Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

GEOPHYSICS, VOL. 82, NO. 4 (JULY-AUGUST 2017); P. T183–T196, 9 FIGS., 3 TABLES. 10.1190/GEO2016-0691.1

Solving the tensorial 3D acoustic wave equation: A mimetic finite-difference time-domain approach

Jeffrey Shragge1 and Benjamin Tapley2

stencils to be applied throughout the model domain, including the boundary region where implementing centered FD stencils can be problematic. The FSG approach combines wavefield information propagated on four complementary subgrids to ensure the existence of all wavefield gradients required for computing the tensorial Laplacian operator, and thereby avoids interpolation approximations. The RBCs are implemented with a flux-preserving mimetic boundary operator that forestalls introduction of nonphysical energy into the grid by enforcing underlying flux-conservation laws. After validating the 3D MFDTD scheme on a sheared Cartesian mesh, we generate 3D wavefield simulation examples for internal boundary (IB) and topographic coordinate systems. The numerical examples demonstrate that the MFDTD scheme is capable of providing accurate and low-dispersion impulse responses for scenarios involving distorted IB meshes conforming to water-bottom surfaces and topographic coordinate systems exhibiting 2.5 km of topographic relief and including steep (65°) slope angles.

ABSTRACT Generating accurate numerical solutions of the acoustic wave equation (AWE) is a key computational kernel for many seismic imaging and inversion problems. Although finite-difference timedomain (FDTD) approaches for generating full-wavefield solutions are well-developed for Cartesian computational domains, several challenges remain when applying FDTD approaches to scenarios arguably best described by more generalized geometry. In particular, how best to generate accurate and stable FDTD solutions for scenarios involving grids conforming to complex topography or internal surfaces. We address these issues by developing a mimetic FDTD (MFDTD) approach that combines four key components: a tensorial 3D AWE, mimetic finite-difference (MFD) operators, fully staggered grids (FSGs), and MFD Robin boundary conditions (RBC). The tensorial formulation of the 3D AWE permits wave propagation to be specified on (semi-) analytically defined coordinate meshes designed to conform to complex domain boundaries. MFD operators allow for higher order FD

boundaries for reasons of numerical expediency or accuracy; for example, facilitating wavefield injection of ocean-bottom node seismic data on an irregular water bottom, or improved handling of rugose internal model surfaces (i.e., water bottom or top salt body) by mitigating the “staircasing” effects commonly observed in discretized Cartesian velocity models. Introducing non-Cartesian geometry in 3D acoustic-wave propagation problems, though, necessarily complicates the simulation of numerical solutions relative to Cartesian approaches, and practitioners may choose from a variety of computational strategies. Integral-based methods are one class of numerical techniques that can be used to solve the AWE on voxelized elements designed to

INTRODUCTION Many seismic modeling, imaging, and inversion problems involve generating full-wavefield solutions of the 3D acoustic wave equation (AWE) on computational meshes arguably best described by nonCartesian geometry. Example scenarios include performing reverse time migration (RTM) or full-waveform inversion (FWI) from irregular topographic surfaces, handling severe irregular crooked-line geometry in 2D land seismic data sets, modeling waveforms from deviated boreholes, and incorporating earth curvature in long-offset crustal seismic surveys. In other situations, it may be advantageous to use computational meshes conforming to internal discontinuous

Manuscript received by the Editor 19 December 2016; revised manuscript received 14 March 2017; published online 13 June 2017. 1 The University of Western Australia, Centre for Energy Geoscience, School of Earth Sciences and School of Physics, Crawley, Western Australia, Australia. E-mail: [email protected]. 2 The University of Western Australia, School of Physics, Crawley, Western Australia, Australia. E-mail: [email protected]. © 2017 Society of Exploration Geophysicists. All rights reserved. T183

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

T184

Shragge and Tapley

conform to the domain boundaries and/or judiciously chosen interior model surfaces. Examples of integral solutions include the finite-element (Marfurt, 1984), spectral-element (Komatitsch and Vilotte, 1998), and discontinuous-Galerkin (Cockburn et al., 2000) methods. Although integral approaches can be developed that generate highly accurate numerical results, they are characterized by increased computational complexity when compared with finite-difference (FD) solutions of the 3D AWE. However, generating stable, efficient, and sufficiently accurate FD solvers for generalized geometries is not a straightforward task. With the aim of developing an efficient 3D propagation kernel for the types of industrial acoustic RTM and FWI applications described above, the work presented herein focuses on generating cost-effective finite-difference timedomain (FDTD) solutions of the tensorial 3D AWE based on “semianalytic” coordinate transformations. Many different authors have examined the use of coordinate transformation approaches for generating solutions to 2D/3D wave-propagation problems. Although most efforts have focused on (visco-) elastic wave-equation modeling (Carcione, 1994; Komatitsch et al., 1996; Ohminato and Chouet, 1997; Hestholm, 1999; Hestholm and Ruud, 2002; Ely et al., 2008, 2009; Appelö and Petersson, 2009; Zhang et al., 2012; de la Puente et al., 2014), several studies have examined this solution strategy for solving the 3D AWE in nonCartesian coordinates (Nichols, 1994; Cheng and Blanch, 2008; Shragge, 2014a, 2014b). Commonly, these studies develop FDTD solutions for a particular coordinate geometry of interest (e.g., cylindrical borehole coordinates); however, when viewed collectively, these examples are members of a more general tensorial formulation of the 3D AWE. Shragge (2014b) extends earlier work on one-way Riemannian wavefield extrapolation theory (Sava and Fomel, 2005; Shragge, 2008) by formulating generalized two-way solutions to the tensorial 3D AWE using a decomposition of the tensorial 3D Laplacian operator into generalized first- and (mixed) second-order derivatives. The developed numerical approach discretizes the resulting partial differential expressions using conventional Taylor-series expansions on a nonstaggered grid (NSG), and it applies a low-order “ghost-point” approximation for handling boundary regions. Shragge (2014b) also provides 3D numerical examples for semianalytically defined meshes, including internal boundary (IB) and topographic coordinate systems. Although generating wavefield solutions to the tensorial 3D AWE on NSGs is a memory-efficient and computationally efficient strategy for many non-Cartesian geometries, this approach has two main drawbacks when dealing with scenarios involving more highly irregular geometry. First, because FD stencils used within the boundary region of computational grids are of lower order than those applied within the mesh interior (i.e., OðΔx2 Þ versus OðΔx4 Þ or higher), solutions computed in the boundary region will degrade the overall accuracy of the numerical solution over the duration of the simulation and contribute to the rise of long-term numerical instabilities. These accuracy and stability issues become increasingly acute for non-Cartesian meshes exhibiting grid distortion of increasing severity. Accordingly, these scenarios evidently require more resilient numerical and computational approaches. A second drawback of NSGs is that they preclude taking advantage of the accuracy improvements afforded by performing wavefield simulation on staggered grids (Madariaga, 1976; Virieux, 1986). Although generating 3D AWE solutions on various forms of standard staggered grids (SSGs) is more memory intensive and less computa-

tionally efficient than on NSGs, the improved effective sampling leads to more accurate wavefield simulations characterized by reduced numerical dispersion for a given fixed grid spacing. However, a corresponding drawback of using SSGs to solve the tensorial 3D AWE formulation is that numerical approximations of mixed secondorder partial derivatives used in tensorial 3D Laplacian operator do not fall on the same SSG grid (de la Puente et al., 2014) and must be estimated from neighboring grid points using high-order (e.g., a bi-cubic) interpolation operators (Hestholm et al., 2006). Following this approach reduces the overall accuracy of SSG wavefield solutions to that of the interpolation operators, and it can generate numerical instabilities earlier than deemed acceptable — though generally at longer simulation times than for comparable solutions computed on NSGs. One class of computational strategies able to address the accuracy and stability issues described above involves the use of fully staggered grids (FSGs) (Lebedev, 1964). An FSG solution of a 3D wave-propagation problem uses a system of four coupled SSGs of unique grid staggerings that collectively form a complementary set of wavefield solutions. By complementary, we mean that all of the mixed second-order partial derivative approximations required to simulate propagation on one subgrid (SG) may be found at a colocated point in one of the other three. FSGs increasingly are being used to compute highly accurate solutions to (Cartesian) 3D anisotropic elastodynamics problems (Lisitsa and Vishnevskiy, 2010; Bernth and Chapman, 2011). However, adopting an FSG approach to generate solutions of the tensorial 3D AWE — while achieving significant gains in numerical accuracy of solutions — does not in itself forestall the rise of instabilities caused by using low-order FD approximations within the computational domain boundary region. The development of mimetic finite-difference (MFD) operators as a method of generating high-order numerical solutions to partial differential equations (PDEs) throughout a computational domain arguably represents an important advance for dealing with accuracy and instability issues associated with finite meshes and implementing boundary conditions. The main benefit of adopting an MFD strategy is that one may design MFD gradient, divergence, and curl operators that “satisfy fundamental properties of the PDE in question including conservation laws, symmetry and positivity of solutions, self-adjointness of differential operators, and exact mathematical identities of vector and tensor calculus” (Lipnikov et al., 2014). Recent work demonstrates that introducing MFD gradient, divergence, and composite Laplacian operators into an FSG + MFD solution of hyperbolic PDEs can achieve high-order accuracy throughout the entire computational domain — including within the problematic boundary region (see, e.g., Castillo and Miranda [2013]; readers interested in the history of MFD theory development are referred to the review paper of Lipnikov et al., 2014). Studies incorporating FSG + MFD solutions of wave equations include: modeling the dynamics of earthquake rupture propagation (Rojas et al., 2008), solving the 2D AWE on unstructured polygonal meshes (Lipnikov and Huang, 2008), solving the 3D elastodynamic wave equation in the presence of free-surface topography by incorporating coordinate transformations (de la Puente et al., 2014), and solving the Cartesian 3D AWE using a tensor product approach (Solano-Feo et al., 2016). This paper develops an FSG + MFD strategy for generating mimetic FDTD (MFDTD) solutions of the tensorial 3D AWE applicable for generalized geometry. This main goal is to mitigate the shortcomings of computing solutions to the tensorial 3D AWE on NSGs

