m at the prototype scale and roughly

    1550016-18

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    19/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    (a) (b)

    (c) (d)

    (e) (f)

    Fig. 12. Tsunami wave approaching Monai valley: (a) wave propagation at t = 14 s, (b) adaptivegrid at t = 14 s, (c) wave propagation at t = 17 s, (d) adaptive grid at t = 17 s, (e) wave propagationat t= 20 s, (f) adaptive grid at t= 20s.

    congruous with the field measurement of over 30 m [Matsuyama and Tanaka,2001].

    As compared in Fig. 13 and Table 6, the numerical predictions obtained on the

    adaptive grid are very similar to those predicted on the refined uniform grid at

    a resolution of 0.014 m. Yet, in order to achieve a similar solution accuracy, the

    1550016-19

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    20/30

    Q. Liang, J. Hou & R. Amouzgar

    0 5 10 15 20 25 300.02

    0.01

    0

    0.01

    0.02

    0.03

    0.04

    0.05

    0.06

    t(s)

    (

    m)

    measuredcomputed on adaptive grid

    computed on uniform grid

    0 5 10 15 20 25 300.02

    0.01

    0

    0.01

    0.02

    0.03

    0.04

    0.05

    0.06

    t(s)

    (

    m)

    measuredcomputed on adaptive grid

    computed on uniform grid

    (a) (b)

    0 5 10 15 20 25 300.02

    0.01

    0

    0.01

    0.02

    0.03

    0.04

    0.05

    0.06

    t(s)

    (

    m)

    measuredcomputed on adaptive gridcomputed on uniform grid

    (c)

    Fig. 13. Tsunami wave approaching Monai valley: time histories of computed and measured waterlevel at: (a) G5, (b) G7, (c) G9.

    adaptive grid-based model is found to be about four times more efficient than its

    uniform grid counterpart.

    4.4. 2011 Tohoku tsunami propagation simulationOn March 11, 2011, a large earthquake with Mw = 9.0 off the Pacific coast of

    Tohoku, Japan with epicenter at 38.1035 N, 142.861 E was reported by Japan Mete-

    orological Agency (JMA) at 14:46:18 local time. The earthquake subsequently caused

    one of the most destructive tsunamis in human history, engulfed part of the eastern

    coast of Japan. Research has been reported for different aspects of this tsunami event

    since2011, for example [Fujii et al.,2011;Lay et al.,2011;Wei et al.,2012;Hooper

    et al.,2013;Melgar and Bock,2013;Satake et al., 2013], due to rich source of data

    availability and important scientific value. In this section, the present adaptive grid

    1550016-20

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    21/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    Fig. 14. (Color online) Tsunami wave approaching Monai valley: computed maximum runup att= 16.9 s around Monai valley. The numbers on the contours denote the elevation of the topography(m); The red line is the computed shoreline.

    Table 6. Tsunami wave approaching Monai valley: RMSE () at gauges andrelative computational cost.

    Gauges G5 G7 G9 Relative CPU time

    RMSE() on adaptive grid [mm] 0.414 0.927 0.416 1.00RMSE () on uniform grid [mm] 0.423 0.881 0.395 3.56

    tsunami model is used to reproduce the tsunami wave propagation for this event to

    demonstrate its robustness for field-scale application.

    On a background uniform grid of 1600 m resolution, dynamic grid adaption canbe carried out up to a resolution of 400 m in the computational domain as shown

    in Fig. 15, according to local flow features. The bathymetry data for the 1600 m

    and 400 m cells are generated from the available 450 m and 1350 m datasets. A

    rectangular fault model assuming static deformation of the Ocean floor is applied

    to process the initial fault parameters to generate water surface deformation that

    initialized the tsunami [Okada,1985]. The initial fault parameters reported byFujii

    et al. [2011] are used and the processed initial water surface level is illustrated

    in Fig. 15, where the origin (0, 0) locates at (137.732 E, 31.5939 N). The initial

    1550016-21

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    22/30

    Q. Liang, J. Hou & R. Amouzgar

    Fig. 15. 2011 Tohoku tsunami propagation simulation: locations of wave gauges and initial tsunamiwaves [Fujii et al., 2011] (unit: m).

    ocean surface disturbance then generates a series of tsunami waves propagating

    outward, which is simulated using the proposed adaptive grid-based model in this

    work. During the simulation, constant Manning coefficient of 0.025 is imposed and

    sa and coar are specified to be 0.17 and 0.8.

