Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

22
Finite Elements in Analysis and Design 40 (2004) 2063 – 2084 www.elsevier.com/locate/nel Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane Sanghoon Lee a , Loukas F. Kallivokas a; b; a Department of Civil Engineering, The University of Texas at Austin, 1 University Station C1748, Austin, TX 78712, USA b Institute for Computational Engineering and Sciences, The University of Texas at Austin, USA Received 1 September 2003; accepted 15 February 2004 Abstract We have recently discussed the performance of local second-order two-dimensional absorbing boundary con- ditions of elliptical shape for scattering and radiation problems involving sound-hard obstacles embedded in a full-plane. In this article, using the method of images, we extend the applicability of elliptically shaped trun- cation boundaries to semi-innite acoustic media. For problems in either the time- or the frequency-domains, involving near-surface structures of elongated cross-sections, we show that signicant computational savings are attainable when compared against semi-circular truncation geometries. ? 2004 Elsevier B.V. All rights reserved. Keywords: Absorbing boundary conditions; Half-space or half-plane; Scalar wave equation; Elliptical shape; Time-domain; Frequency-domain 1. Introduction When domain discretization methods are used in the numerical simulation of acoustic waves in un- bounded domains, whether of innite or semi-innite extent, there always arises the need to truncate the unbounded domain to a nite computational one. The truncation is achieved through the intro- duction of an articial boundary that limits the extent of the unbounded domain. For well-posedness of the ensuing initial-and-boundary-value problem (IBVP) over the nite domain, an appropriate (absorbing, silent, articial, etc.) condition need also be prescribed on the truncation boundary. Typ- ically, it is dicult to obtain an exact absorbing condition for arbitrary truncation geometries (if it Corresponding author. Department of Civil Engineering, The University of Texas at Austin, 1 University Station C1748, Austin, TX 78712, USA. Tel.: +1-512-232-5769; fax: +1-512-471-7259. E-mail addresses: [email protected] (S. Lee), [email protected] (L.F. Kallivokas). 0168-874X/$ - see front matter ? 2004 Elsevier B.V. All rights reserved. doi:10.1016/j.nel.2004.02.005

Transcript of Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

Page 1: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

Finite Elements in Analysis and Design 40 (2004) 2063–2084www.elsevier.com/locate/ nel

Local absorbing boundaries of elliptical shape for scalar wavepropagation in a half-plane

Sanghoon Leea, Loukas F. Kallivokasa;b;∗

aDepartment of Civil Engineering, The University of Texas at Austin, 1 University Station C1748, Austin,TX 78712, USA

bInstitute for Computational Engineering and Sciences, The University of Texas at Austin, USA

Received 1 September 2003; accepted 15 February 2004

Abstract

We have recently discussed the performance of local second-order two-dimensional absorbing boundary con-ditions of elliptical shape for scattering and radiation problems involving sound-hard obstacles embedded in afull-plane. In this article, using the method of images, we extend the applicability of elliptically shaped trun-cation boundaries to semi-in nite acoustic media. For problems in either the time- or the frequency-domains,involving near-surface structures of elongated cross-sections, we show that signi cant computational savingsare attainable when compared against semi-circular truncation geometries.? 2004 Elsevier B.V. All rights reserved.

Keywords: Absorbing boundary conditions; Half-space or half-plane; Scalar wave equation; Elliptical shape;Time-domain; Frequency-domain

1. Introduction

When domain discretization methods are used in the numerical simulation of acoustic waves in un-bounded domains, whether of in nite or semi-in nite extent, there always arises the need to truncatethe unbounded domain to a nite computational one. The truncation is achieved through the intro-duction of an arti cial boundary that limits the extent of the unbounded domain. For well-posednessof the ensuing initial-and-boundary-value problem (IBVP) over the nite domain, an appropriate(absorbing, silent, arti cial, etc.) condition need also be prescribed on the truncation boundary. Typ-ically, it is di:cult to obtain an exact absorbing condition for arbitrary truncation geometries (if it

∗ Corresponding author. Department of Civil Engineering, The University of Texas at Austin, 1 University StationC1748, Austin, TX 78712, USA. Tel.: +1-512-232-5769; fax: +1-512-471-7259.