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Solving tensorial 3D AWE: MFDTD approach and SSGs. The improved accuracy and stability of the MFDTD solutions extend the tensorial approach to more complex computational geometries (i.e., of increasing irregularity and roughness). We begin with a discussion on the theory of 3D tensorial AWE, including some aspects of differential geometry governing acoustic wave propagation in generalized coordinate systems. Next, we describe the MFD operators and Robin boundary conditions (RBCs) used in the numerical solutions, and present the 3D grid staggerings used in this study. After developing and validating the specific OðΔt2 ; Δx4 Þ MFD operators, we present 3D wave-propagation examples generated in sheared Cartesian, IB, and topographic coordinates. The paper concludes with a discussion on computational merits and drawbacks of the developed MFDTD approach.

TENSORIAL 3D AWE The tensorial form of the (constant-density) AWE in a 3D generalized coordinate system defined by variables ξ ¼ ½ξ1 ; ξ2 ; ξ3  is given by



 1 ∂2 2 ∇ξ − 2 2 U ξ ¼ F ξ ; cξ ∂t

(1)

where ∇2ξ is a tensorial Laplacian operator, cξ is the P-wave velocity field, U ξ is a scalar acoustic wavefield, and F ξ is the source distribution. Here and throughout, we use a notation where subscripts ξ designate fields and operators defined in a generalized ξ-coordinate system that represents a regularly sampled computational domain throughout which one computes a 3D AWE solution. A subscript x indicates an irregularly sampled physical domain over which one desires a 3D AWE solution. We assume that the mapping relationship between the ξ- and x-coordinate systems is known, unique (i.e., one-to-one), and expressible through a set of analytic, semianalytic, or fully numerical mapping relationships:

xi ¼ xi ðξÞ

ξi ¼ ξi ðxÞ for i ¼ 1; 2; 3:

and

2

g11 ½gij  ¼ 4 g12 g13

g12 g22 g23

3 g13 g23 5; g33

(3)

whose elements gij provide a link between the ξ- and x-coordinate systems through

∂xk ∂xk gij ¼ : ∂ξi ∂ξj

The tensorial 3D AWE requires a contravariant representation of the metric tensor (i.e., gij ), defined as the inverse of the covariant metric tensor ½gij 

2

3 g22 g33 −g223 g13 g23 −g12 g33 g12 g33 −g13 g22 1 ½gij  ¼ 4 g13 g23 −g12 g33 g11 g33 −g213 g12 g13 −g11 g23 5; jgj g12 g33 −g13 g22 g12 g13 −g11 g23 g11 g22 −g212 (5) where the metric tensor discriminant jgj is given by

jgj ¼ jg11 g22 g33 − g212 g33 − g223 g11 − g213 g22 þ 2g12 g13 g23 j; (6) and j · j is the absolute value operator. Computing the partial derivatives of ∇2ξ in equation 1 involves introducing the contravariant metric tensor elements from equations 5 and 6 into a tensorial expression for the Laplacian operator (Guggenheimer, 1977)

∇2ξ

  1 ∂ pffiffiffiffiffiffi ij ∂ ¼ pffiffiffiffiffiffi jgjg ; ∂ξj jgj ∂ξi

(4)

Unless stated otherwise, we use summation notation over repeated indices and use a convention by which matrix subscripts and superscripts (e.g., gij and gij ) indicate covariant and contravariant tensors, respectively (Synge and Schild, 1978).

(7)

where partial derivatives ∂∕∂ξi are evaluated with respect to the generalized ξ-coordinate variables. Inserting the Laplacian expression of equation 7 into equation 1 generates the tensorial 3D AWE appropriate for any ξ-coordinate system satisfying the aforementioned one-to-one mapping relationship. However, we emphasize that it is the geometric characteristics of any specific choice of coordinate system that determines the overall stability, computationally tractability, and/or resulting numerical efficiency of MFDTD simulations. For the purposes of discussion below, we note that the tensorial Laplacian operator when applied to scalar field f may be decomposed into tensorial gradient

(2)

The second-order partial differentials of ∇2ξ , though specified in a regular ξ-coordinate system, are affected by spatially varying geometry in physical x-coordinates. Correctly formulating the tensorial Laplacian operator requires introducing transformations from differential geometry in the form of a (symmetric rank-two) metric tensor ½gij 

T185

∇ξ f ¼ gij

∂f ; ∂ξj

(8)

and tensorial divergence

  1 ∂ pffiffiffiffiffiffi i ∇ξ · v ¼ pffiffiffiffiffiffi jgjv ; jgj ∂ξi

(9)

operators, where v is a generic vector field. Finally, all quantities hereafter are assumed to be in a ξ-coordinate system and for brevity, all ξ subscripts will be omitted unless required for clarity.

MFDs ON FULLY STAGGERED GRIDS This section describes an MFDTD approach for solving the tensorial 3D AWE on an FSG. We first discuss the mimetic tensorial gradient, divergence, and boundary operators used in the MFD numerical scheme, and then we describe how the boundary operator is used to enforce RBCs at the computational domain boundaries. We conclude with a discussion on the staggerings of the four complementary SG chosen to comprise the FSG, as well as some implementation considerations.

Shragge and Tapley

T186

umþ1 ¼ 2um − um−1 þ Δt2 C2 ðLum þ Fm Þ;

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Mimetic finite differences Over the past few decades, MFD operators increasingly have been used to compute numerical solutions for a variety of different PDEs. One of the main benefits of following an MFD strategy is to exploit its ability to satisfy fundamental properties of the physical system including conservation laws, symmetry, and exact mathematical identities of vector and tensorial calculus (Lipnikov et al., 2014). The MFD operators can be constructed with a high order of numerical accuracy not just throughout a 3D computational domain M, but also within the boundary region MBND , and right up to the domain boundary ∂M. Although satisfying these properties is generally important for generating stable and accurate numerical solutions for any PDE, it is especially critical when applying FDbased methods on non-Cartesian geometries exhibiting significant topological complexity due to the presence of spatially irregular “geometric fluxes” (i.e., numerical approximations of gradient operators gij ð∂U∕∂ξj Þ) within MBND . Herein, we solve the tensorial 3D AWE by implementing an MFD approximation of the Laplacian operator in equation 7 based on composite MFD gradient and divergence operators, Gi and Di , where the subscript indicates the direction in which the operator acts. These operators satisfy the aforementioned fundamental properties of the MFD approach operators throughout M due to four key factors: (1) using an FSG approach comprised of four complementary SG defined on judiciously staggered nodes, (2) incorporating additional points on domain boundary ∂M to facilitate higher order FD approximations within MBND , (3) introducing higher order asymmetric FD stencils within MBND that ideally are of the same accuracy order as those used within the domain interior, and (4) specifying a boundary operator Bn to enforce flux-conserving boundary conditions. As one moves away from MBND into the computational domain interior, though, the MFD Gi and Di operators rapidly converge toward the familiar Taylor-series FD stencils of the equivalent order of accuracy. Without yet specifying the order of spatial discretization of MFD operators, one can develop an MFDTD scheme for the 3D tensorial AWE in equation 1 by: (1) replacing the continuous wavefield with its discrete equivalent, U ↔ um , where integer superscript m ¼ ½0; N t  is the temporal index, (2) substituting a OðΔt2 Þ Taylor-series approximation for the second-order temporal partial derivative, and (3) replacing the analytic tensorial Laplacian operator with the MFD equivalent L formed by the composition of mimetic gradient

∇f ¼ gij

∂f ↔ Γij Gj f; ∂ξj

(10)

and mimetic divergence

  1 ∂ pffiffiffiffiffiffi i ∇ · v ¼ pffiffiffiffiffiffi i jgjv ↔ ψ −1 Di ψvi ; jgj ∂ξ

(11)

operators according to

∇2 f ↔ ψ −1 Di ψΓij Gj ≡ L;

(12)

where p ffiffiffiffiffi ψ is a matrix version of the metric tensor determinant (i.e., jgj → ψ) and Γij represents the discretized metric tensor that introduces coupling between the i ≠ j directions for nonorthogonal metrics. This leads to the following MFDTD update formula:

(13)

where C is a (diagonal) matrix representation of velocity c, Δt is the time-sampling interval, and Fm is a matrix representing the analytic source function f. Appendix A provides a matrix representation of 1D Gi and Di derivative operators of accuracy OðΔx4 Þ. Applying these operators in the second or third dimension requires sequentially acting on a wavefield array extracted from memory with the required 1D or 2D stride. The MFDTD scheme in equation 13 is appropriate for computing solutions to the tensorial AWE in unbounded computational domains; however, practical implementation requires introducing boundary conditions that account for the finiteness of the computational mesh.

