Algo for Modelling (Pre)2005

download Algo for Modelling (Pre)2005

of 34

Transcript of Algo for Modelling (Pre)2005

  • 8/2/2019 Algo for Modelling (Pre)2005

    1/34

    An algorithm for modeling the interaction of a flexible rod with a

    two-dimensional high-speed flow

    D. Tam1, R. Radovitzky1, and R. Samtaney2

    1 Department of Aeronautics and Astronautics, Massachusetts Institue of Technology, Cambridge, MA,

    U.S.A.

    2 Princeton Plasma Physics Laboratory, Princeton University, Princeton, NJ, U.S.A.

    SUMMARY

    We present an algorithm for modeling coupled dynamic interactions between very thin flexible

    structures immersed in a high-speed flow. The modeling approach is based on combining an Eulerian

    finite volume formulation for the fluid flow and a Lagrangian large-deformation formulation for the

    dynamic response of the structure. The coupling between the fluid and the solid response is achieved via

    an approach based on extrapolation and velocity reconstruction inspired in the Ghost Fluid Method.

    The algorithm presented does not assume the existence of a region exterior to the fluid domain

    as it was previously proposed and, thus, enables the consideration of very thin open boundaries and

    structures where the flow may be relevant on both sides of the interface. We demonstrate the accuracy

    of the method and its ability to describe disparate flow conditions across a fixed thin rigid interface

    without pollution of the flow field accross the solid interface by comparing with analytical solutions

    of compressible flows. We also demonstrate the versatility and robustness of the method in a complex

    Correspondence to: Department of Aeronautics and Astronautics, Massachusetts Institue of Technology,

    Cambridge, MA, 02139, U.S.A.

  • 8/2/2019 Algo for Modelling (Pre)2005

    2/34

    1

    fluid-structure interaction problem corresponding to the transient supersonic flow past a transverse,

    highly flexible structure.

    Copyright c 2004 John Wiley & Sons, Ltd.

    key words: fluid-solid interaction, compressible flows, flexible structures

    1. Introduction

    Current and future interplanetary exploration missions demand the availability of numerical

    tools for the design of light structures such as gossamer spacecraft and parachutes, [ 1, 2, 3]. In

    many situations of interest, an adequate description of the continuum fields in both the fluid

    flow and the solid structure dynamic deformations, as well as of their coupled interactions, is

    necessary. In this work we propose a computational strategy for modeling the coupled response

    of a thin structure immersed in a supersonic flow.

    In general, the dynamic deformation of solid structures is most adequately described in a

    Lagrangian framework, especially in the case of large deformations. The main advantage of the

    Lagrangian approach lies in its natural ability to track the evolution of properties at material

    points in materials with history, as well as in the treatment of boundary conditions at material

    surfaces such as free boundaries or fluid-solid interfaces. In contrast to Eulerian approaches,

    boundary conditions are enforced at material surfaces ab initio and therefore require no special

    attention. In this work, we propose a Lagrangian formulation for describing the large dynamic

    deformations of two-dimensional thin structures (rods) having both bending and membranal

    stiffness.

    By contrast, Lagrangian formulations are inadequate in the case of high-speed flows or

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    3/34

    2

    flows involving significant vorticity due to the unavoidable mesh distortion incurred during

    deformation which reduces the stable time step and the overall accuracy of the simulation

    and, eventually, breaks the numerical method. This problem can be partially remedied by the

    use of remeshing [4]. However, remeshing increases the complexity of the algorithm and of its

    implementation and suffers from robustness problems in the three-dimensional case. Eulerian

    approaches, in which the field equations are formulated in terms of spatial variables and fixed

    or adaptivealbeit not distortingmeshes, are more adequate for most fluid flows. We concern

    ourselves with flows where the viscous time scales far exceed the convection time scales, i.e.

    we model the fluid flow with the compressible Euler equations. In this work, the supersonic,

    unsteady flow conditions are modeled by recourse to a finite volume formulation of the Euler

    equations of compressible flow following Samtaney et al [5, 6].

    A number of different strategies for coupling fixed-grid Eulerian fluid dynamics formulations

    with Lagrangian solid mechanics formulations have been proposed. For incompressible viscous

    flows, the immersed boundary method of Peskin and McQueen [7] has received significant

    attention, especially owing to its success in modeling the complex conditions of blood flow in

    the heart. A recent review of the method may be found in Peskin [ 8]. Several extensions of

    this method have been recently proposed by Liu [9].

    Our work is concerned with problems involving high-speed compressible flows. For this type

    of problems, the so-called Cartesian boundary method [10, 11] and the Embedded boundary

    approach of Colella et al [12] have recently gained significant popularity. In these approaches,

    the computational domain is discretized by rectangular finite volume cells and the geometry

    is represented by intersections with the underlying Cartesian grid. This leads to cut-cells

    in those grid locations where the boundary intersects the grid. A detailed description of this

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    4/34

    3

    approach, along with some issues related to the unavoidable appearance of small cells, is given

    by Colella [12] and references therein. An alternative approach that explicitly avoids these

    issues from the outset by replacing the imposition of boundary conditions with an approach

    based on field extrapolation into exterior ghost cells has been proposed [13, 14, 15, 16, 17]. This

    class of methods is inspired in the Ghost Fluid Method of Fedkiw et al [ 18]. The convergence

    properties of this Eulerian-Lagrangian coupling approach have been carefully studied by

    Arienti et al [17]. A similar treatment of irregular boundaries in cartesian grid approaches

    including second order accurate formulation of boundary conditions has been recently given

    by Sussman [19, 20].

    The algorithms presented in the references above are adequate for flows interacting with

    bulk solids [13, 14] or thin shells [21, 22]. However, in the case of thin shells they are limited

    to situations in which the shell is closed and flow takes place only on one side of it. This

    restriction is imposed by the assumption that the fluid domain has a well-defined interior and

    exterior, which is the basis of the coupling algorithm based on level sets.