E-mail addresses: [email protected] (S. Lee), [email protected] (L.F. Kallivokas).

0168-874X/$ - see front matter ? 2004 Elsevier B.V. All rights reserved.doi:10.1016/j. nel.2004.02.005

Page 2: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2064 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

at all exists). Such an exact condition can be seen as a near- eld instantiation of the Sommerfeldradiation condition—realized closer to the origin rather than at in nity.Exact conditions on truncation boundaries are referred to as DtN (Dirichlet-to-Neumann) maps,

for they typically relate the Dirichlet to the Neumann datum on the boundary [1,2]. DtN conditionsare, by construction, spatially and temporally non-local, even for the few canonical geometries forwhich the apparatus of separation of variables allows for their construction. The non-locality of DtNmaps imposes onerous computational requirements. By contrast, and in order to overcome the exactcondition’s di:culties, local approximants in both time and space of the DtN maps oJer the bene tof reduced computational cost at the expense of accuracy. Oftentimes, but not always, the trade-oJis user-controllable, via: (a) the order of the local approximant (increasing orders lead to increasedspatial and temporal coupling); and (b) the location of the truncation boundary (the farther theboundary is placed, the fewer the reLections are).For the computational bene ts they aJord, it is without surprise that a signi cant amount of work

has been devoted to devising local constructs of the truncation condition. Extensive surveys, bothearly and recent, of developments in this area include the works in [2–6]. Local constructs can beroughly classi ed in two categories: those based on rational approximations of the dispersion relation,and those based on asymptotic expansions of the far- eld solution or the solution exterior to thecomputational domain. As examples, we mention here the pioneering works of Engquist and Majda[7,8], and Bayliss and Turkel [9–11]. Most of these and related works, however, were limited tocanonical geometries (straight, circular, or spherical boundaries). It is conceivable though that othergeometries may oJer computational advantages, especially in the presence of elongated scatterers.However, numerical results pertaining to non-circular or non-spherical geometries are scant.For example, in the context of two-dimensional elliptically shaped local boundaries we mention

ad hoc (as opposed to systematic) developments [12–17] largely based on generalizations of theBayliss–Turkel conditions to elliptical or general convex geometries without, however, the bene t ofgeneralization to higher-order conditions, and the systematic works of Grote and Keller [18], Antoineet al. [19], Barry et al. [20], and Kallivokas et al. [21–23]. Finite element implementations of up tosecond-order conditions have recently appeared for elliptical and arbitrarily shaped geometries [24,25]for frequency-domain applications only, whereas comparisons of elliptical boundary developmentsup to 1996 have also been reported [26,27].The works discussed thus far pertain mostly to the full-plane or full-space case. In [28], we showed

how a family of absorbing boundary conditions prescribed on arbitrarily-convex truncation surfaces(three-dimensional case) can be applied to the half-space problem; therein, numerical results werepresented for semi-spherical truncation boundaries endowed with second-order conditions. To thebest of our knowledge, local absorbing conditions applicable to both time- and frequency-domainproblems for half-plane problems prescribed on non-circular boundaries have not appeared. Thus,this paper lls this gap by extending past work [29,28], most notably the approach followed for thehalf-space three-dimensional case [28]; herein, we present numerical results for elliptically-shapedboundaries for the half-plane case in both the frequency- and time-domains using a local second-orderabsorbing boundary condition.The rest of the article is organized as follows: we treat rst the continuous problem starting with the

full-plane setting, and repeat the main theoretical results from recent work [30]; using the full-planedevelopment, we particularize next the conditions to the half-plane case and provide the frameworkfor resolving the waves in the semi-in nite medium using the method of images. In the numerical

Page 3: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2065

section we present results pertaining to both canonical and arbitrary scatterers using both ellipticallyshaped and circular truncation geometries for applications in the time- and frequency-domains andestablish the computational advantages elliptically shaped boundaries oJer. Lastly, in Appendix A,we provide an exact solution to a prototype problem involving cylindrical obstacles, which is usedto compare against the numerical results.