MFD Robin boundary conditions One important application of the MFD gradient and divergence operators described above is that they can be modified so as to implement flux-conserving boundary conditions enforced by an MFD boundary operator (Castillo and Miranda, 2013). Herein, we provide an overview of this approach; for a more thorough mathematical treatment, the reader is referred to Castillo and Miranda (2013). The approach for generating a flux-preserving MFD boundary operator is based on evoking the Green-Gauss-Stokes theorem (GGST) in boundary region MBND . Recall that for arbitrary scalar and vector fields in a continuum, f and v, the GGST is expressed as

ZZ ZZZ ⬭ðfv · nÞdA ¼ ðv · ∇f þ f∇ · vÞdV; ∂M

M

(14)

where n is the normal to area element dA of boundary surface ∂M and dV is a volume element. This relationship states that the net flux entering a closed surface (left side) is equal to the net sources of flux enclosed by the surface (right side). Specifically for the problem at hand, f and v could represent an acoustic wavefield and its spatial derivative, respectively. For scenarios in which (1) ∂M is a free-surface boundary completely encompassing an object described by the domain M (e.g., cube or cylinder), (2) there is perfect internal reflection of acoustic waves encountering ∂M, and (3) there are no acoustic sources external to M; we would expect both sides of equation 14 to be identically zero because there should be zero net flux in or out of M throughout the duration of simulation. Note that the GGST is a statement of the conservation of energy and the conservation of flux, and it is generally applicable regardless of the physical system under consideration. A key challenge in developing an accurate MFDTD scheme is implementing a discrete flux-preserving boundary condition. One approach for achieving this is to use a numerical strategy that encodes flux conservation into a boundary operator that naturally enforces the GGST within MBND . The theoretical work on MFD boundary operators summarized in the Castillo and Miranda (2013) monograph demonstrates how this can be achieved using commonalities between inner-product spaces defined on continuua and discrete vector spaces. We follow their approach by stating the discrete equivalence of the GGST:

hBξn u; viI ¼ hGn u; viP þ hu; Dn viQ ;

(15)

where h·; ·iA represents a standard inner product of the form hu; viA ¼ uT Av for arbitrary vectors u and v and a generic weight matrix A; Bξn

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Solving tensorial 3D AWE: MFDTD approach is the boundary operator applied in a direction normal n to the computational domain boundary ∂M; and I, P, and Q are the identity and two weight matrices, respectively (note that the order of terms in equation 15 is written to correspond with those in equation 14). The main diagonal entries of matrices P and Q represent midpoint integration weights designed to ensure that MFD operators Gi and Di , when used in equation 15, satisfy the GGST throughout M. Weight matrix operators may be developed using a coupled Taylor-series plus Vandermonde matrix approach described in Castillo and Miranda (2013). The particular set of matrix weight coefficients will depend on the particular choice of grid staggerings and on the order of the implemented FD scheme. Appendix A provides a matrix representation of the 1D P and Q operators of accuracy OðΔx4 Þ. Using the matrix operator notation of Solano-Feo et al. (2016), we write boundary operator Bξn as

Bξn

¼ GTn P þ QDn ;

(16)

where the superscript T indicates matrix transpose and the superscript ξ emphasizes that this operator is applicable on the numerical solution domain. Appendix A also provides an example of a 1D OðΔx4 Þ boundary operator Bξn. Note that the structure of this Bξn operator contains only nonzero values in the upper and lowermost 6 × 6 matrix entries and, when applied to a wavefield vector, would otherwise not affect the computational grid (generally speaking, an OðΔxN Þ Bξn operator will affect a width of 2ðN − 1Þ grid points). Analogous to the role of the GGST in a continuum, the discrete boundary operator Bξn is specified independently of the physical system under consideration, and it is solely determined by the user-specified order of accuracy, the structure of the Gn and Dn coefficients, and the corresponding P and Q weights. Castillo and Miranda (2013) further demonstrate how a discrete version of the GGST can be enforced by using Bn in an MFD-RBC scheme. An analytic expression for the RBC, which represents a weighted linear superposition of Dirichlet and Neumann boundary conditions, is given by

½βn · ∇ þ αUj∂M ¼ h;

T187

where the summation over n is in the three principal coordinate directions, A is a boundary selection operator having only nonzero entries of unitary values at locations corresponding to the selected domain boundaries, and 0 is the null matrix. Introducing the MFDRBC directly into the MFDTD scheme in equation 13 yields

umþ1 ¼ 2um − um−1 þ Δt2 C2 ½ðL − RÞum þ Fm ;

(20)

which is the desired MFDTD approach that enforces an MFD-RBC scheme over a finite-computational domain. We use this numerical scheme to generate solutions of the tensorial 3D AWE in the examples below.

Grid staggering The aforementioned MFD operators have been introduced without reference to FSGs; however, this represents an important point for developing accurate solutions of the tensorial 3D AWE. We begin by outlining 1D and 2D grid staggerings that are useful for developing intuition as to how a 3D FSG scheme can be implemented using four complementary SGs. We follow the notation of Castillo and Miranda (2013), which differentiates between fields defined on half-integer (f) or full-integer (v) grids. 1D staggerings Figure 1a presents a 1D example of staggered f and v grids for an n1 ¼ 8 point mesh, in which the grids are of dimension f ∈ Rn1 þ2 and v ∈ Rn1 þ1 and share points on the domain boundary. To compute an MFDTD solution of the 1D tensorial AWE on the grids shown in Figure 1a, one assumes wavefield um ∈ Rn1 þ2 is defined on f (blue points). The first step applies G1 um ≡ v1 to approximate

(17)

where α and β are the scalar weights determined by the particular physical problem; n is the unit normal to boundary ∂M; and h ¼ hðtÞ represents time-dependent behavior on ∂M. For the applications described herein, we assume time-independent behavior on ∂M, and thus enforce h ¼ 0. Solano-Feo et al. (2016) demonstrate that boundary operator Bξn can be used to specify an MFD-RBC matrix operator, which herein we term R. Numerical implementations of equation 17 on a Cartesian mesh have outward-oriented normal vectors n on each of the three principal axes and thus numerically approximate n · ∇ ↔ δni ð∂∕∂xi Þ in the i ¼ 1; 2; 3 directions. Implementation in generalized coordinates introduces the differential expression for the gradient operator in equation 8 (i.e., n · ∇ ↔ gnj ð∂∕∂ξj Þ) that projects the wavefield-gradient operator into ξ coordinates prior to applying the MFD-RBC scheme. In our examples, we use α ¼ β ¼ 1, which leads to the following matrix expressions for the MFD-RBC scheme:

Rum ≡ ðBξn Gn þ AÞum ¼ 0 Cartesian;

(18)

Rum ≡ ðBξn Γnj Gj þ AÞum ¼ 0 Generalized;

(19)

Figure 1. Examples of MFD staggered grids. (a) 1D grid staggering showing the difference between the f half-integer (blue) and v integer (pink) grid locations. Note that the two grids share boundary points (i.e., f 0 ¼ v0 and f 8 ¼ v8 ). (b) The 2D staggered grid example on f (pink) and v (blue) two SGs that share corners but not side points.