    The computed process of tsunami propagation is demonstrated in Fig. 16, from

    the elapsing time of earthquake up to 20 min. The initial tsunami wave starts to

    propagate radially in the ocean, approaching the east coast of Japan where the

    first wave arrives the near-shore gauges after about 20 min. Qualitative comparison

    with the animation provided in the supplementary material of Wei et al. [2012]

    indicates that the current adaptive grid model correctly predicts the overall tsunamipropagation process.

    The change of wave surface is also recorded and compared with field measure-

    ments. In this study, measured data is available at 15 gauges as shown in Fig. 15(a),

    including 6 nearshore GPS buoys, 6 NOWPHAS (The Nationwide Ocean Wave infor-

    mation network for Ports and HArbourS, Japan) wave gauges close to the coast, 2

    cabled pressure-gauges and 1 DART buoy (Deep-ocean Assessment and Reporting

    of Tsunamis) system in offshore about 500 km away from the epicenter, as detailed

    in Table7.

    1550016-22

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    23/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    (a) (b)

    (c) (d)

    Fig. 16. 2011 Tohoku tsunami propagation simulation: computed wave propagation at: (a) t =5 min, (b)t = 10 min, (c) t= 15 min, (d) t= 20 min (unit: m).

    From these field records, the maximum tsunami wave amplitude is approximately

    6 m at 804 and 802 (Iwate South), located in the sea with the water depths of 200 m

    and 204 m, respectively. Looking at that GPS buoys plots in Figs.17and18, it is

    observed that the model reasonably reproduced the first leading wave in most of

    1550016-23

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    24/30

    Q. Liang, J. Hou & R. Amouzgar

    Table 7. 2011 Tohoku tsunami propagation simulation: gauges under consideration.

    Gauge name Type Depth (m) N Latitude() E Longitude()

    807-Iwate North GPS buoy 125 40.11667 142.06667

    804-Iwate Central GPS buoy 200 39.62722 142.18667802-Iwate South GPS buoy 204 39.25861 142.09694803-Miyagi North GPS buoy 160 38.85778 141.89444801-Miyagi Central GPS buoy 144 38.2325 141.68361806-Fukushima GPS buoy 137 36.97139 141.18556613 NOWPHAS wave gauge 50 42.8844 144.410602 NOWPHAS wave gauge 50.7 42.5427 141.441202 NOWPHAS wave gauge 43.8 40.9347 141.415203 NOWPHAS wave gauge 27.7 40.5528 141.546219 NOWPHAS wave gauge 49.5 40.2111 141.862205 NOWPHAS wave gauge 21.3 38.2412 141.047D2148 Deep ocean tsunameter 5660 38.8210 148.6550TM1 Pressure gauge 1600 39.1855 142.767

    TM2 Pressure gauge 1000 39.2036 142.444

    the gauges, as well as the wave series that arrived within the following 5 h affected

    by refraction and reflection. A proper match of the arrival times and amplitudes

    proves the correct estimation of the position and initial ocean surface deformation

    resulting from the earthquake dynamics. However, the source estimation involves

    the major uncertainty of the simulation that may lead to discrepancies between

    numerical predictions and field measurements, especially in the wave series after

    the leading wave. Tsunami waves propagating toward the shore also recorded by

    a number of NOWPHAS (The Nationwide Ocean Wave information network forPorts and HArbourS, Japan) wave gauges deployed in 20 to 50 m deep water mostly

    in the nearshore area north of Tohoku and south of Hokkaido (Figs. 17 and 18).