2. The continuous problem

2.1. Propagation in full-plane

We consider rst the exterior problem in a full-plane governed by the time-dependent waveequation. Let be a closed curve with exterior ⊂ R2; is occupied by a linear, inviscid, andcompressible Luid. Without loss of generality we focus on the scattering problem in which a travelingplane wave p0 impinges upon a rigid structure 1 (Fig. 1a), and seek to determine the scattered (ortotal) pressure eld within . 2 The strong form of the problem can be stated asGiven p0(x; t) with x∈, nd p(x; t) such that

Op(x; t) = c2Qp(x; t); x∈; t ¿ 0; (1)

p(x; t) + p0(x; t) = 0; x∈; t ¿ 0; (2)

limr→∞

√r(pr +

1cp) = 0; (3)

p(x; 0) = 0; p(x; 0) = 0; x∈ S; (4)

where p denotes scattered pressure, x is the position vector, t is time, is the outward unit normalon , c is the velocity of wave propagation, is the Laplace operator, an overdot denotes derivativewith respect to time, and p denotes the normal derivative of the scattered pressure p. Condition (2)implies that the normal Luid acceleration vanishes on the boundary of the scatterer, or equivalently,that the normal derivative of the total pressure also vanishes on . Condition (3), in which r isradial distance and pr the derivative of the pressure along the radial direction, is the Sommerfeldradiation condition. Condition (4) indicates that the system is taken to be initially at rest.Following classical lines for resolving unbounded domain problems, the exterior in nite region

is truncated via the introduction of an arti cial boundary a. The process gives rise to the nitecomputational domain f (Fig. 1b). For well-posedness of the associated IBVP over the nite f ,a boundary condition need also be prescribed on the truncation boundary a. This condition, whichcan be determined in terms of the actual solution p on a, admits the form [23]:

p(x; t) =F[pt(·; ·)](x); x∈a; (5)

in which the dots following pt indicate dummy variables, and F is an integral operator that dependson pt , the time history of p, i.e.,

pt(t) = p(t − );∀: 06 6 t: (6)

1 Throughout this article, the scatterer is considered to be immobile.2 Radiation problems can be similarly treated.

Page 4: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2066 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

Ω

RigidScatterer

RigidScatterer

p0

p0Incident Pressure Field

Γa

Γ Γ

AbsorbingBoundary

Ωf~

(a) (b)

Fig. 1. (a) Exterior scattering problem in unbounded full-plane; and (b) reduced model with nite Luid region f andarti cial boundary a.

F is non-local in time and space, that is to say, the motion at any given instant t at every point onthe arti cial boundary a is coupled with the time histories of all other points on a. The non-localcharacter of the exact F makes it unsuitable for implementation in the context of domain dis-cretization methods. In [21], we developed a family of localized approximants to F for arbitrarilyconvex truncation boundaries and have used its members in applications implicating circular trun-cation boundaries in a full-plane. Recently, in [30], we presented numerical results for ellipticallyshaped boundaries also in a full-plane. Here, we use a second-order approximant to F, denoted byF2, prescribed on an elliptical boundary for half-plane applications; this local second-order conditionis particularly well-suited for numerical implementation using nite elements. In the next section,we discuss the essential theoretical results from [30] that lead to the second-order approximant. Weremark that, with the introduction of the arti cial boundary and the approximant F2, the IBVP(1)–(4) is now replaced by

Op(x; t) = c2Qp(x; t); x∈f ; t ¿ 0; (7)

p(x; t) = −p0(x; t); x∈; t ¿ 0; (8)

p(x; t) =F2[p(x; t)]; x∈a; t ¿ 0; (9)

p(x; 0) = 0; p(x; 0) = 0; x∈ Sf : (10)

2.2. The absorbing boundary

Let a (the absorbing boundary) be smooth and convex and let it be described by the parametricrepresentation X( ), where X denotes the position vector on a, and is an arc-length parameter(Fig. 2). Using a Fermi-type coordinate system, it can be shown [30,20,21] that a second-orderabsorbing boundary condition on a is given by

p + p = −1cOp+

(12 −

c

)p+

12cp +

(182c +

12

)p; (11)

where is an arbitrary non-negative parameter, and denotes the curvature of a (for convex a,(x)¡ 0;∀x∈a).