    In this work, we extend this approach to the case of thin structures that at the same time are

    open and in which the flow on both sides of the structure may be relevant. These situations

    arise in important applications such as the deployment of parachutes used as decelaration

    devices during planet entry in space exploration missions, [1, 2, 3], The extended approach

    retains the basic concepts of the original algorithm, but allows an unbiased consideration of the

    flow conditions on both sides of the immersed structure as well as an adequate treatment of the

    boundary conditions on both sides of the boundary. Among the advantages of the approach one

    finds its simplicity, robustness and ease of implementation especially considering the minimal

    modifications required in each solver. Another advantage of this class of fluid-solid coupling

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    5/34

    4

    methods is their suitability for parallel implementation. In [14, 22], the three-dimensional

    parallel implementation of this class of fluid-solid coupling algorithms was demonstrated,

    including scalability properties on up to 1856 processors. For simplicity, we restrict our

    attention to the two dimensional problem.

    In the following sections we first present the formulation and numerical approach for

    describing large dynamic deformations of a thin rod structure. This is followed by a review of

    the numerical method adopted for the fluid. Subsequently, we describe the fluid-solid coupling

    algorithm and the proposed extension to thin open immersed structures. The last section of the

    paper is devoted to establishing the feasibility and properties of the method. We first present

    verification simulations and a convergence study corresponding to the supersonic flow past a

    very thin flat rigid boundary at different angles of attack. These simulations demonstrate the

    ability and accuracy of the proposed approach to describe the flow on both sides of a very thin

    structure. We finally present a fully coupled simulation of a supersonic flow initially normal to

    a flexible structure which demonstrates the versatility and robustness of the overall method in

    simulating complex fluid-structure interaction problems.

    2. Large-displacement rod dynamics model

    In this section, we briefly summarize the model adopted for describing the dynamics of slender

    rods. For more general beam or slender rod models, the vast literature on the subject, which

    traces its origins to the work of Euler and Bernoulli [23], may be consulted, see for example

    [24, 25, 26] and references therein. A representative rod element is shown in Figure 1. The rod

    element is allowed to undergo a motion consisting of a finite rotation, a finite uniform stretch

    and a small bending distortion. With the conventions shown in this schematic, the deformation

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    6/34

    5

    Figure 1. Schematic describing the conventions and kinematics of the rod model proposed

    mapping of the rod element follows as:

    x1 l

    LX1 (X1)X2 (1)

    x2 X2 (2)

    and the stretch of the longitudinal fibers of the rod element follows as

    =x1X1

    l

    L (X1)X2 (3)

    The small bending distortion measured from the rotated and stretched configuration (axes

    x1,x2) is assumed to follow the classic Euler-Bernoulli hypothesis:

    w

    x1(4)

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    7/34

    6

    A strain energy density per unit undeformed area of the rod of the form:

    W() = E(1 + log) (5)

    is assumed, where E is the Youngs modulus. This energy density gives a linear relation between

    the nominal stress and the logarithmic or true strain.

    The strain energy of the rod is obtained by integrating equation (5) over the volume of the

    undeformed rod after inserting the assumed rod kinematics, equation (3), with the result

    U EA [L l + l log(l/L)] +EI

    2 l

    0

    2

    (x1)dx1 (6)

    where A and I are the area and moment of inertia with respect to the axis normal to the

    bending plane of the undeformed cross section of the rod. In evaluating the integral along

    the undeformed axis of the rod, a change of variables to the deformed configuration has been

    conveniently taken advantage of.

    Linear momentum balance is enforced weakly by recourse to Hamiltons principle for

    continuous media, i.e. by finding the paths between two arbitrary times t1 and t2 for which

    the action integral is stationary:

    t1t0

    Ldt = 0 (7)

    where L is the associated Lagrangian defined as L = K , K is the kinetic energy of the

    system, = U+ ext is the potential energy of the system and ext is the potential of the

    external forces. Equation (7) must hold for any variationally admissible virtual displacement.

    A complete derivation of Hamiltons principle for continuous systems may be found in standard

    references, see for example [27].

    We take Hamiltons Principle, Equation (7), as the basis for finite element discretization.

    The explicit derivation of the first variation of the action integral, Equation ( 7), leading to

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    8/34

    7

    the explicit expression of the continuous variational form is omitted for conciseness, since it

    is not necessary for the numerical formulation. This derivation for different linearized beam

    models may be found in standard textbooks, see for example Reddy [ 28]. We use hermitian

    cubic interpolation to represent w and its derivatives as a function ofx1. This automatically

    satisfies the requirement ofC1 interelement continuity ofw. In the finite element formulation

    proposed, the unknowns represent the physical displacements and rotation at extremity (node)

    a = 1, 2 of the rod element. Explicit expressions of the strain and kinetic energy of the rod

    element in terms of the degrees of freedom are derived in [29]. Upon spatial discretization, the

    stationarity condition (7) leads to the semi-discrete system of nonlinear ordinary differential

    equations:

    Mhxh +Finth (xh) = F

    exth (t) (8)

    In these expressions, Mh is the mass matrix, xh the array of nodal accelerations and

    Finth (xh) =U

    xh(9)

    Fexth = extxh(10)

    are the arrays of internal forces and time-varying external forces, respectively. In the fluid

    structure interaction problems of interest in this work, the array Fexth in equation (10)

    represents the external nodal forces equivalent to the traction boundary conditions imposed

    by the flow on the structure. The computation of these forces is discussed in section 4.