    Comparison between the measured and computed water levels at these wave gauges

    shows a good agreement in wave amplitudes, phase and periods, indicating that the

    current model can produce reliable numerical prediction in the nearshore areas. In

    those deep water gauges, the measurement for TM1 and TM2 are reported up to

    30 min as shown in Fig.18. The recorded peak is about 5 m after about 18 min of

    earthquake. The model successfully reproduces such wave forms that compare well

    with the measurements although wave amplitude is slightly underestimated. The

    main wave features predicted by the current model also agree with those predictedby an alternative model by Fujii et al. [2011]. Only one tsunameter (D21418) in

    eastern deep ocean is considered, located about 500 km east of the epicenter with

    the water depth of 5660 m. It measured a peak of 1.64 m approximately 33 min after

    the earthquake. As plotted in Fig.18(g), the wave amplitude, form, arrival time and

    second depression are well-predicted by the current model. However, the amplitude

    of the arrival wave is again somehow underestimated. The reason might lie in the

    sub-fault positions and the amplitude of the instantaneous uplift of the sub faults

    near to the epicenter does not account for the progressive fault procedure.

    1550016-24

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    25/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    (a) (b)

    (c) (d)

    (e) (f)

    (g) (h)

    Fig. 17. 2011 Tohoku tsunami propagation simulation: computed and measured water levelevolution: I.

    1550016-25

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    26/30

    Q. Liang, J. Hou & R. Amouzgar

    (a) (b)

    (c) (d)

    (e) (f)

    (g)

    Fig. 18. 2011 Tohoku tsunami propagation simulation: computed and measured water level evolu-tion: II [Fujii et al., 2011].

    1550016-26

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    27/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    Table 8. 2011 Tohoku tsunami propagation simulation: RMSE ().

    Gauges 804 801 803 806

    RMSE () on adaptive grid [m] 1.441 1.736 1.743 0.726

    RMSE () on uniform grid [m] 1.498 1.889 1.635 0.733

    To further demonstrate the performance of the adaptive grid model, the results

    computed by the model on a uniform gird with the resolution of 400 m are compared

    at gauges 804, 801, 803 and 806 (Fig. 17). Predictions on the two grid systems

    appears to be very similar, which is confirmed by the calculated RMSE () listed in

    Table8,indicating that the key wave features are well captured at high-resolution

    by the adaptive grid. The computation on the adaptive grid is 2.87 times as fast as

    that on the uniform one.

    5. Discussions and Conclusions

    A dynamically adaptive grid-based 2D shallow water equation model is presented in

    this work for tsunami simulation. The model is validated against three laboratory-

    scale benchmark tests and then applied to reproduce a field event. Numerical results

    compare satisfactorily with experimental measurements and field data wherever

    available, demonstrating the models capability in providing reliable and efficient

    predictions of tsunami propagation over complex domains.

    The model solves the 2D shallow water equations using a finite volume Godunov-

    type scheme implemented on the current adaptive grid system. Simulation resultsshow that the second-order Godunov-type scheme is able to reliably predict the

    amplitude and arrival times of the tsunami waves. It should be noted that the

    proposed model is designed for simulation of long and non-breaking waves, due to

    the restriction of the shallow water equations (the ratio between the wave length

    and height should be higher than 20 [Hinkelmann, 2005]). The model performs

    well when simulating tsunami propagation in relatively deep oceans where long

    wave assumption is valid, as shown from the simulation results for the 2011 Tohoku

    tsunami. However, in the nearshore zones where wave dispersion becomes important,

    a tsunami model based on the shallow water equations may not be able to provide

    satisfactory predictions due to its inherent limitation in resolving wave dispersion.Under such circumstances, a model based on Bousinesq equations may be more

    appropriate and further improvement to the current model is needed.

    On the current simplified adaptive grid system, dynamic adaptation is achieved

    by increasing or reducing the subdivision levels of the coarse background cells accord-

    ing to certain criteria reflecting local flow features, e.g. water surface gradient and

    wetting-drying condition as used in this work. Since no data structure is actually

    required to store grid information, the adopted grid system provides flexibility for

    grid manipulation and is straightforward to implement comparing to other adaptive

    1550016-27

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    28/30

    Q. Liang, J. Hou & R. Amouzgar

    grid system, e.g. quadtree grid. As demonstrated, the current adaptive grid-based

    model is able to capture the key features of the tsunami waves, e.g. steep wave

    fronts and wet-dry interfaces, where localized high-resolution meshes are essential

    for accurate predictions. By intelligently allocating computational resources by cre-

    ating local refined meshes, the current adaptive grid-based tsunami model is able toprovide computationally efficient simulations without compromising solution accu-

    racy, as confirmed by the comparison of the numerical results with those predicted

    by a uniform grid model based on similar numerical scheme. Therefore, the present

    adaptive grid-based model provides an alternative for tsunami simulations.