Page 5: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2067

Γ

v

RigidScatterer

v

R(λ,ξ)

λ Arc-lengthparameterΩf

v

Γa

X(λ)

ξv

Fig. 2. Arti cial boundary a in a Fermi-type coordinate system characterized by the planar curve parameter and scalardistance metric .

The parameter in (11) has been formally introduced for stability in time-domain applications.Speci cally, in [30] we derived (11) (and all members of the family of approximants) using asymp-totic expansions, in the Laplace transform domain, of the solution to the problem exterior to a. Theparameter was introduced into these expansions to ensure the stability of the solution when theexpansions are transformed back to the time-domain. From a physical point of view, representsdamping introduced through the boundary a. In [20,21], it was shown that, to ensure the dissipa-tivity of the conditions, need be greater than a critical value above which exponential error growthin time-domain applications is prevented. In [30], for elliptically shaped boundaries, we used:

(x) = −c(x): (12)

With choice (12), the second-order condition (11) becomes

p − cp = −1cOp+

32p+

12cp − 3

82cp: (13)

It can be easily veri ed that either (11) or (13) cannot be readily incorporated into a weak formtantamount to the strong statement of the modi ed IBVP (7)–(10), since (11) or (13) contain boththe normal derivative p and its rst time derivative, i.e., p. To do this e:ciently, we introduce twoauxiliary variables on the arti cial boundary, q(1) and q(2), and decompose (13) into the followingequivalent set of three equations:

− p =1cp − 1

2p − c

2q(1) − c

82q(2); (14)

p − q(1) + cq(1) = 0; (15)

p − q(2) + cq(2) = 0: (16)

Eqs. (14)–(16) allow now the ready incorporation of the second-order condition into a variationalstatement for the IBVP (7)–(10), at the minimal expense of two additional variables on the absorbingboundary.

2.3. Propagation in a half-plane

We consider next the half-plane propagation problem, and in particular the scattering problemthat arises when a traveling wave pinc impinges upon a submerged near-surface object (Fig. 3).

Page 6: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2068 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

Free surface Γf

Half-Plane Ω Γ

v

RigidScattererp

inc

pincIncident Pressure Field

Γ

vΓa

AbsorbingBoundary

Ωf~Rigid

Scatterer

(a) (b)

Fig. 3. (a) Exterior scattering problem in a half-plane; and (b) reduced model with nite Luid region f , arti cial boundarya, and free surface f .

Γ

v

RealScatterer

Pinc

Γa

Γf

Prefl

v

xy

AbsorbingBoundry

Real IncidentPressure field

Image IncidentPressure field

(Free-field reflected)

Free surface

ΓImageScatterer

Fig. 4. Method of images: full-plane problem adjoint to the half-plane problem shown in Fig. 3b.

Typically, in full-plane problems, one need only be concerned with the motion within the nite Luiddomain: it is expected that the arti cial boundary will resolve any outgoing waves to the degreeallowed by the order of the approximant to the exact truncation condition (second order here). Bycontrast, in half-plane scattering problems where full-plane absorbing boundary conditions are used,the introduction of an arti cial boundary might give rise to additional errors, particularly in regionsclose to the intersection of the arti cial boundary with the free surface. This is especially true if onewere to solve for the scattered wave eld in a manner identical to that of the preceding section (thescattered wave eld on the free surface is non-zero). The peculiarity of the half-plane problem stemsfrom the presence of a free surface which the absorbing boundary conditions, developed originallyfor full-plane problems, do not readily account for.To overcome the di:culty imposed by our desire to use full-plane conditions in half-plane prob-