    The equations of motion (8) are integrated in time using Newmarks family of algorithms:

    xn+1 = xn + txn + t212 xn + xn+1

    xn+1 = xn + t

    (1 )xn + xn+1

    Mxn+1 +Fint(xn+1) = F

    ext(tn+1)

    (11)

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    9/34

    8

    where and are the Newmark algorithm parameters. For = 0, a conventional implicit

    predictor-corrector algorithm [30] is adopted to solve the system of equations (11), leading to

    the incremental nonlinear algebraic system:

    MU

    t2+ Fint(xn+1 +U) = Fextn+1 (12)

    where U = t2xn+1. A consistent linearization of this nonlinear algebraic equation about

    the predictor configuration leads to the computation of the tangent stiffness matrix:

    K =Fintx

    bxn+1

    (13)

    which enables a quadratic convergence of the Newton Raphson algorithm used to obtain

    dynamic equilibrium at t = tn+1. Explicit expressions for the mass matrix, array of internal

    forces and consistent tangent moduli in our slender rod model are derived in [ 29].

    3. Eulerian compressible fluid solver

    In this section we summarize the formulation of the fluid solver. Further details may be found

    in the original references. [31, 5, 6]. The flow is modeled as compressibe and inviscid, leading

    to the governing Euler equations of compressible flow. These equations may be expressed in

    the following strong conservative form:

    U,t + F,x(U) + G,y(U) = 0 (14)

    where

    U = ,u,v,ET

    F(U) =u, u2 +p, uv, (E+

    p

    )uT

    G(U) =v,uv,v2 +p, (E+

    p

    )vT

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    10/34

    9

    where is the density, u and v are the Cartesian components of the velocity vector, p is the

    pressure, E is the specific total energy, U is the vector of conservative variables, F(U) and

    G(U) are the components of the flux vector. An additional equation of state closes the system

    of equations. In this work, the equation of state of perfect gases:

    p = ( 1)e (15)

    is adopted, where is the specific heat ratio and e is the specific internal energy with

    E= e + 12u2.

    A finite volume formulation is adopted as the numerical approximation of these equations.

    The discretized equations may be written as:

    Ui,jt

    =Fi 1

    2,j Fi+ 1

    2,j

    h+Gi,j1

    2 Gi,j+ 1

    2

    h(16)

    where Ui,j is an average value ofU over the (i, j)th cell, and Fi 1

    2,j , Fi+ 1

    2,j , Gi,j 1

    2and Gi,j+ 1

    2

    are the fluxes at the cell interfaces. This formulation is numerically conservative and thus, the

    variation ofU over one cell of the mesh is equal to the inward flux. For stability reasons, flows

    with strong compressibility effects leading to the formations of shocks are best modeled by a

    conservative formulation [32, 33].

    The fluxes at the cell interfaces may be calculated either by the Equilibrium Flux Method

    (EFM) (a kinetic flux vector splitting scheme) [31], or the Godunov [34] or Roe method [35] (a

    flux difference splitting scheme). These discretization schemes are first order in space and can

    be taken as a starting point for the formulation of higher order schemes. In our case, second

    order accuracy is achieved via linear reconstruction with Van Leer type slope limiting applied

    to projections in characteristic state space. This method is often referred to as the MUSCL

    approach (Monotone Upstream-centered Schemes for Conservation Laws) [36, 33, 37, 38].

    Equations (16) are integrated explicitly in time using the second-order Runge-Kutta algorithm:

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    11/34

    10

    First step:

    Un+ 1

    2

    i,j = Uni,j +

    t2h

    [Fi1/2,j(Un) Fi+1/2,j(U

    n)

    + Gi,j1/2(Un) Gi,j+1/2(U

    n)] (17)

    Second step:

    Un+1i,j = Uni,j +

    t

    h[Fi1/2,j(U

    n+ 12 ) Fi+1/2,j(U

    n+ 12 )

    + Gi,j1/2(Un+ 1

    2 ) Gi,j+1/2(Un+ 1

    2 )] (18)

    The fluid solver imposes restrictions on the stable time step given by the Courant-Friedrichs-

    Levy (CFL) stability condition [32, 38, 33]. The resulting fluid model is second order in time

    and space. Details of the parallel implementation of this algorithm, including adaptive mesh

    refinement capability may be found in Ref. [39].

    4. Eulerian-Lagrangian coupling algorithm

    Our objective in this work is to develop a fluid-structure coupling algorithm with the ability

    to describe situations in which the details of the flow on both sides of a very thin structure

    are of equal importance. Such situations arise, for example, when the structure is a manifold

    with boundary (e.g. an open shell) or, if the manifold is closed, when there is fluid in the

    shell interior as well as in the exterior. This objective is relatively easy to achieve with an

    unstructured mesh, finite element, Arbitrary Lagrangian Eulerian (ALE) formulation [ 40] or

    with other alternative mesh moving techniques[41] in the case that the thin structure is fixed,

    i.e., in the case of a thin rigid boundary, or when the structure deformations are relatively

    small. However, when the deformations are large, these methods usually suffer from stability

    [42] and excessive mesh distortion problems [43].

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    12/34

    11

    In previous work, [16], we have focused on flow geometries with closed boundaries and a well

    defined exterior region. In this work we extend this algorithm in a way that this restriction

    can be eliminated. For simplicity, we restrict our attention to the two-dimensional case.

    The Eulerian fluid solver and the Lagrangian solid solver are weakly coupled by applying

    appropriate boundary conditions at the fluid-solid interface at the beginning of each time step.

    Other possible implicit and staggering schemes in coupled systems have been proposed and

    studied in detail in [44, 45]. In the case of inviscid flows considered, these boundary conditions

    correspond to continuity of the normal component of the velocity field:

    [ v n ] = 0, on the fluid-solid boundary (19)

    where [ . ] represents field jumps, and continuity of the normal component of the traction across

    the fluid-solid interface:

    [ t n] = [ ijninj ] = [ n ] = 0, on the fluid-solid boundary (20)

    which enforce conservation of mass and linear momentum, respectively. For simplicity, heat

    transfer across the fluid-solid interface is neglected.