    Acknowledgment

    This work is supported by the National Natural Science Foundation of China (Grant

    No. 51379074), the Royal Society under the International Exchanges 2013 NSFC costshare scheme (IE131297) and the Chinese Government through the Recruitment

    Program of Global Experts. The authors also give big thanks to Dr Tomohiro

    Yasuda and Professor Hajime Mase from the Disaster Prevention Research Institute

    at Kyoto University for providing data to support the simulations of the 2011 Tohoku

    tsunami.

    References

    Abadie, S. M., Harris, J. C., Grilli, S. T. & Fabre, R. [2012] Numerical modeling of tsunami waves

    generated by the flank collapse of the Cumbre Vieja Volcano (La Palma, Canary islands):Tsunami source and near field effects, J. Geophys. Res. 117, C05030.

    Arcos, M. E. M. & LeVeque, R. J. [2014] Validating velocities in the geoclaw tsunami model usingobservations near hawaii from the 2011 Tohoku tsunami, Pure Appl. Geophys 172, 849867.

    Berger, M. J. & LeVeque, R. J. [1998] Adaptive mesh refinement using wave-propagation algo-rithms for hyperbolic systems,SIAM J. Numer. Anal. 35, 22982316.

    Berger, M. J. & Oliger, J. [1984] Adaptive mesh refinement for hyperbolic partial differentialequations, J. Comput. Phys. 53(3), 484512.

    Bradford, S. F. & Sanders, B. F. [2002] Finite-volume model for shallow-water flooding of arbitrarytopography, J. Hydraul. Eng. 128, 289298.

    Briggs, M., Synolakis, C., Harkins, G. & Green, D. [1995] Laboratory experiments of tsunamirunup on a circular island, Pure Appl. Geophys. 144, 569593.

    Bussing, T. R. & Murmant, E. M. [1988] Finite-volume method for the calculation of compressiblechemically reacting flows, AIAA J. 26, 10701078.

    Choi, B. H., Kim, D. C., Pelinovsky, E. & Woo, S. B. [2007] Three-dimensional simulation oftsunami run-up around conical island, Coastal Eng. 54, 618629.

    Clain, S. & Clauzon, V. [2010] L stability of the MUSCL methods, Numer. Math. 116, 3164.Courant, R., Friedrichs, K. & Lewy, H. [1928] Ueber die partiellen differenzengleichungen der

    mathematischen physik, Math. Ann. 100(1), 3274.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 Space63(7), 815820.Funke, S., Pain, C., Kramer, S. & Piggott, M. [2011] A wetting and drying algorithm with a

    combined pressure/free-surface formulation for non-hydrostatic models, Adv. Water Resources34, 14831495.

    1550016-28

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    29/30

    Simulation of Tsunami Propagation Using Adaptive Cartesian Grids

    George, D. L. & LeVeque, R. J. [2008] High-Resolution Methods and Adaptive Refinement forTsunami Propagation and Inundation, Hyperbolic Problems: Theory, Numerics, Applications(Springer-Verlag, Berlin, Heidelberg).

    Hinkelmann, R. [2005] Efficient Numerical Methods and Information-Processing Techniques forModeling Hydro- and Environmental Systems (Springer, Berlin, Heidelberg).

    Hooper, A., Pietrzak, J., Simons, W., Cui, H., Riva, R., Naeije, M., Terwisscha van Scheltinga,A., Schrama, E., Stelling, G. & Socquet, A. [2013] Importance of horizontal seafloor motionon tsunami height for the 2011 mw = 9.0 tohoku-oki earthquake, Earth Planetary Sci. Lett.361, 469479.

    Horrillo, J., Wood, A., Kim, G.-B. & Parambath, A. [2013] A simplified 3-D NavierStokes numer-ical model for landslide-tsunami: Application to the gulf of Mexico, J. Geophys. Res. Oceans118(12), 69346950.

    Hou, J., Simons, F., Mahgoub, M. & Hinkelmann, R. [2013] A robust well-balanced model onunstructured grids for shallow water flows with wetting and drying over complex topography,Computer Methods Appl. Mech. Eng. 257, 126149.