Shragge and Tapley

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

T188

∇U on v (pink points). Note that the action of G1 on um reduces the resulting vector dimensions according to Rn1 þ2 → Rn1 þ1 . The second step applies D1 v1 to approximate ∇2 U, which returns the Laplacian L of um on f (blue points). Thus, the dimensions are altered according to Rn1 þ1 → Rn1 þ2 . Note that although Lum is not explicitly defined on the exterior points of f, one may insert zero values at these locations by augmenting the upper- and lowermost rows of divergence operator D1 with zero values (see the example in Appendix A). To illustrate the cyclicity in grid dimensions when applying MFD G1 and D1 operators, we use a symbolic notation where ½f!½v!½f indicates successive applications of partialD1 G1 derivative operators. We note that specifying 1D wavefield values of um to fall on f represents one particular choice; alternatively, one could specify um to fall on v, and thus ½v!½f!½v using modified D1 G1 grid dimensions. 2D staggerings Figure 1b presents a 2D staggered grid example for n1 ¼ n2 ¼ 8 using two complementary SGs defined on ½f; f ∈ Rðn1 þ2Þ×ðn2 þ2Þ and ½v; v ∈ Rðn1 þ1Þ×ðn2 þ1Þ . Computing a 2D AWE solution on this mesh largely follows the 1D procedure, in which the contributions of the second-order partial-derivative expressions in equation 20 follow a similar grid cyclicity when applying the Gi and Di operators. For example, applying the second partial derivative along the first dimension cycles between ½f; f!½v; f!½f; f or ½v; v!½f; v! D1

G1

G1

formation from neighboring grid points to approximate the mixed partial derivatives of the Laplacian operator using OðΔx2 Þ bilinear or OðΔx4 Þ bicubic interpolation. Using an interpolation approach, though, diminishes the accuracy of numerical solutions to that of the interpolation scheme and gives rise to an earlier onset of numerical instability for scenarios involving complex geometry. 3D staggerings Solving the 3D tensorial AWE using an FSG approach requires four complementary SGs to ensure the availability of all mixed partial-differential operators. The first row of Table 1 presents the four chosen SGs, on which wavefields um are defined. Using the same symbolic notation as above, rows two through five show the cyclicity in grid dimensions when applying the MFD Laplacian operator to wavefield um . Note that a full complementary set of wavefields is defined for each and every staggered grid. Wavefield coupling is also required to apply the MFD-RBCs presented in equation 18 for scenarios involving nonorthogonal coordinate systems, the operations of which are illustrated in six two through eight. Wavefield injection/extraction weights Achieving the desired 3D source impulse at a given point with index ½i; j; k requires a judicious injection of wavefield energy on

D1 ½v; v for the two SGs, respectively. However, applying the following progression for mixed second-order partial derivatives ½f; f!½v; f!½v; v and ½v; v!½f; v!½f; f leads to wavefield

Table 2. Wavefield-injection locations and weights for the four FSG SGs assuming a primary injection point of i; j; k in the f; f; f SG.

derivatives defined only at grid points of the complementary SG. All required wavefield partial derivatives, though, are made available by swapping colocated grid information, which effectively couples the two SGs to form a single complementary FSG system. Note that the combination of ½f; v ∈ Rðn1 þ2Þ×ðn2 þ1Þ and ½v; f ∈ Rðn2 þ1Þ×ðn2 þ2Þ also represents a set of 2D complementary SGs that form an FSG. Finally, an alternate solution approach would be to apply the MFDTD formula on an SSG, and then use wavefield in-

Value

D2

G1

G1

D2

Table 1. The four complementary SG staggerings for wavefield um used to generate MFDTD solutions of the 3D tensorial AWE. The first line shows the four SG staggerings for wavefield um . Rows two through five present the SG cyclicity for all second-order partial derivatives required to compute the 3D tensorial Laplacian operator. Rows six through eight present the grid cyclicity for gradient operators required to evaluate the MFD-RBCs in equation 20. Value um D1 G1 um D1 G2 um D1 G3 um D2 G3 um G1 um G2 um G3 um

, , , ,

D2 G2 um , D3 G3 um D2 G1 um D3 G1 um D3 G2 um

SG 1

SG 2

SG 3

SG 4

½f; f; f ½f; f; f ½v; v; f ½v; f; v ½f; v; v ½v; f; f ½f; v; f ½f; f; v

½f; v; v ½f; v; v ½v; f; v ½v; v; f ½f; f; f ½v; v; v ½f; f; v ½f; v; f

½v; f; v ½v; f; v ½f; v; v ½f; f; f ½v; v; f ½f; f; v ½v; v; v ½v; f; f

½v; v; f ½v; v; f ½f; f; f ½f; v; v ½v; f; v ½f; v; f ½v; f; f ½v; v; v

SG 1

um ½f; f; f Location 1 ½i; j; k Location 2 — Location 3 — Location 4 — pffiffiffiffiffi−1 Weighting (× jgj ) 1

SG 2

SG 3

SG 4

½f; v; v ½i; j − 1; k ½i; j þ 1; k ½i; j; k − 1 ½i; j; k þ 1 0.25

½v; f; v ½i − 1; j; k ½i þ 1; j; k ½1; j; k − 1 ½i; j; k þ 1 0.25

½v; v; f ½i − 1; j; k ½i þ 1; j; k ½i; j − 1; k ½i; j þ 1; k 0.25

Table 3. Pseudocode for generating an MFDTD solution of the 3D AWE. Step description

Equation

For all m in N t steps: — For all four complementary SGs: — m 1) Apply three Gj operators to u . 27 2) Apply nine ψ −1 Di ψΓij operators. 28 3) Apply MFD-RBC on all six faces. 17 4) Apply free surface as required. — m 5) Inject source wavelet F with appropriate weights. Table 2 6) Apply MFDTD update formula. 20 7) Cycle wavefields: um → um−1 and umþ1 → um . — Sum output wavefield using all four SGs with weighting Table 2 factors.

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Solving tensorial 3D AWE: MFDTD approach all four SGs (Davydycheva et al., 2003; de la Puente et al., 2014). This can be accomplished by spraying out the source wavelet at the locations listed inpTable ffiffiffiffiffi 2 with the corresponding weightings. Multiplication by jgj−1 is required to achieve the correct (volumetric) balance of injected wavefield energy. Generating the desired output single physical wavefield at index ½i; j; k requires a summation of energy from the four SGs using the adjoint operation (i.e., the locations and weights) of that presented in Table 2.

along with relevant equations. The first two steps involve successively applying the three MFD gradient operators Gi to um , the appropriate metric weighting factors, three MFD divergence operators Di to the three wavefield gradients (for nine fields in total), and finally the geometric postweighting factor to complete the calculation of the tensorial 3D Laplacian operator. One then applies the MFDRBC operator, injects the source wavelet, updates the MFDTD wavefield, and cycles the wavefields for the ensuing time step.

Numerical implementation The MFDTD algorithm described above represents the kernel operation in simulating a wavefield solution at the future time step, umþ1 , from the current and previous solutions, um and um−1 , respectively. Table 3 presents an overview of the key procedural steps

T189

3D NUMERICAL EXAMPLES This section presents 3D numerical examples of the implemented MFDTD scheme for acoustic wavefield simulation in sheared Cartesian, IB, and topographic coordinates. Each example uses a (semi-) analytic mapping between the x- and ξ-coordinates (equation 2) that

a)

b)

c)

d)

Figure 2. Wavefield snapshots for the MFDTD solution for a 3D sheared Cartesian mesh. (a) Wavefield at t ¼ 0.162 s, (b) t ¼ 0.28 s, (c) t ¼ 0.40 s, and (d) wavefield generated in 3D Cartesian coordinates for comparison purposes with panel (c). The locations of the extracted traces shown in Figure 3 are illustrated by the lines at ½x2 ; x3  ¼ ½0.625; 0.625 km in the upper left sections of the four panels.

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

T190

Shragge and Tapley

is based on either linear or quadratic Bézier interpolation functions. We also provide the resulting inverse metric tensor (equation 3) and the square root of the metric tensor discriminant (equation 6), which are sufficient for specifying the tensorial 3D Laplacian operator (equation 7). To help guide our choices of grid sampling and time-step selection (i.e., for stability and numerical dispersion considerations), we follow the approach prescribed in Shragge (2014b). In particular, we select sampling rates such that there is effectively a minimum of eight points per wavelength sampling in the physical domain. We note that numerical accuracy and stability formally depends on many characteristics of an FD implementation (e.g., geometric stretching and skewness, the order of FD operator and stencil coefficients applied, boundary implementation, overall runtime, and solution strategy), and we emphasize that this approach only represents a guideline for parameter selection. Finally, for visualization purposes, we use 3D sinc-based operators to interpolate simulated

ξ-coordinate wavefields to a uniformly sampled Cartesian domain (for details, see Shragge, 2014b).