    The formulation of the algorithmic steps to enforce these conditions is described in the

    following. In the model proposed, we consider that the only aerodynamic force acting on the

    structure is due to the fluid pressure. Equation (20) is enforced weakly by directly applying

    the pressure exerted by the fluid on the structure at time tn as traction boundary conditions

    for time step tn+1 in a variationally consistent manner. This results in the following expression

    for the external force array Fexth in Equation (10):

    Fextia

    n+1=

    So2

    pnNanids (21)

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    13/34

    12

    where i is the degree of freedom, Na the shape function of node a and pn is the local value of

    the pressure, which is interpolated bilinearly from the computed flow field at time tn.

    On the fluid, mass, momentum and energy conservation at the boundary are enforced via

    extrapolation and a flow reconstruction step. First velocity, pressure and density cell averages

    from the physical fluid domain are extrapolated to a narrow band of ghost cells across the

    boundary. This extrapolation is done by advection in pseudo-time :

    q

    + n q = 0 (22)

    where q = (,p,u,v) is the array of extrapolated quantities and n is the normal to the interface

    computed from the level set function. We employ simple upwinding along the normal n to

    march forward in in the extrapolation step above. When the steady state (q/ = 0)

    is reached, the non-physical extrapolated velocity field in the ghost cells is reconstructed

    according to the expression:

    vF = (2vS vF) n

    n + (vF t)t (23)

    where vF is the fluid velocity extrapolated from the active fluid cells and vS is the velocity of

    the solid interface. The resulting ghost velocities correspond to a reflection of the normal

    component of the fluid velocity relative to the moving boundary, whereas the tangential

    component is left unchanged. The fluxes thus computed from real and ghost values at the

    boundary, see section 3, in effect, enforce Equation (19). In the case of flows with a well-

    defined exterior domain, ghost and real flow values can be supported on the same grid. By

    contrast, in the case of open boundaries a separate data structure is required to store the

    extrapolated ghost values, as ghost and real fluid regions overlap. To this end, two arrays are

    used: one storing the real values of the conserved variables on the whole domain on which

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    14/34

    13

    the main computation takes place and another one storing the values of the extrapolated

    variables in the ghost fluid cells next to the interface, as shown in Figure 2, which is used for

    the application of boundary conditions in the fluid. It bears emphasis that significant storage

    savings may be achieved by adopting specialized sparse arrays [46] to store ghost values, which

    effectively reduce the dimensionality of the required storage size. It should be noted that the

    use of an additional array to store and access ghost data imposes some additionalalbeit

    straightforwardmodifications to the fluid solver, as compared with conventional ghost fluid

    approaches one of whose attractive features is the minor solver modifications required.

    Figure 2. Real fluid and ghost fluid arrays

    The location of the boundary as well as the boundary normal required to apply

    these boundary conditions, as described above, need to be computed efficiently to avoid

    computational bottlenecks. To this end, the level set function (x) which gives the minimum

    distance to the fluid-solid interface at each grid point of the fluid domain is used. The normal

    n to the interface is also interpolated directly from the level set function:

    n =

    (24)

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    15/34

    14

    The boundary is located where (x) = 0. The computation of the distance function at

    each time step is accomplished with an optimal algorithm developed by Mauch [47], whose

    complexity is O(m + n), where m is the number of elements in the shell mesh and n is the

    number of grid points in the subset of the fluid grid where the level set is required. In the

    case of flows with a well-defined exterior, as assumed in [13, 21], a sign is assigned to the level

    set function. (x) is taken as negative (positive) in the interior (exterior) of the physical fluid

    domain. This facilitates the immediate identification of real and ghost fluid cells. The main

    limitation of this approach is that it precludes the possibility of flows coexisting on both sides

    of a thin boundary.

    In the case of open boundaries, by contrast, interior and exterior cannot be defined. However,

    a key observation is that the boundary remains an orientable manifold, i.e., it has two

    unequivocally identifiablealbeit arbitrarily chosenpositive and negative sides which can

    be conveniently assigned to the adjacent fluid domain. Both in the two and three dimensional

    case, the manifold may be endowed with an orientation by a suitable choice of a specific

    parametrization, from which the surface tangent vector(s) and a positive normal may be

    defined. Based on this observation, the side of the boundary is identified on the fluid grid

    by endowing the distance function i,j with a sign next to the interface. This sign is obtained

    from thedimension independentformula:

    sign(i,j) =

    +1 if di,j n 0

    1 otherwise

    (25)

    where di,j is the distance vector from grid point i, j to the boundary and n is the local normal

    to the boundary. Thus, cells lying in the half space pointed to by the surface normal are

    assigned a positive sign and all other a negative one, as shown in Figure 3. Therefore, this

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    16/34

    15

    pseudo-signed level set function offers a straightforward manner to determine if two grid

    points lie on the same side of the boundary or not.

    Figure 3. Pseudo sign defined by the orientation of the interface

    Once the pseudo-signed distance function is computed, values from each side of fluid-solid

    interface are symmetrically extrapolated to the corresponding other side and stored on the

    auxiliary ghost array. The velocity fields are then reconstructed in the entire ghost region to

    impose non-penetration boundary condition, as described earlier.

    After the explicit application of boundary conditions by extrapolation in the fluid, Equations

    (22) and (23), and in the structure, Equation (21), at the beginning of the time step, time

    integration proceeds independently in each solver as described in sections 2 and 3. The

    computation of the fluxes in the fluid, Equation (16), for cells next to the fluid-solid boundary,

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    17/34

    16

    i.e. where i,j changes sign, needs to be done using both real and ghost values. The latter are

    obtained from the auxiliary array described above. Apart from this consideration, the solution

    is computed on the whole fluid grid, without the need of any additional special treatment of

    the boundary.

    One of the main advantages of this explicit coupling approach is its suitability for parallel

    implementation. The scalability properties of this coupling scheme on upwards of 1800

    processors has been reported in [14].