lems, we resort to the method of images in order to: (a) ensure the appropriateness of the useof the full-plane condition (13); and (b) ensure that the wave eld one needs to solve for remainszero at all times on the free surface. To this end, we consider the half-plane problem depicted inFig. 3(b) and its adjoint full-plane problem depicted in Fig. 4. In the latter gure, preL representsthe image incident eld of the incident pressure eld pinc in the positive half-plane (y¿ 0). In otherwords, in the negative half-plane preL represents the free- eld reLected wave eld, i.e., the wave eldgenerated when pinc impinges on the free-surface in the absence of any scatterers. We remark thatwe require that the free surface be an axis of symmetry of the convex planar region bounded by

Page 7: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2069

the (closed-curve) arti cial boundary a. Having turned our attention to the half-plane problem,henceforth a is used to denote the (open-curve) arti cial boundary in the negative half-plane only(Fig. 4).Let us de ne the image incident pressure eld as

preL(x; y; t) = −pinc(x;−y; t); ∀y: (17)

The total pressure eld is

ptot(x; y; t) = p0(x; y; t) + p(x; y; t); ∀y; (18)

where p denotes part of the scattered eld, and

p0(x; y; t) = pinc(x; y; t) + preL(x; y; t); ∀y: (19)

On account of symmetry,

ptot(x; y; t) = −ptot(x;−y; t); ∀y: (20)

Therefore, on the free surface f (y = 0), one has

ptot(x; 0; t) = 0; (21)

and from (17)

pinc(x; 0; t) + preL(x; 0; t) = 0: (22)

Clearly, from (19) and (22), there holds on the free-surface f (y = 0):

p0(x; 0; t) = pinc(x; 0; t) + preL(x; 0; t) = 0: (23)

Thus, p0 above denotes the total free- eld wave eld that would have been generated in the absenceof any scatterers. p in (18) denotes a scattered pressure wave eld; in physical terms, p is equal tothe sum of the scattered elds generated by the real and image scatterers when they are insoni edby pinc and preL, respectively. With these de nitions, p is equal to the part of the scattered eldunderstood as the total wave eld less the free- eld wave p0. Notice that by virtue of (18), (21),and (23), there also holds on f

p(x; 0; t) = ptot(x; 0; t) − p0(x; 0; t) = 0: (24)

We remark that the total scattered wave eld psc generated in the negative half-plane when pinc

impinges on the scatterer is given by

psc(x; y; t) = p(x; y; t) + preL(x; y; t);∀y6 0: (25)

It can be easily seen now that the adjoint problems (Figs. 3(b) and 4) are equivalent, provided thatin the half-plane case one solves for the scattered wave eld p, prescribes (24) on the free surface,and uses p0 (19) as the excitation. These observations, together with the fact that the right-hand sideof the absorbing boundary condition (13) is an odd function with respect to the y coordinate, allowfor the ready use of the full-plane condition (13) in the half-plane problem without any modi cation,i.e., one needs to pose the problem for p only over the lower half-plane in Fig. 4, while prescribing(13) on the restriction of the full-plane arti cial boundary in the half-plane. The associated strongform can be cast as

Page 8: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2070 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

Given an incident eld pinc(x; t) with x∈f (Fig. 4), nd p(x; t) such that

Op(x; t) = c2Qp(x; t); x∈f ; t ¿ 0; (26)

p(x; t) = −p0(x; t) = −9pinc(x; t)

9 − 9preL(x; t)9 ; x∈; t ¿ 0; (27)

− p(x; t) =1cp(x; t) − 1

2(x)p(x; t) − c

2q(1) (x; t)

−c82(x)q(2)(x; t); x∈a; t ¿ 0; (28)

p (x; t) − q(1) (x; t) + c(x)q(1) (x; t) = 0; x∈a; t ¿ 0; (29)

p(x; t) − q(2)(x; t) + c(x)q(2)(x; t) = 0; x∈a; t ¿ 0; (30)

p(x; 0) = 0; p(x; 0) = 0; x∈ Sf : (31)