Example 1: 3D sheared Cartesian coordinates We first validate the developed 3D MFDTD algorithm using analytically and numerically generated wavefield simulation results on a sheared Cartesian coordinate system. Although this test is not intended to illustrate a practical solution for any particular seismic application, we include it to demonstrate the accuracy of the above theory and of our implemented numerical solution. This represents a reasonable test because a sheared Cartesian mesh introduces nonorthogonal geometry in the MFDTD propagation (equation 20) and the MFD-RBC scheme (equation 19), but it still affords an analytic result against which to verify generated numerical solutions. A 3D sheared Cartesian coordinate transformation is given by

2

3 2 3 x1 ξ1 4 x2 5 ¼ 4 ξ2 þ ξ1 cos θ2 5; x3 ξ3 þ ξ1 cos θ3

a)

(21)

where θ2 and θ3 are the shear angles of the coordinate system in the ξ2 and ξ3 directions. The inverse metric tensor associated with this coordinate transformation is

2

1 ½gij  ¼ 4 − cos θ2 − cos θ3

b)

Figure 3. Extracted waveform traces and differences: (1) Cartesian; (2) Cartesian-sheared Cartesian difference, (3) sheared Cartesian, (4) analytic-sheared Cartesian difference, and (5) analytic. Waveform example at (a) time t ¼ 0.174 s prior to boundary reflection, and (b) time t ¼ 0.246 s postboundary reflection. Note that the freesurface condition is applied only to the left.

− cos θ2 1 þ cos2 θ2 cos θ2 cos θ3

3 − cos θ3 cos θ2 cos θ3 5: (22) 1 þ cos2 θ3

Thep square ffiffiffiffiffi root of the metric tensor for this transformation is given by jgj ¼ 1. We perform our 3D wavefield simulation test on a ½N ξ1 ; N ξ2 ; N ξ3  ¼ ½161; 201; 201 computational domain sampled at Δξi ¼ 6.25 m, using dip angles of θ2 ¼ θ3 ¼ 75°, a 17.5 Hz Ricker wavelet, and time step of Δt ¼ 0.6 ms. The total simulation time is of sufficient duration to ensure that the propagated wavefield reflects from all of the bounding surfaces including the free surface with the required −1 reflection coefficient. Figure 2 presents the wave-propagation results interpolated to a 3D Cartesian mesh, in

Figure 4. Bathymetry adapted from a region of Australia’s Northwest Shelf. The black lines indicate the locations of mesh extracted at ξ3 ¼ 0.3125 km in Figure 5a, and at ξ2 ¼ 1.025 in Figure 5b.

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Solving tensorial 3D AWE: MFDTD approach which the active computational domain is illustrated by the darker background color. Figure 2a shows the wavefield prior to interaction with the boundary, whereas Figure 2b depicts the wavefield interaction with the domain boundary. Figure 2c shows the wavefield postreflection from the free surface with the required polarity reversal. For comparison purposes, Figure 2d presents the wavefield corresponding to Figure 2c computed directly in 3D Cartesian coordinates. Note that the free-surface and bottom reflections appear very similar to those generated in the 3D sheared Cartesian; however, the reflections from the other boundaries are significantly different due to the sheared boundary. Figure 3 presents several extracted waveform trace and trace differences among the following: the interpolated sheared 3D Cartesian result shown in Figure 2a–2c (trace 3 in each subplot); the 3D Cartesian result presented in Figure 2d (trace 1 in each subplot); and a theoretically calculated wavefield for the given geometry, wavelet, and velocity model (trace 5 in each subplot). The locations of the trace extraction are indicated in Figure 2 by the crosshairs in the uppermost slices in the four panels. Trace 2 shows the waveform differences between the 3D sheared Cartesian and (nonsheared) Cartesian, whereas trace 4 presents the waveform differences between the numerically generated 3D sheared Cartesian and analytic results. Figure 3a depicts the wavefields before they encounter the boundary. We observe a good match a) between the sheared and nonsheared 3D Cartesian coordinate results, in which the minor differences in waveforms can be attributed to differences in Shragge (2016): (1) source wavefield injection onto the computational grids, (2) numerical dispersion due to inclusion of different FD terms with different coefficients (i.e., the mixed partial derivatives), and (3) interpolation of output wavefields to a Cartesian mesh. The differences between the analytic and the sheared 3D Cartesian solution are again very similar (trace 4), with a slightly higher mismatch owing to modest amplitude differences. We note that these waveforms are within three wavelengths of the injection point, and thus they are still moderately affected by near-field numerical behavior. Figure 3b shows the wavefields after encounterb) ing the wavefield boundary, in which we apply the free-surface boundary condition only on the left side at ξ1 ¼ 0. The numerically generated waveforms to the left are again quite consistent, which indicates that nonorthogonal geometry in the 3D sheared Cartesian example is accounted for by incorporating wavefield fluxes (i.e., Γnj Gj ) in the boundary operator Bξn (equation 18). The somewhat lower differences between the theoretical and 3D sheared Cartesian result than in Figure 3a are attributed to better amplitude matching due to being roughly an additional wavelength from the source injection location.

T191

tom, is a useful strategy for improving the accuracy of FDTD simulation results in several scenarios. For example, this reduces or even precludes introducing staircasing artifacts that arise when coarsely discretizing a curvilinear surface on a Cartesian mesh. We use quadratic Bézier curves in x1 to smoothly interpolate the coordinate mesh between three surfaces: (1) the free surface at x1 ¼ ξ1 ¼ 0; (2) a flat layer at depth x1 ¼ ξ1 ¼ a, where a is the maximum depth; and (3) a water-bottom surface W ¼ W½ξ1 ¼ ζ; ξ2 ; ξ3 , where ζ is the location of the water-bottom surface in the computational grid (i.e., 0 < ζ < a). A 3D IB coordinate system (Shragge, 2014b) may then be expressed as

2

3 2 ξ1 3 1 1 x1 ðζ−1Þζ ðWðξ − 1Þ þ aζðζ − ξ ÞÞ 4 x2 5 ¼ 4 5: ξ2 3 3 x ξ

(23)

The inverse metric tensor associated with the IB coordinate transformation is given by

Example 2: Internal boundary coordinates Modeling wavefield propagation on meshes conforming to an IB, such as a rugose water bot-

Figure 5. Cross sections of a 3D IB coordinate system conforming to the water-bottom surface (thick line) in Figure 4. (a) Inline section extracted at ξ3 ¼ 0.3125 km. (b) Crossline section extracted at ξ2 ¼ 1.025 km.

Shragge and Tapley

T192 2

3 γ −1 ½ζ2 ðζ −1Þ2 þðW 22 þW 23 Þðξ1 Þ2 ðξ1 −1Þ2  W 2 ξ1 ð1−ξ1 Þ W 3 ξ1 ð1−ξ1 Þ 7 1 1 W 2 ξ ð1−ξ Þ 1 0 4 5; γ W 3 ξ1 ð1−ξ1 Þ 0 1

16 ½gij  ¼

(24)

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

2ξ1 Þ

aζð2ξ1

where γ ¼ Wð1 − þ − ζÞ, and W i ¼ ð∂W∕∂ξi Þ represents the spatial partial derivative of the water-bottompsurface in ffiffiffiffiffi the ith direction. The metric tensor square root is jgj ¼ ½aζ ðζ − 2ξ1 Þ þ 2ð2ξ1 − 1Þ∕½ζðζ − 1Þ. To provide a 3D IB-coordinate propagation test, we examine a water-bottom model adapted from bathymetric observations from Australia’s Northwest Shelf (see Figure 4). The minimum and maximum values of the bathymetric profile are 0.33 and 0.69 km, respectively. Figure 5a and 5b presents inline and crossline sections of the developed 3D mesh, in which the controlling IB surface W is visible as the thicker black line. Assuming a ¼ 1 km and ζ ¼ 0.5 km, we construct a mesh that is somewhat rough (i.e., locally more nonorthogonal) in locations where the surface changes rapidly in depth, but it is otherwise smooth. We use water and sediment velocities of 1.5 and 2.5 km∕s, respectively, for the upper (0.0 km ≤ ξ1 < 0.5 km) and lower (0.5 km ≤ ξ1 ≤ 1.0 km) layers. We perform our 3D wavefield simulation test on a ½N ξ1 ; N ξ2 ; N ξ3  ¼ ½161; 321; 321 computational domain sampled

Figure 7. Sample topographic surface from San Francisco Bay area exaggerated by a 2.5× factor. Minimum and maximum elevations are 0.0 and 2.5 km, respectively. The black lines indicate the locations of mesh extracted at ξ3 ¼ 2.50 and ξ2 ¼ 4.32 km in Figure 8a and 8b, respectively.

a)

c)

b)

d)

Figure 6. Wavefield snapshots for the MFDTD solution for a 3D IB coordinate system for the water-bottom surface shown in Figure 4, for water (light gray) and sediment (dark gray) velocities of 1.5 and 2.5 km∕s, respectively. Wavefield snapshots are shown at (a) t ¼ 0.162 s, (b) t ¼ 0.28 s, (c) t ¼ 0.40 s, and (d) t ¼ 0.52 s.