    It is important to remark that in order to avoid singularities stemming from the violation of

    the Sobolev cone condition at the ends of vanishing-thickness boundaries, structures embedded

    in the flow are endowed with a finite thickness, as real structures are expected to have.

    The staggering method adopted remains stable if the time step is chosen as:

    t min(tCFL, tb) (26)

    where tCFL is the stable time step for the fluid and:

    tb =min (x,y)

    vS(27)

    which prevents the solid boundary from crossing more than one fluid cell per time step. As

    explained in section 2, the structure does not impose additional time step restrictions related to

    stability, as the integration is done implicitly. The stability of different weak coupling schemes

    in coupled systems has been studied in detail in [45].

    The resulting fluid-structure coupling algorithm is summarized in Algorithm 1.

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    18/34

    17

    Algorithm 1. Fluid-Solid Coupling

    1. - Apply boundary conditions using solution from previous time-step:

    (a) For the fluid solver:

    i. Compute pseudo-signed distance function from updated location of solid boundary

    ii. Extrapolate flow field values to the ghost region using Equation (22) and store

    extrapolated fields in auxiliary ghost array.

    iii. Reconstruct velocity field in the ghost region using Equation (23)

    (b) For the solid solver:

    i. Interpolate pressure field at the interface directly from fluid grid

    ii. Apply pressure as external loading on the structure using Equation (21)

    2. - Compute stable time step:

    (a) Compute stable time step for the fluid solver tf according to CFL condition

    (b) Compute time step restrictions at the interface tb, Equation (27)

    (c) Adopt stable time step as: t = min(tf,tb), Equation (26)

    3. Integrate solution in time:

    (a) Integrate in time the fluid solution Second order Runge-Kutta integration, Equations (17),

    (18). Next to the boundary, access ghost values from auxiliary array to compute fluxes in

    Equation (16).

    (b) Integrate in time the solid solution using Newmarks algorithm, Equations (11).

    (c) Increment time-step in both fluid and solid: t tn + t

    (d) Update location of the interface, i.e., the reference configuration of the solid.

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    19/34

    18

    Figure 4. Schematic solution of the supersonic flow past a thin rigid plate

    5. Numerical examples

    5.1. Supersonic flow past a fixed thin plate

    The first example is intended to assess the ability of the extended ghost fluid method to

    describe supersonic flows past open thin boundaries. To this end, the flow past a fixed thin

    plate immersed in a high-speed flow at different angles of attack , see schematic in Figure 4,

    is computed using the numerical method described in the foregoing. This problem is amenable

    to analytical treatment [48, 49] and, therefore, provides a convenient means of assessing the

    accuracy and convergence properties of the extended ghost fluid method. The plate profile

    induces a weak shock attached to its leading edge on the side where the cross section of the

    flow decreases and an expansion wave on the opposite side. The pressure behind the shock and

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    20/34

    19

    the expansion wave is uniform and can be computed analytically as a function of using the

    Rankine-Hugoniot relations and, respectively, the Prandtl-Meyer function [49].

    Figure 5. Computed solution of the supersonic flow past a thin rigid plate. In the case shown, the Mach

    number is M = 1.8 and the angle of attack is = 15 degrees. Contours show pressure normalized

    with upstream value

    The Mach number adopted in these calculations is M = 1.8 and the initial properties of

    the gas are: p = 1.0atm, = 1.293 kgm3 , and = 1.4. The fluid domain is discretized with a

    400 680-cell fluid grid. In order to avoid singularities in the solution, the rigid boundary is

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    21/34

    20

    (a) y=-0.245 (b) y=0.095

    (c) y=0.275

    Figure 6. Comparison of numerical and analytical horizontal pressure profiles for different values of

    the vertical coordinate y. Analytical values are shown in thick gray lines. Numerical results are shown

    in thin black lines and + symbols. Values shown are normalized with upstream pressure.

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    22/34

    21

    given a finite thickness, as real structures are expected to have. The numerically-computed

    flow field for the case of = 15 degrees is shown in Figure 5. The contours indicate the value

    of the pressure normalized with its upstream value. Figures 6 (a)-(c) show comparisons of the

    normalized horizontal pressure profiles against the analytical solution for different values of

    the vertical coordinate. Figure 6 (b) corresponds to a cross section through the center of the

    plate (y = 0.095) and shows that the pressure values behind the shock and in the expansion

    region behind the plate are accurately computed up to the interface. It bears emphasis that

    there is no pollution of the solution from one side of the boundary to the other, which usually

    constitutes a challenge in methods based on extrapolation. Figures 6-(a) and (c) correspond

    to cross sections one grid cell away from the bottom (y = 0.245) and top (y = 0.275) tips of

    the rigid boundary, respectively. As it can be seen in these figures, the quality of the numerical

    solution is very good both far and near the rod ends.

    Several simulations were conducted for angles of attack ranging from = 5 to 18 degrees.

    Above 18 degrees and for an upstream Mach number of M = 1.8 the shock at the leading

    edge of the profile detaches. Figure 7 shows the variation of the pressure behind the shock

    as a function of the angle of attack. As expected, the pressure behind the shock increases

    with . The corresponding dependence of the pressure in the expansion region behind the

    plate on the angle of attack is shown in Figure 8. As is increased, the pressure behind the

    expansion wave decreases, as expected. The numerically computed values are plotted on the

    same Figures 7 and 8. In both cases, a very good agreement between the exact and numerical

    results is obtained.

    A critical aspect of fluid-structure interaction models is the ability to compute the

    aerodynamic loads on the structure with sufficient accuracy, as these loads determine the

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    23/34

    22

    Figure 7. Comparison of analytical and numerical values of the pressure behind the shock vs. angle of

    attack

    structural response. In the following, we study the convergence of the pressure load on the

    structure in the supersonic flow past a fixed plate problem. Other aspects of the convergence

    of the coupling approach based on the ghost fluid method were previously reported by Arienti

    et al [17].