3. The discrete problem

3.1. Finite elements

Following classical lines the weak form corresponding to (26)–(31) can be easily shown to be

1c2

∫f

p Op df +∫f

∇p · ∇p df +1c

∫a

pp da − 12

∫a

pp da

+c2

∫a

p q(1) da − c

8

∫a

2pq(2) da

+c2

∫a

q(1) p da − c2

∫a

q(1) q(1) da +c2

2

∫a

q(1) q(1) da

− c8

∫a

2q(2)pa +c8

∫a

2q(2) q(2) da − c8

∫a

3q(2) q(2) da

=∫pp0

da: (32)

The semi-discretized equations corresponding to the variational form (32) can be obtained read-ily by using standard piecewise polynomials to represent the test and trial functions implicated in(32) (details were given in [30]). The resulting ordinary diJerential equations have the followingsecond-order structure:

M OU + CU + KU = F ; (33)

Page 9: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2071

Standard acoustic element(e.g.4-noded quadrilateral)

Absorbing finite element(e.g.2-noded line element)

Absorbing Boundary Γa

Finite Computational Domain Ωf

Fig. 5. Representation of the absorbing boundary element.

where

M =

Mpp Mppf

0

Mpf p Mpf pfMpf pa

0 MpapfMpapa

; (34)

C =

0 0 0

0 0 0

0 0 C a

; K =

Kpp Kppf

0

Kpf p Kpf pfKpf pa

0 KpapfKpapa

+ K a

: (35)

U is the vector of the unknown nodal quantities, i.e.

UT = [pT; pTf; SpTa

] (36)

and

FT = [f Tp; 0T; 0T]: (37)

In the above, SpTa= [pTa

; q(1)Ta; q(2)Ta

], fp is the discretized form of the right-hand side of (32), andpT, p

Tf, pTa

denote partitions of p over , f , and a, respectively. Notice that the contributions ofthe absorbing boundary enter through the bottom right blocks in the damping and stiJness matrices(35).For completeness, we repeat the expressions given in [30,21], that characterize the contributions

of the absorbing boundary. As it can be seen from the weak form (32) or the global matrices(35), the contributions of the absorbing boundary give rise to a boundary-only element (Fig. 5) thatis completely characterized by the damping and stiJness matrices C a and K a; the corresponding

Page 10: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2072 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

element matrices are given as

ka =

−12

∫e T c

2

∫e T

−c8

∫e2 T

c2

∫e T

c2

2

∫e T

0

−c8

∫e2 T 0 −c2

8

∫e3 T

; (38)

ca =

1c

∫e T 0 0

0 −c2

∫e T

0

0 0c8

∫e2 T

: (39)

and above are the shape functions and their tangential derivatives, respectively (details onthe shape functions are provided in Appendix B). The subscript e denotes a line element on thetruncation boundary a, and the line diJerential (de

a) has been dropped throughout for brevity. Weremark that one need only mesh the nite region f and simply attach the absorbing element onthe boundary a without any further discretization within the in nite exterior region. Notice that allmatrices are frequency-independent and symmetric, thus readily allowing for applications in eitherthe time- or frequency-domain, while maintaining the overall symmetry of the algebraic systemsresulting from the discretization of the interior computational domain.

4. Numerical results

Two types of rigid scatterers submerged in a half-plane at diJerent depths of immersion areconsidered; a circular cylindrical, and a cigar-shaped scatterer. We consider insoni cation by bothtime-harmonic and transient plane waves. Numerical results are obtained using the second-orderabsorbing boundary condition prescribed on both semi-circularly- and semi-elliptically-shapedboundaries. All subsequent references to scattered pressure pertain to the partial scattered eld ob-tained as a solution to (33); to recover the total scattered eld one need only add preL (see A.2).