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Solving tensorial 3D AWE: MFDTD approach at Δξi ¼ 6.25 m, which represents a range of 3.85 m ≤ Δx1 ≤ 8.64 m in the physical Cartesian coordinates. We use a 25 Hz Ricker wavelet with a Δt ¼ 0.5 ms time step that is injected into the boundary layer. The total simulation time is of sufficient duration to ensure that the propagated wavefield reaches the base of the IB mesh. Figure 6 presents wavefield simulation results computed in IB coordinates and then interpolated back to a Cartesian grid. The four panels sequentially illustrate evolving acoustic wave propagation and interaction with the irregular water-bottom surface. In particular, water-bottom reflections in later panels are free of wavefield diffractions associated with staircasing effects common in Cartesian meshing of complex bathymetry.

Example 3: 3D topographic coordinates To develop a 3D topographic mesh, we define a generic 2D topographic surface, T ¼ Tðξ2 ; ξ3 Þ, as a function of the two lateral coordinates. Using this definition, we specify a 3D topographic coordinate system using linear Bézier interpolation functions in x1 :

3 2 1 3 x1 ξ1 þ Tð1 − ξa Þ 4 x2 5 ¼ 4 5; ξ2 x3 ξ3

T193

The memory complexity of an FSG approach scales according to 4 × 13N 3 , where four is the number of SGs and 13N 3 is due to the requirement of concurrently holding in memory 3N 3 time steps, a N 3 velocity model, and 3 × 3N 3 wavefield gradients for each SG. Because we follow an approach in which the geometric transforms are defined semianalytically, this forestalls an additional 12N 3 memory requirement that otherwise would be required to store the three x fields and their nine gradients used in equation 4. The computational complexity of the FSG approach roughly scales as 4× that of an SSG implementation, save for a modest reduction caused by precluding the need for bilinear or bicubic interpolation. However, relative to a SSG approach, the computational overhead is also somewhat offset by the reduced numerical dispersion that results from a weighted summation and corresponding destructive interference of four um

a)

2

(25)

where the range of depth axis is defined by ξ1 ¼ ½0; a. The associated inverse metric tensor is

2 ½gij  ¼

a2 þða−ξ1 Þ2 ðT 22 þT 23 Þ ða−TÞ2 6 6 ðξ1 −aÞT 2 4 a−T ðξ1 −aÞT 3 a−T

ðξ1 −aÞT 2 a−T

ðξ1 −aÞT 3 a−T

1 0

0 1

3 7 7; 5

(26)

where T i ¼ ð∂T∕∂ξi Þ. The square root of the metric tensor is given pffiffiffiffiffi by jgj ¼ 1 − T∕a. Figure 7 presents the 2D topographic surface chosen to demonstrate MFDTD wavefield simulation in 3D topographic coordinates. The minimum and maximum elevations are 0.0 and 2.5 km, respectively, with slopes reaching 65° at some locations. Figure 8 presents inline and crossline sections of the mesh plotted at a 1:1 scale. The strongest mesh distortion occurs at the surface of the grid, which “heals” as one moves to greater depths. We conduct a 3D wavefield simulation test on a ½N ξ1 ; N ξ2 ; N ξ3  ¼ ½201; 301; 301 mesh sampled at Δξi ¼ 40.0 m, which represents a range of 40.0 m ≤ Δx1 ≤ 52.0 m in physical Cartesian coordinates. We use a 10 Hz Ricker wavelet sampled at 2.5 ms and use a total simulation time of 1.75 s. Figure 9 presents the wavefield snapshots taken before (Figure 9a) and after (Figure 9b–9d) interaction with the free surface. As the wavefield evolves, we observed complex scattering from the free surface, which is best observed in the map-view sections, in which the scattered energy is completely encircled by the unaffected horizontally propagating wavefield components.

b)

DISCUSSION Although applying the developed MFDTD algorithm on an FSG is demonstrably beneficial for generating accurate and low-dispersion solutions of the 3D tensorial AWE for complex geometry, following this approach does come with additional computational overhead.

Figure 8. Cross sections through 3D topographic mesh conforming to elevation map in Figure 7. The maximum slope shown is 65°. (a) Inline section at crossline ξ3 ¼ −2.5 km. (b) Crossline section at inline ξ2 ¼ −4.32 km.

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

T194

Shragge and Tapley

with different dispersion characteristics (Davydycheva et al., 2003; Lisitsa and Vishnevskiy, 2010). Alternatively, one could use somewhat larger γΔξi sampling intervals (where γ ≈ 1.15 in our experience) in an FSG approach to generate dispersion roughly equivalent to that calculated using an equivalent SSG approach on a Δξi sampled grid, but at a reduction in computational complexity roughly proportional to γ 4 (i.e., 1.154 ¼ 1.75× in our experience). In this work, we have not presented any strategies for eliminating unwanted reflections arising from wavefield interactions with the domain boundary. Solano-Feo et al. (2016) implement an MFD ap-

proach based on Reynolds (1978) work on (now low-order) one-way attenuating boundaries. However, this work is important in that it demonstrates that an attenuating boundary operator can be codified directly within an MFD matrix operator framework. Our initial tests indicate that the well-established absorbing boundary condition (Clayton and Engquist, 1977) and the more recent double absorbing boundary condition (Baffet et al., 2014; Hagstrom et al., 2014) perform well in attenuating boundary reflections — even when taking generalized geometry into account. However, due to the significant theoretical treatment required to demonstrate the validity of these

a)

c)

b)

d)

Figure 9. Wavefield snapshots for the MFDTD solution for a 3D topographic mesh are shown in Figure 8. The domain boundary denoted by the dark to light gray transition. Wavefield snapshots at: (a) t ¼ 0.75 s, (b) t ¼ 1.0625 s, (c) t ¼ 1.375 s, and (d) t ¼ 1.6875 s.

Solving tensorial 3D AWE: MFDTD approach 2

operators within an MFD framework, we feel that this topic falls beyond the scope of the present work.

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

CONCLUSION This paper has demonstrated the utility of applying an MFD + FSG approach to generate solutions of the 3D tensorial AWE. We demonstrate that generalized geometry can be directly incorporated into analytic and discrete versions of the 3D tensorial Laplacian operator, and subsequently develop an MFDTD strategy appropriate for generating uniformly accurate OðΔx4 Þ solutions of the 3D AWE. We apply the 3D MFDTD approach on an FSG comprised of a set of four complementary SGs, which are defined such that any cross-derivative not available on its own grid is always available at a colocated point on another SG. This, along with judicious weighting of injected and extracted wavefield energy, ensures that accurate and stable MFDTD modeling results can be generated for 3D meshes exhibiting complex geometry. We use a 3D sheared Cartesian mesh to validate the implemented MFD operators and MFD + RBC scheme against an analytic solution. We simulate accurate and stable numerical solutions on a 3D IB mesh that smoothly conforms to an irregular water-bottom surface, and the resulting impulse response demonstrably forestalls staircasing effects commonly observed in Cartesian simulations involving meshed curvilinear water-bottom surfaces. We then present 3D wavefield simulation results on a topographic coordinate system in the presence of 2.5 km topographic relief with slopes approaching 65°. The resulting impulse responses illustrate the wavefield complexity generated by the interaction of an acoustic wavefield with a complex free surface. Finally, although the computational and memory complexity of the MFDTD approach are more significant than using an NSG or an SSG plus wavefield interpolation, the improved accuracy and stability suggest that the developed technique would be applicable for a variety of RTM and FWI scenarios involving complex geometry.

ACKNOWLEDGMENTS J. Shragge acknowledges the support of Woodside Pty Ltd. and the UWA:RM consortium sponsors. Woodside Pty Ltd. and J. Claerbout are also thanked for the Northwest Shelf bathymetric survey and San Francisco Bay area data set, respectively. Simulation results were computed using Pawsey Centre HPC facilities (https:// www.pawsey.org.au/). Reproducible numerical examples were generated using the Madagascar package (http://www.ahay.org). Analytic coordinate system geometry results were verified using the Mathematica software package (http://www.wolfram.com/ mathematica/).

APPENDIX A

T195

Gi ¼

−1152 10;063 2483 −3309 2099 −697 9768 3256 3256 4884 6 407 3256 17 3 −5 1 6 0 −11 12 24 8 24 24 6 1 −9 9 −1 6 0 0 24 8 8 24 6 1 −9 9 −1 6 0 0 24 8 8 24 6

1 6 6 Δξi 6 6 6 6 6 6 4

.. . ::: ::: ::: :::

.. . ::: ::: ::: :::

.. . ::: ::: ::: :::

..

. ::: ::: ::: :::

..

..

.

::: ::: ::: ::: .. .

0 0 0 0 ..