    A series of simulations corresponding to the case of upstream Mach number M = 1.8 and

    angle of attack = 18 degrees is conducted for grid resolutions starting at 85 50. In each

    subsequent simulation, the resolution is increased by a factor of 2 in each direction. The

    finest grid resolution is 680 400. The exact nondimensional value of the aerodynamic lift

    (= 1.5759) for M = 1.8 and = 18 degrees is readily obtained from the difference between

    the analytical pressure values in the windward ( p1p = 2.5515) and leeward (p2p

    = 0.3453)

    sides, where p =1

    = 0.7143, and the normalized length of the flat plate (=1

    cos=1.051).

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    24/34

    23

    Figure 8. Comparison of analytical and numerical values of the pressure in the expansion region behind

    the plate vs. angle of attack

    The error in the computed lift, defined as the absolute value of the difference between the

    numerical and the analytical values normalized by the analytical value, as well as the rate of

    convergence, is reported in Table 5.1. The first order convergence rate obtained is attributed

    Grid resolution Computed value Error Error in % Convergence Rate

    85 x 50 1.4864950 0.08946629 5.6772 0.92

    170x100 1.5285353 0.04734034 3.0041 1.29

    340x200 1.5564916 0.01938427 1.2300 1.31

    680x400 1.5680707 0.00780513 0.4952 -

    Table I. Convergence analysis of the aerodynamic lift on the solid structure.

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    25/34

    24

    to the first order description of the geometry, and the first order interpolation of the fluid

    pressures on the boundary. Second order schemes for similar treatment of irregular boundaries

    in cartesian grids have recently been proposed [19, 20], but are more expensive in terms of

    CPU and memory. It can be concluded from these results that the algorithm proposed applies

    boundary conditions on both sides of the thin profile in a consistent manner and results in

    convergent pressure distributions caused by the flow on the solid boundary. It can therefore be

    expected that this, in turn, will result in correct traction boundary conditions on the structure

    in coupled simulations.

    5.2. Supersonic flow past a highly-flexible structure

    In this section, we demonstrate the versatility of the overall computational methodology in

    describing complex fluid solid interactions. The simulation corresponds to a supersonic flow

    transverse to an initially-flat structure made of an elastic fabric with a Youngs Modulus

    E = 6.0 109Pa and mass density = 1000.0 kgm3 . The length of the structure is 1.m and is

    discretized with 50 elements as described in section 2, its thickness is 3.0 103m, its cross

    sectional area A = 1.0 103m2 and its moment of inertia I = 2.25 109m4. A schematic of

    the simulation set up is provided in Figure 9.

    The initial properties of the gas are: upstream pressure p = 1.0atm, mass density

    = 1.293kgm3 , and = 1.4. The flows Mach number is M = 2.0. The size of the

    computational fluid domain is 5.20m9.60m and the grid resolution adopted in this calculation

    is 260 480 fluid cells.

    At first, the structure is held fixed and the steady-state flow around the flat structure is

    computed. A strong shock develops upstream of the structure. The highly flexible structure is

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    26/34

    25

    Figure 9. Schematic of simulation of supersonic flow past a flexible structure tranverse to the flow

    then released, except at its tips which are restrained horizontally, and starts inflating under the

    pressure of the flow, inducing complex interactions between the flow and the thin structure,

    see Figure 10.

    Figure 10 (a) shows the initial steady-state flow past the fixed structure. A strong shock

    forms in front of the structure which causes a very high (low) pressure pp 5.0 (p

    p 1.0) on

    the windward (leeward) side of the structure. When released, the structure starts to accelerate

    rigidly except at the extremities where the horizontal supports create a flexural wave which

    propagates towards the center, Figure 10 (b). The forward motion of the structure releases

    an expansion wave in the flow, which travels upstream towards the strong shock lowering the

    upstream pressure, Figures 10 (b)-(c). The flexural waves in the structure converge at its center

    at t = 2.80ms. As the structure deforms until it reaches a maximum deflection, the flow in

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    27/34

    26

    (a) Step 0, t = 0.00ms (b) Step 150, t = 1.20ms

    (c) Step 350, t = 2.80ms (d) Step 450, t = 3.60ms

    Figure 10. Simulation of supersonic flow past a flexible structure

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    28/34

    27

    (a) Step 650, t = 5.20ms (b) Step 850, t = 6.80ms

    (c) Step 1050, t = 8.40ms (d) Step 3050, t = 24.40ms

    Figure 11. Simulation of supersonic flow past a flexible structure (continued)

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    29/34

    28

    the windward side stagnates and the pressure increases again, as shown in Figure 10 (c). At

    t = 3.60ms, Figure 10 (d), the expansion wave propagating upstream reaches the shock front,

    affecting its shape and causing it to move downstream, thus following with some delay the

    initial forward motion of the rod. In the same Figure, it can be observed that, at this point,

    the elastic strain energy stored in the structure starts causing it to recoil. As the elastic energy

    is released, Figure 11 (a), the structure pushes the upstream flow generating a compression

    wave. This emphasizes the ability of the method to describe aspects of the flow caused by the

    dynamic deformation of the structure. Figure 11 (b) clearly shows the compression wave shed

    by the structure propagating upstream towards the shock front. At t = 8.40ms, Figure 11

    (c), the compression wave reaches the shock front, affecting its shape and causing it to move

    backwards, following again with some delay the motion of the structure.

    This process in which the structure first inflates storing elastic energy and then recoils

    restoring the energy to the flow continues ostensibly unchanged, as no physical dissipation

    mechanisms are taken into account in the model. In order to reach a steady-state inflated

    configuration, a small numerical dissipation is added to the structure by a convenient choice

    of the parameters of Newmarks algorithm, equation (11), as = 0.5 and = 1. Figure 11 (d)

    shows the steady-state configuration reached at time t = 24.40ms after the structure was first

    released.