4.1. Cylindrical scatterer

The geometry of the problem is depicted in Fig. 6; we denote the major semi-axis of thesemi-elliptical boundary with sM, the minor semi-axis with sm, and their ratio with = sM=sm.The cylindrical scatterer has a radius of a. In all cases, the mesh density is kept the same; it ischaracterized by a typical element size metric of 0:02a. For the time-harmonic cases we use a fre-quency sweep between ka=0:1 and ka=10 (low to medium range). With these choices the smallestwavelength is 0:6a; Therefore, for the highest frequency we consider, the mesh density results inroughly 30 points per wavelength—more than enough to resolve locally the waves. We use linear

Page 11: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2073

Pinc

α

θ

y

xsm

sM

d/2

a

Fig. 6. Model of a rigid circular scatterer in a half-plane surrounded by a semi-elliptical absorbing boundary.

Table 1Computational cost in terms of the total number of degrees-of-freedom per case for the circular cylindrical scatterer

sM Total DOF d=2a

= 1 = 2

4:0a 57,850 23,554 2.03:0a 28,658 10,400 1.23:0a 28,058 10,248 1.5

isoparametric elements for both domain Luid elements and for the absorbing boundary elements; forall numerical experiments we used unstructured quadrilateral meshes.We consider three diJerent depths of immersion, three angles of incidence (=90 (vertically prop-

agating wave), 45, and a grazing angle of 5), and both semi-circular and semi-elliptical boundaries.Table 1 summarizes for all cases the associated computational cost in terms of degrees-of-freedom.Figs. 7–9 depict the amplitude of the scattered pressure eld as computed at various points on

the cylindrical scatterer’s boundary (denoted in the plots by the polar angle ). Both exact andapproximate solutions are shown; the latter correspond to a semi-circular boundary ( = 1), anda semi-elliptical boundary ( = 2). Clearly, there is excellent agreement between the exact andcircular-boundary cases; the agreement between the exact and the elliptical boundary is also quitesatisfactory in all cases. We observe the greatest deviation from the exact solution in the case of theelliptical boundary (=2) when the wave impinges at the grazing angle (Fig. 9). We remark though,that the use of the semi-elliptical geometries has allowed for considerable computational savings overthe semi-circular boundaries; referring to Table 1, the savings, again in terms of degrees-of-freedom,are of the order of 60%. We further remark that the accuracy of both boundaries at the low-frequencyend is also quite satisfactory.Whereas Figs. 7–9 depict the response point-wise, in Fig. 10 we compare the numerical solution

against the exact for the distribution of the scattered pressure amplitude along both the scatterer’sboundary (), and the absorbing boundary (a) for a single scattering case (angle of incidence= 45, ka= 1). Notice again that the agreement is quite satisfactory.

Page 12: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2074 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

0 1 2 3 4 5 6 7 8 9 10

ka0 1 2 3 4 5 6 7 8 9 10

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Scat

tere

dP

ress

ure

Am

plitu

de

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Exactη = 1 η = 2

θ = 270°α = 90°

θ = 270°α = 45°

θ = 0°α = 45°

θ = 0°α = 90°

θ = 90°α = 45°

θ = 90°α = 90°

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

Fig. 7. Scattered pressure amplitude for a circular cylindrical scatterer; depth of immersion d=2a= 2:0; sM = 4:0a.

Fig. 11 depicts a relative error measure E on the scatterer’s boundary using L2-norms for variousratios and for a single immersion depth d=2a = 2:0 and angle of incidence = 45. The error isde ned as

E =[∫ ‖pex − papp‖2 d]1=2[∫ ‖pex‖2 d]1=2

× 100%; (40)

where pex is the exact solution derived from (A.16), and papp is the nite element solution obtainedby solving (33). In all cases, it appears that the circular boundary outperforms the elliptical one at an

Page 13: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2075

0 1 2 3 4 5 6 7 8 9 10

ka0 1 2 3 4 5 6 7 8 9 10

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Scat

tere

dP

ress

ure

Am

plitu

de

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Exactη = 1

η = 2

0.0

0.5

1.0

1.5

2.0

2.5

3.0

3.5

4.0

4.5

θ = 270°α = 90°

θ = 270°α = 45°

θ = 0°α = 45°

θ = 90°α = 45°

θ = 90°α = 90°

θ = 0°α = 90°

Fig. 8. Scattered pressure amplitude for a circular cylindrical scatterer; depth of immersion d=2a= 1:2; sM = 3:0a.

increased computational cost. We remark however that the error levels for the elliptical boundaries(e.g. = 1:5), are still quite satisfactory.