.

.

::: ::: ::: ::: .. . 0

::: ::: ::: ::: .. . 0 0 0

3

−9 9 −1 8 8 24 1 −9 9 −1 0 24 8 8 24 −1 5 −3 −17 11 24 24 8 24 12 697 −2099 3309 −2483 −10;063 1152 3256 4884 3256 3256 9768 407 1 24

7 7 7 7 7 7 7 7 7; 7 7 7 7 7 7 5

(A-1)

where three vertical and horizontal dots indicate null entries and three diagonal dots indicate a repetition of stencil coefficients along the corresponding diagonal. The corresponding 1D MFD divergence operator is given by 2

Di ¼

0

0

0

0

0

.. . ::: ::: ::: :::

.. . ::: ::: ::: :::

.. . ::: ::: ::: :::

..

..

0

909 6091 −1165 129 −25 6 −4751 15;576 5192 2596 15;576 6 5192 1298 1 −9 9 −1 6 0 0 24 8 8 24 6 1 −9 9 −1 6 0 0 24 8 8 24 6

1 6 6 Δξi 6 6 6 6 6 4

. ::: ::: ::: :::

.

..

.

0 0 0 0 ..

.

::: ::: ::: ::: .. .

::: ::: ::: ::: .. . 0

0

0

::: ::: ::: ::: .. . 0 0

3

7 7 7 7 7 7 7 7. 7 7 7 7 7 4751 5

−9 9 −1 8 8 24 1 −9 9 −1 0 24 8 8 24 25 −129 1165 −6091 −909 15;576 2596 5192 15;576 1298 5192 1 24

0

0

0

0 (A-2)

Where, again, three vertical and horizontal dots indicate null entries and three diagonal dots indicate a repetition of stencil coefficients along the corresponding diagonal. The diagonal elements of weight matrix P corresponding to an OðΔx4 Þ operator defined on a f grid are given by



 407 473 343 1177 1177 343 473 407 ; ; ; ;1; :::;1; ; ; ; P ¼ diag ; 1152 384 384 1152 1152 384 384 1152 (A-3)

where all unspecified entries are unity. The diagonal elements of weight matrix Q corresponding to an OðΔx4 Þ operator defined on a v grid are given by

 649 143 75 551 551 75 143 649 ; ; ; ;1; :::;1; ; ; ; ;0 . Q ¼ diag 0; 576 192 64 576 576 64 192 576 (A-4) 

MFD OPERATORS This appendix presents OðΔx4 Þ expressions for MFD operators Gi , Di , and Bn , as well as weight matrices P and Q derived in Castillo and Miranda (2013). The 1D MFD gradient operator used in the numerical examples is given by

Where, again, all unspecified entries are unity. Weight operators P and Q are used with Gi and Di to form boundary operator, Bn according to equation 15. The OðΔx4 Þ 1D MFD boundary operator is given by

Shragge and Tapley

T196

Downloaded 06/19/17 to 203.17.70.53. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

2

−1 6 187 6 3072 6 3341 6 27;648 6 −1103 6 3072 6 2099 6 6 9216 6 −697 6 13;824 1 6 6 .. Bn ¼ . Δξn 6 6 6 ::: 6 6 ::: 6 6 ::: 6 6 ::: 6 6 4 ::: :::

0

−1567 4608 319 9216 523 1024 −2365 9216 473 9216

.. . ::: ::: ::: ::: ::: :::

0

0

13211 27;648 −171 1024 −321 1024 73 27;648

0 .. . ::: ::: ::: ::: ::: :::

−1165 4608 319 27;648 173 1024 75 1024 −25 27;648

.. . ::: ::: ::: ::: ::: :::

0 .. . 0

::: ::: ::: ::: ::: ::: .. . 0

::: ::: ::: ::: ::: ::: .. .

0

0

0

0

0

0

0

43 −25 768 13;824 −11 0 1536 −25 25 512 13;824

0 .. . ::: ::: ::: ::: ::: :::

::: ::: ::: ::: ::: ::: .. . 0

25 27;648 −25 25 −75 −73 13;824 512 1024 27;648 11 −173 321 0 1536 1024 1024 −319 171 0 0 27;648 1024 25 −43 1165 −13;211 13;824 768 4608 27;648

0

::: ::: ::: ::: ::: ::: .. .

::: ::: ::: ::: ::: ::: .. .

3

7 7 7 7 7 7 7 7 7 7 7 7 7; 7 −473 697 7 7 9216 13;824 7 2365 −2099 7 9216 9216 7 −523 1103 7 1024 3072 7 −319 −3341 7 9216 27;648 7 1567 −187 7 5 4608 3072 0 1

(A-5)

where all unspecified entries are zeros values.

REFERENCES Appelö, D., and N. A. Petersson, 2009, A stable finite difference method for the elastic wave equation on complex geometries with free surfaces: Communications in Computational Physics, 5, 84–107. Baffet, D., T. Hagstrom, and D. Givoli, 2014, Double absorbing boundary formulations for acoustics and elastodynamics: SIAM Journal on Scientific Computing, 36, A1277–A1312, doi: 10.1137/130928728. Bernth, H., and C. Chapman, 2011, A comparison of the dispersion relations for anisotropic elastodynamic finite-difference grids: Geophysics, 76, no. 3, WA43–WA50, doi: 10.1190/1.3555530. Carcione, J. M., 1994, The wave equation in generalized coordinates: Geophysics, 59, 1911–1919, doi: 10.1190/1.1443578. Castillo, J. E., and G. F. Miranda, 2013, Mimetic discretization methods: CRC Press. Cheng, A., and J. O. Blanch, 2008, Numerical modeling of elastic wave propagation in a fluid-filled borehole: Communications in Computational Physics, 3, 33–51. Clayton, R., and B. Engquist, 1977, Absorbing boundary conditions for acoustic and elastic wave equations: Bulletin of the Seismological Society of America, 67, 1529–1540. Cockburn, B., G. Karniadakis, and C. Shu, 2000, Discontinuous Galerkin methods: Theory, computational and applications: Springer-Verlag. Davydycheva, S., V. Druskin, and T. Habashy, 2003, An efficient finite difference scheme for electromagnetic logging in 3D anisotropic inhomogeneous media: Geophysics, 68, 1525–1536, doi: 10.1190/1 .1620626. de la Puente, J., M. Ferrer, M. Hanzich, J. Castillo, and J. Cela, 2014, Mimetic seismic wave modeling including topography on deformed staggered grids: Geophysics, 79, no. 3, T125–T141, doi: 10.1190/geo20130371.1. Ely, G., S. Day, and J.-B. Minster, 2008, A support-operator method for viscoelastic wave modeling in 3-D heterogeneous media: Geophysical Journal International, 172, 331–344, doi: 10.1111/j.1365-246X.2007.03633.x. Ely, G., S. Day, and J.-B. Minster, 2009, A support-operator method for 3-D rupture dynamics: Geophysical Journal International, 177, 1140–1150, doi: 10.1111/j.1365-246X.2009.04117.x. Guggenheimer, H., 1977, Differential geometry: Dover Publications Inc. Hagstrom, T., D. Givoli, D. Rabinovich, and J. Bielak, 2014, The double absorbing boundary method: Journal of Computational Physics, 259, 220–241, doi: 10.1016/j.jcp.2013.11.025.