    This example illustrates the robustness of the coupling algorithm in describing complex

    fluid-solid interactions in which the structure undergoes large nonlinear elastic deformations

    which, in turn, affect the flow in a non-trivial manner. The ability of the model to describe

    these interactions across a very thin structure without cross pollution of the flow across the

    interface is particularly noteworthy.

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    30/34

    29

    Summary and conclusions

    We have proposed a computational strategy for the coupling of high-speed flows interacting

    with the large, dynamic deformations of very thin open structures. Among the necessary

    components of the overall computational framework, a formulation is presented for the large

    dynamic deformations of thin rod structures including the bending and membrane response.

    The coupling algorithm constitutes an extension of the ghost fluid method without the

    restrictions of thick solid structures and closed boundaries in which a well-defined exterior

    to the fluid domain exists. The new algorithm was verified against the analytical solution of

    the supersonic flow past a flat rigid plate at different angles of attack. The numerical solution

    is shown to converge to the analytical solution on both the shocked and rarefied regions on the

    windward and leeward side of the plate without pollution of the solution across the infinitely

    thin boundary. A convergence analysis of the lift load on the structure confirms the theoretical

    first order accuracy of the coupling approach. As an example of a coupled application, a

    simulation of the transient supersonic flow normal to a highly-flexible structure is presented.

    The simulation shows that a complex pattern of highly unsteady coupled interactions are set

    in motion between the flow and the structure, leading to the large oscillations of the structure

    until a steady-state is reached in its final inflated configuration.

    ACKNOWLEDGEMENTS

    The support of the U.S. Department of Energy through the ASC Center for the Simulation of the

    Dynamic Response of Materials (DOE W-7405-ENG-48, B523297) is gratefully acknowledged.

    REFERENCES

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    31/34

    30

    1. G. Guglieri and F. Quagliotti. Validation of a simulation model for a planetary entry capsule. Journal

    of Aircraft, 40(1):127136, 2003. 1, 3

    2. Y.H. Cao and H. Xu. Parachute flying physical model and inflation simulation analysis. Aircraft

    Engineering and Aerospace Technology, 76(2):215220, 2004. 1, 3

    3. M.B. Quadrelli, J.M. Cameron, and V. Kerzhanovich. Multibody dynamics of parachute and balloon flight

    systems for planetary exploration. Journal of Guidance Control and Dynamics, 27(4):564571, 2004. 1,

    3

    4. R. Radovitzky and M. Ortiz. Lagrangian finite element analysis of newtonian fluid flows. International

    Journal For Numerical Methods In Engineering, 43(4):607617, 1998. 2

    5. R. Samtaney and N. J. Zabusky. Circulation deposition on shock-accelerated planar and curved density-

    stratified interfaces: models and scaling laws. Journal of Fluid Mechanics, 269:4578, 1994. 2, 8

    6. R. Samtaney and D. I. Meiron. Hypervelocity Richtmyer-Meshkov instability. Physics of Fluids,

    9(6):17831803, 1997. 2, 8

    7. C.S Peskin and D.M. McQueen. A threedimensional computational method for blood flow in the heart: (i)

    immersed elastic fibers in a viscous incompressible fluid. Journal of Computational Physics, 81:372405,

    1989. 2

    8. Charles S. Peskin. The immersed boundary method. Acta Numerica, pages 139, 2002. 2

    9. X. Wang and W. K. Liu. Extended immersed boundary method using fem and rkpm. Computer Methods

    in Applied Mechanics and Engineering, 193:13051321, 2004. 2

    10. J. J. Quirk. An alternative to unstructured grids for computing gas dynamic flows around arbitrarily

    complex two-dimensional bodies. Computers & Fluids, 23:125142, 1994. 2

    11. J. J. Quirk. A parallel adaptive grid algorithm for computational shock hydrodynamics. Applied Numerical

    Mathematics, 20:427453, 1996. 2

    12. P. Colella. Volume-of-fluid methods for partial differential equations. In E. F. Toro, editor, Godunov

    Methods: Theory and Applications. Kluwer Academic/Plenum Publishers, New York, 2001. 2, 3

    13. D. Meiron, R. Radovitzky, and R. Samtaney. The virtual test facility: An environment for simulating

    the nonlinear dynamic response of solids under shock and detonation wave loading. In Proceedings of

    the Sixth U.S. National Congress on Computational Mechanics, Dearborn, MI, 2001. U.S. Association for

    Computational Mechanics. 3, 14

    14. J. Cummings, M. Aivazis, R. Samtaney, R. Radovitzky, S. Mauch, and D. Meiron. A virtual test facility

    for the simulation of dynamic response in materials. Journal Of Supercomputing, 23(1):3950, 2002. 3,

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    32/34

    31

    4, 16

    15. R.P. Fedkiw. Coupling an Eulerian fluid calculation to a Lagrangian solid calculation with the ghost fluid

    method. Journal of Computational Physics, 175:200224, 2002. 3

    16. F. Cirak and R. Radovitzky. A new fluid-shell coupling algorithm based on level sets. In Proceedings of

    the 44th AIAA/ASME/ASCE/AHS Structures, Structural Dynamics, and Materials Conference, Norfolk,

    VA, April 7-10 2003. American Institute of Aeronautics and Astronautics. 3, 11

    17. M. Arienti, P. Hung, and J. Morano, E.and Shepherd. A level set approach to Eulerian-Lagrangian

    coupling. Journal of Computational Physics, 185:213251, 2003. 3, 22

    18. R.P. Fedkiw, T. Aslam, B. Merriman, and S. Osher. A non-oscillatory Eulerian approach to interfaces in

    multimaterial flows (the ghost fluid method). Journal of Computational Physics, 152:457492, 1999. 3