    Hubbard, M. E. & Dodd, N. [2002] A 2d numerical model of wave run-up and overtopping,Coastal Eng. 47, 126.

    Imamura, F. [1996] Review of tsunami simulation with a finite difference method, in Long-WaveRunup Models, eds. Yeh, P. L. H. & Synolakis, C. (Word Scientific Publishing Co., Singapore),pp. 2542.

    Ji, H., Lien, F.-S. & Yee, E. [2010] A new adaptive mesh refinement data structure with anapplication to detonation, J. Comput. Phys. 229(23), 89818993.

    Kazolea, M., Delis, A., Nikolos, I. & Synolakis, C. [2012] An unstructured finite volume numericalscheme for extended 2D Boussinesq-type equations, Coastal Eng. 69, 4266.

    Lay, T., Yamazaki, Y., Ammon, C. J., Cheung, K. F. & Kanamori, H. [2011] The 2011 Mw 9.0off the pacific coast of tohoku earthquake: Comparison of deep-water tsunami signals withfinite-fault rupture model predictions, Earth Planets Space 63(7), 797801.

    Lee, W.-K., Borthwick, A. G. & Taylor, P. H. [2011] A fast adaptive quadtree scheme for a two-layer shallow water model, J. Comput. Phys. 230(12), 48484870.

    Liang, Q. [2011] A structured but non-uniform cartesian grid-based model for the shallow waterequations, Int. J. Numer. Methods Fluids66(5), 537554.

    Liang, Q. [2012] A simplified adaptive cartesian grid system for solving the 2D shallow waterequations, Int. J. Numer. Methods Fluids69(2), 442458.

    Liang, Q. & Borthwick, A. G. [2009] Adaptive quadtree simulation of shallow flows with wet dryfronts over complex topography, Computers Fluids38, 221234.

    Liang, Q., Hou, J. & Xia, X. [2015] Contradiction between the c-property and mass conservationin adaptive grid based shallow flow models: Cause and solution, Int. J. Numer. Methods Fluids78(1), 1736.

    Liang, Q. & Marche, F. [2009] Numerical resolution of well-balanced shallow water equations withcomplex source terms, Adv. Water Resources 32, 873884.

    Liu, P. L. F., Cho, Y.-S., Briggs, M. J., Kanoglu, U. & Synolakis, C. E. [1995] Runup of solitarywave on a circular island, J. Fluid Mech. 302, 259285.

    Liu, P. L.-F., Yeh, H. & Synolakis, C. eds. [2008] Advanced Numerical Models for SimulatingTsunami Waves and Runup (World Scientific Publishing Co., Inc., Singapore).

    Lynett, P. J., Wu, T.-R. & Liu, P. L. F. [2002] Modeling wave runup with depth-integratedequations, Coastal Eng. 46, 89107.

    Matsuyama, M. & Tanaka, H. [2001] An experimental study of the highest run-up height in1993 Hokkaido Nansei-oki earthquake tsunami, in Proc. Int. Tsunami Symp., Seattle, U.S.A.,pp. 879889.

    Melgar, D. & Bock, Y. [2013] Near-field tsunami models with rapid earthquake source inversionsfrom land- and ocean-based observations: The potential for forecast and warning, J. Geophys.Res. Solid Earth118(11), 59395955.

    Nicolsky, D. J., Suleimant, E. N. & Hansen, R. A. [2011] Validation and verification of a numericalmodel for tsunami propagation and runup, Pure Appl. Geophys. 168, 11991222.

    1550016-29

  • 7/24/2019 Simulation of Tsunami Propagation.pdf

    30/30

    Q. Liang, J. Hou & R. Amouzgar

    Nikolos, I. & Delis, A. [2009] An unstructured node-centered finite volume scheme for shallowwater flows with wet/dry fronts over complex topography, Computer Methods Appl. Mech.Eng. 198, 37233750.