Hestholm, S., 1999, Three-dimensional finite difference viscoelastic wave modeling including surface topography: Geophysical Journal International, 139, 852–878, doi: 10.1046/j.1365-246x.1999.00994.x. Hestholm, S., M. Moran, S. Ketcham, T. Anderson, M. Dillen, and G. McMechan, 2006, Effects of free-surface topography on moving seismic-source modeling: Geophysics, 71, no. 6, T159–T166, doi: 10.1190/1.2356258. Hestholm, S., and B. Ruud, 2002, 3D free-boundary conditions for coordinate-transform finite-difference seismic modeling: Geophysical Prospecting, 50, 463–474, doi: 10.1046/j.1365-2478.2002.00327.x. Komatitsch, D., F. Coutel, and P. Mora, 1996, Tensorial formulation of the wave equation for modeling curved interfaces: Geophysical Journal International, 127, 156–168, doi: 10.1111/j.1365-246X.1996.tb01541.x. Komatitsch, D., and J.-P. Vilotte, 1998, The spectral element method: An efficient tool to simulate the response of 2D and 3D geological structures: Bulletin of the Seismological Society of America, 88, 368–392. Lebedev, V., 1964, Difference analogues of orthogonal decompositions, basic differential operators and some boundary value problems: USSR Computational Mathematics and Mathematical Physics, 4, 449–465. Lipnikov, K., and L. Huang, 2008, A mimetic finite-difference method for acoustic-wave modeling on arbitrary meshes: 78th Annual International Meeting, SEG, Expanded Abstracts, 2067–2071. Lipnikov, K., G. Manzini, and M. Shashkov, 2014, Mimetic finite difference method: Journal of Computational Physics, 257, 1163–1227, doi: 10 .1016/j.jcp.2013.07.031. Lisitsa, V., and D. Vishnevskiy, 2010, Lebedev scheme for the numerical simulation of wave propagation in 3D anisotropic elasticity: Geophysical Prospecting, 58, 619–635, doi: 10.1111/j.1365-2478.2009.00862.x. Madariaga, R., 1976, Dynamics of an expanding circular fault: Bulletin of the Seismological Society of America, 66, 163–182. Marfurt, K., 1984, Accuracy of finite-difference and finite-element modeling of the scalar and elastic wave equations: Geophysics, 49, 533–549, doi: 10 .1190/1.1441689. Nichols, D., 1994, Imaging complex structures using band-limited Green’s functions: Ph.D. thesis, Stanford University. Ohminato, T., and B. A. Chouet, 1997, A free-surface boundary condition for including 3D topography in the finite-difference method: Bulletin of the Seismological Society of America, 87, 494–515. Reynolds, A., 1978, Boundary conditions for the numerical solution of wave propagation problems: Geophysics, 43, 1099–1110, doi: 10 .1190/1.1440881. Rojas, O., S. M. Day, J. Castillo, and L. Dalguer, 2008, Modeling of rupture propagation using high-order mimetic finite-differences: Geophysical Journal International, 172, 631–650, doi: 10.1111/j.1365-246X.2007.03651.x. Sava, P., and S. Fomel, 2005, Riemannian wavefield extrapolation: Geophysics, 70, no. 3, T45–T56, doi: 10.1190/1.1925748. Shragge, J., 2008, Riemannian wavefield extrapolation: Non-orthogonal coordinate systems: Geophysics, 73, no. 2, T11–T21, doi: 10.1190/1.2834879. Shragge, J., 2014a, Reverse time migration from topography: Geophysics, 79, no. 4, S141–S152, doi: 10.1190/geo2013-0405.1. Shragge, J., 2014b, Solving the 3D acoustic wave equation on generalized structured meshes: A finite-difference time-domain approach: Geophysics, 79, no. 6, T363–T378, doi: 10.1190/geo2014-0172.1. Shragge, J., 2016, Acoustic wave propagation in tilted transversely isotropic media: Incorporating topography: Geophysics, 81, no. 5, C265–C278, doi: 10.1190/geo2015-0311.1. Solano-Feo, F., J. M. Guevara-Jordan, O. Rojas, B. Otero, and R. Rodriguez, 2016, A new mimetic scheme for the acoustic wave equation: Journal of Computational and Applied Mathematics, 295, 2–12, doi: 10.1016/j.cam .2015.09.037. Synge, J. L., and A. Schild, 1978, Tensor calculus: Dover Publications Inc. Virieux, J., 1986, P-SV wave propagation in heterogeneous media: Velocitystress finite-difference method: Geophysics, 51, 889–901, doi: 10.1190/1 .1442147. Zhang, W., Z. Zhang, and X. Chen, 2012, Three-dimensional elastic wave numerical modeling in the presence of surface topography by a collocatedgrid finite-difference method on curvilinear grids: Geophysics Journal International, 190, 358–378, doi: 10.1111/j.1365-246X.2012.05472.x.

Solving the tensorial 3D acoustic wave equation: A ...

finite-difference time-domain approach. Jeffrey Shragge1 and Benjamin Tapley2. ABSTRACT. Generating accurate numerical solutions of the acoustic wave.

2MB Sizes 8 Downloads 272 Views

Recommend Documents

Solving the 3D acoustic wave equation on generalized ...
difference time-domain (FDTD) methods represent the most com- mon numerical approach for generation solutions of the 3D acoustic wave equation, and have ...

Solving 3D anisotropic elastic wave equations on ...
try-sized 3D anisotropic elastic data sets, we parallelized computation across multiple GPU devices ... 1Formerly the University of Western Australia, Centre for Petroleum Geoscience and CO2 Sequestration, Crawley, Australia. Presently the University

Cylindrical wave, wave equation, and method of ...
Sep 19, 2007 - Schrodinger wave equation in which the method of separation of variables is used for obtaining the general spreading of the wave.) 2 Invalidity of the separation of variables for obtaining cylindrical wave function from the wave equati

Surface acoustic wave device and communication device
Mar 31, 2004 - (10) Patent Number: (45) Date of Reissued Patent: USO0RE39975E. US RE39,975 E. Jan. 1, 2008. (54) SURFACE ACOUSTIC WAVE DEVICE AND. COMMUNICATION DEVICE. (75) Inventor: Michio Kadota, Kyoto (JP). (73) Assignee: Murata Manufacturing Co.

Universal Wave Equation Worksheet.pdf
... dock measure about 1.5 m. from crest to crest and appear to be traveling at about 3.0 m/s. Determine the frequency and period of. these waves. Page 1 of 2 ...

Universal Wave Equation Worksheet.pdf
Whoops! There was a problem loading this page. Universal ... ksheet.pdf. Universal ... ksheet.pdf. Open. Extract. Open with. Sign In. Details. Comments. General Info. Type. Dimensions. Size. Duration. Location. Modified. Created. Opened by me. Sharin

Film Resonance on Acoustic Wave Devices: The ... - Jonathan D. Gough
50-Ω coaxial cable to a HP8512A transmission/reflection unit. Materials and ..... This material is available free of charge via the Internet at http://pubs.acs.org.

Film Resonance on Acoustic Wave Devices: The ... - Jonathan D. Gough
These opportunities make polymer-modified electrodes one of the most active ... electronic and mechanical properties by using two materials in a composite is ..... signature of film resonance.19 A condition for peak splitting to be observable is ...

Interpreting Freeform Equation Solving
tom [email protected]. Abstract. Learners' step-by-step solutions can offer insight into their ... could be beneficial for student learning, it could be frustrating to some learners. Entering individual actions may be ..... The first five action

1 On the Complexity of Non Universal Polynomial Equation Solving ...
computing and numerical analysis developed their own techniques for solving .... f1,...,fs ∈ C[X1,...,Xn] of degree at most d, the complex algebraic variety they ..... The mapping Φ associates to every parameter α ∈ W a system of ...... G.W. St

pdf-0730\acoustic-wave-sensors-theory-design-and-physico ...
... OUR ONLINE LIBRARY. Page 3 of 10. pdf-0730\acoustic-wave-sensors-theory-design-and-phys ... ons-of-modern-acoustics-by-d-s-ballantine-jr-robe.pdf.

1 On the Complexity of Non Universal Polynomial Equation Solving ...
of solving them in any real or future computer. The reason is that this ... on the number of variables and the maximum of the degrees, such that the following equivalence ... not seem to be the best methods of dealing with HN. A third syntactical ...

The Hardness of Polynomial Equation Solving
Mar 20, 2003 - Our main result is the following: let be given such a data structure .... objects is a well known subject in theoretical computer science and the .... the degree of the canonical output equation may increase exponentially with.

CHECKERBOARD MODES AND WAVE EQUATION 1 ...
mesh, the checkerboard modes do not exist at low Mach number when the compressible Navier-Stokes ...... [14] Guillard H. and Murrone A. – Behavior of upwind scheme in the low Mach number limit: III. Preconditioned dissipation for a five equation tw

Acoustic wave propagation in tilted transversely ...
Sep 2, 2016 - The developed finite- difference time-domain numerical solution adapts existing. Cartesian TTI operators to this more generalized geometry.

Least-squares shot-profile wave-equation migration
example, allowing for migration velocity models that vary in depth only (Gazdag, ... for example, generalized inversion can compensate for incomplete data.

Least-squares shot-profile wave-equation migration
Department of Physics, University of Alberta, Edmonton, Alberta, Canada .... (with a chosen parameterization) are migrated images, c is the Earth's velocity, and ...

Exact controllability of the 1-d wave equation from a ...
data (u0,u1) and final data (v0,v1), find a control f such that the solution u of (1.1) satisfies ...... operator defined by the left hand side of (2.5), i.e. ..... [8] A. Khapalov, Mobile point controls versus locally distributed ones for the contro

Extended wave-equation imaging conditions for ...
Sep 7, 2015 - The extended image volume is robust to sparsely sampled data and high levels ... 1The University of Western Australia, School of Earth and Environment and School of Physics, Centre for Energy Geoscience, Crawley, Australia. E-mail: ...

Degeneracy loci and polynomial equation solving
11 Dec 2013 - which became introduced in symbolic semi-numeric computation by the already clas- sical Kronecker algorithm [9,11,17–19,21]. Here we refer to procedures whose inputs are measured in the usual way by syntactical, extrinsic parameters a