    19. M. Sussman. A second order coupled level set and volume-of-fluid method for computing growth and

    collapse of vapor bubbles. Journal of Computational Physics, 187:110136, May 2003. 3, 24

    20. M. Sussman. A parallelized, adaptive algorithm for multiphase flows in general geometries. Computers

    and Staggered, 83:435444, 2005. 3, 24

    21. F. Cirak and R. Radovitzky. A general algorithm for coupling Lagrangian-shell withe Eulerian-fluid

    formulations. In Proceedings of the IUTAM Symposim on Integrated Modeling of Fully Coupled Fluid-

    Structure Interactions Using Analysis, Computations and Experiments, New Brunswick, NJ, June 1-6

    2003. International Union of Theoretical and Applied Mechanics. 3, 14

    22. F. Cirak and R. Radovitzky. A Lagrangian-Eulerian shell-fluid coupling algorithm based on level sets.

    Computers and Structures, 83:491498, 2005. 3, 4

    23. Gere and Timoshenko. Mechanics of Materials. Van Nostrand Reinhold Co., New York, 1972. 4

    24. D. H. Hodges. Finite rotation and nonlinear beam kinematics. Vertica, 11:297307, 1987. 4

    25. D. H. Hodges. Nonlinear beam kinematics for small strains and finite rotations. Vertica, 11:573589,

    1987. 4

    26. M. R. M. Crespo da Silva. Equations for nonlinear analysis of 3d motions of beams. Applied Mechanics

    Reviews, 44:S51S59, 1991. 4

    27. J. E. Marsden and T. J. R. Hughes. Mathematical foundations of elasticity. Prentice-Hall, Englewood

    Cliffs, N.J., 1983. 6

    28. J. N. Reddy. Energy Principles and Variational Methods in Applied Mechanics, Second Edition. John

    Wiley and Sons, Hoboken, NJ, 2002. 7

    29. D. Tam. A two-dimensional fluid-structure coupling algorithm for the interaction of high-speed flows with

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    33/34

    32

    open shells. Masters thesis, Massachusetts Institute of Technology, Cambridge, MA, June 2004. 7, 8

    30. Thomas J. R. Hughes. The Finite Element Method. Dover, 2000. 8

    31. D. I. Pullin. Direct simulation methods for compressible ideal gas flow. Journal of Computational Physics,

    34:231244, 1980. 8, 9

    32. C. Hirsch. Numerical Computation of Internal and External Flows, volume 1, Fundamentals of Numerical

    Discretization. John Wiley and Sons, 1988. 9, 10

    33. Culbert B. Laney. Computational Gasdynamics. Cambridge University Press, 1998. 9, 10

    34. S.K. Godunov. Finite-difference method for the numerical computation of discontinuous solutions of the

    equations of fluid dynamics. Mat. Sbornik (Mosk.), 47:271306, 1959. 9

    35. P. Glaister. An approximate linerarised Riemann solver for Euler equations for real gases. Journal of

    Computational Physics, 74:382408, 1988. 9

    36. B. Van Leer. Towards the ultimate conservative difference scheme iv: A new approach to numerical

    converction. Journal of Computational Physics, 23:276299, 1977. 9

    37. Randall J. LevVeque. Finite Volume Methods for Hyperbolic Problems. Cambridge University Press,

    2002. 9

    38. C. Hirsch. Numerical Computation of Internal and External Flows, volume 2, Computational Methods

    for Inviscid and Viscous Flows. John Wiley and Sons, 1990. 9, 10

    39. ASCI Alliance Center for the Simulation of Dynamic Response of Materials, FY00 Annual Report. URL:

    http:// www.cacr.caltech.edu/ ASAP/onlineresources/publications/, 2000. 10

    40. C. Farhat, P. Geuzaine, and C. Grandmont. The discrete geometric conservation law and the nonlinear

    stability of ale schemes for the solution of flow problems on moving grids. Journal of Computational

    Physics, 174:669694, 2001. 10

    41. M. Cruchaga, D. Celentano, and T. Tezduyar. A moving Lagrangian interface technique for flow

    computations over fixed meshes. Computer Methods in Applied Mechanics and Engineering, 191:525

    543, 2001. 10

    42. T. E. Tezduyar. Computation of moving boundaries and interfaces and stabilization parameters.

    International Journal for Numerical Methods in Fluids, 43:555575, 2003. 10

    43. K. Stein, T. Tezduyar, and R. Benney. Mesh moving techniques for fluid-structure interactions with large

    displacements. Journal of Applied Mechanics-Transactions of the ASME, 70:5863, 2003. 10

    44. K. C. Park, C. A. Felippa, and J. A. Deruntz. Stabilization of staggered solution procedures for fluid-

    structure interaction analysis. In T. Belytschko and T. L. Geers, editors, Computational Methods for

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00

    Prepared using nmeauth.cls

  • 8/2/2019 Algo for Modelling (Pre)2005

    34/34

    33

    Fluid-Structure Interaction Problems, pages 94124, New York, 1997. ASME. 11

    45. Q. Zhang and T. Hisada. Studies of the strong coupling and weak coupling methods in fsi analysis.

    International Journal for Numerical Methods in Engineering, 60:20132029, 2004. 11, 16

    46. Sean Mauch. Efficient Algorithms for Solving Static Hamilton-Jacobi Equations. PhD thesis, California

    Institute of Technology, Pasadena, CA, 2003. 13

    47. S. Mauch. A fast algorithm for computing the closest point and distance transform. Preprint,

    http://www.acm.caltech.edu/seanm/software/cpt/cpt.html, 2001. 14

    48. A. Bonnet and J. Luneau. Theorie de la Dynamique des Fluides. Cepadues Editions, Toulouse, France,

    1989. 18

    49. P. Thompson. Compressible Fluid Dynamics. McGraw-Hill, New York, 1972. 18, 19

    Copyright c 2004 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2004; 00:00