    Oishi, Y., Imamura, F. & Sugawara, D. [2015] Near-field tsunami inundation forecast using theparallel tsunami-n2 model: Application to the 2011 tohoku-oki earthquake combined with source

    inversions,Geophys. Res. Lett. 42(4), 10831091.Okada, Y. [1985] Surface deformation due to shear and tensile faults in a half-space, Bull. Seis-

    mological Soc. Am. 74, 11351154.Popinet, S. [2003] Gerris: A tree-based adaptive solver for the incompressible Euler equations in

    complex geometries, J. Comput. Phys. 190(2), 572600.Popinet, S. [2011] Quadtree-adaptive tsunami modelling, Ocean Dyn. 61, 12611285.Popinet, S. [2012] Adaptive modelling of long-distance wave propagation and fine-scale flooding

    during the Tohoku tsunami, Nat. Hazards Earth Syst. Sci. 12, 12131227.Rogers, B., Fujihara, M. & Borthwick, A. G. L. [2001] Adaptive q-tree godunov-type scheme for

    shallow water equations, Int. J. Numer. Methods Fluids35(3), 247280.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), 14731492.Synolakis, C. E. [1986] The runup of long waves, PhD Thesis, California Institute of Technology,

    Pasadena, California.Synolakis, C. E., Bernard, E. N., Titov, V. V., Kanoglu, U. & Gonzalez, F. I. [2007] Standards,

    criteria, and procedures for NOAA evaluation of tsunami numerical models, NOAA Techni-cal Memorandum OAR PMEL-135. NOAA/Pacific Marine Environmental Laboratory, Seattle,WA.

    Tang, L., Titov, V. V., Bernard, E. N., Wei, Y., Chamberlin, C. D., Newman, J. C., Mofjeld, H. O.,Arcas, D., Eble, M. C., Moore, C. et al. [2012] Direct energy estimation of the 2011 Japantsunami using deep-ocean pressure measurements, J. Geophys. Res. 117, C08008.

    Titov, V. V. & Synolakis, C. E. [1995] Modeling of breaking and nonbreaking long-wave evolutionand runup using vtcs-2, J. Waterway Port Coastal Ocean Eng. 121, 308316.

    Titov, V. V. & Synolakis, C. E. [1998] Numerical modeling of tidal wave runup, J. WaterwayPort Coastal Ocean Eng. 124, 157161.

    van Leer, B. [1984] On the relation between the upwind-differencing schemes of godunov, engquistosher and roe, SIAM J. Sci. Stat. Comput. 5, 120.

    Wang, X. & Liu, P. L.-F. [2006] An analysis of 2004 Sumatra earthquake fault plane mechanismsand Indian ocean tsunami, J. Hydraul. Res. 44(2), 147154.

    Watanabe, Y., Mitobe, Y., Saruwatari, A., Yamada, T. & Niida, Y. [2012] Evolution ofthe 2011 Tohoku earthquake tsunami on the Pacific coast of Hokkaido, Coast. Eng. J.54(01), 1250002.

    Wei, Y., Bernard, E. N., Tang, L., Weiss, R., Titov, V. V., Moore, C., Spillane, M., Hopkins, M. &Kangolu, U. [2008] Real-time experimental forecast of the peruvian tsunami of August 2007for U.S. coastlines, Geophys. Res. Lett. 35(4), L04609.

    Wei, Y., Chamberlin, C., Titov, V. V., Tang, L. & Bernard, E. N. [2012] Modeling of the 2011

    Japan tsunami: Lessons for near-field forecast, Pure Appl. Geophys. 170(68), 13091331.Yiu, K., Greaves, D., Cruz, S., Saalehi, A. & Borthwick, A. [1996] Quadtree grid generation:

    Information handling, boundary fitting and CFD applications, Computers Fluids25(8), 759769.

    Zhang, Y. J. & Baptista, A. M. [2008] An efficient and robust tsunami model on unstructuredgrids. part i: Inundation benchmarks, Pure Appl. Geophys. 165(1112), 22292248.

    Zhou, J. G., Causon, D. M., Ingram, D. M. & Mingham, C. G. [2002] Numerical solutions of theshallow water equations with discontinuous bed topography, Int. J. Numer. Methods Fluids38(8), 769788.

    Zijlema, M., Stelling, G. & Smit, P. [2011] SWASH: An operational public domain code for simu-lating wave fields and rapidly varied flows in coastal waters, Coastal Eng. 58(10), 9921012.