4.2. Cigar-shaped scatterer

To illustrate the applicability of the boundaries and the corresponding absorbing nite elementsto problems of practical interest, we consider next an elongated cigar-shaped scatterer (Fig. 12),

Page 14: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2076 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

0 1 2 3 4 5 6 7 8 9 10ka

0 1 2 3 4 5 6 7 8 9 10

Scat

tere

dPr

essu

reA

mpl

itud

e

0.00.20.40.60.81.01.21.41.61.82.0

Exactη = 1η = 2

0.00.20.40.60.81.01.21.41.61.82.02.2

θ = 0°α = 5°

θ = 180°α = 5°

θ = 270°α = 5°

θ = 90°α = 5°

Fig. 9. Scattered pressure amplitude for a circular cylindrical scatterer; depth of immersion d=2a= 1:5; sM = 3:0a.

for which the semi-elliptical boundary oJers tighter circumscription, and therefore the potential forincreased computational gains.Since there is no exact solution in this case, we obtain a reference solution to compare against

by using a semi-circular absorbing boundary with a radius of 22a endowed with the second-ordercondition. Table 2 summarizes the number of degrees-of-freedom used for each of the four casesfor which we computed the response. The typical mesh element metric is 0:25a. Fig. 13 depictsthe distribution of the real and imaginary parts of the scattered eld on the cigar-shaped scatterer’sboundary for two frequencies (ka = 1 and ka = 3) and for a single incidence angle of 45. Theagreement between the exact and all approximate solutions is excellent. Notice though (Table 2)that, for comparable accuracy, the computational cost associated with = 2 is less than half that ofthe semi-circular boundary.To illustrate the applicability of the absorbing elements to transient simulations, Fig. 14 pertains to

a direct time-domain solution of (33) for the scattered pressure eld due to the insoni cation of thecigar-shaped scatterer by a traveling plane wave. We use the implicit trapezoidal rule to integratein time the semi-discrete form (33) with a time-step Qt c=a = 0:1; the time signal for the planewave is given by a nite-duration modi ed Ricker pulse with a central dimensionless frequency of 1(see [30] for details). Again, as seen from Fig. 14 the agreement between the reference solution andthe various elliptical boundaries at discrete points on the surface is excellent.

Page 15: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2077

-180 -150 -120 -90 -60 -30 30 60 90 120 150 1800.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

1.6

1.8

Exact

Scat

tere

d Pr

essu

re A

mpl

itude

θ

η = 2.0On Scatterer

On Absorbing Boundary

0

Fig. 10. Distribution of scattered pressure amplitude for a circular cylindrical scatterer; depth of immersion d=2a = 1:2;sM = 3:0a; = 45, ka= 1.

ka

Rel

ativ

e L

2 (Γ

) E

rror

E (

%)

00 1 2 3 4 5 6 7 8 9 10

1

2

3

4

5

6

7η = 1.00η = 1.50η = 2.00

Fig. 11. Relative L2() error E (in percent) for elliptical and circular boundaries; depth of immersion d=2a= 2:0.

5. Conclusions

In this paper we discussed an e:cient methodology for resolving scattered or total pressure eldsarising in the numerical modeling of near-surface scatterers submerged in an acoustic half-plane. Theprimary objectives of this article were: (a) to provide the theoretical basis for extending full-planeconditions to the half-plane using the method of images; and (b) to demonstrate the signi cant

Page 16: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2078 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

1.5a

4.5a 0.5a

a

Point 3

Point 4Point 2 Point 1

pinc

α

η = 1.5

η = 1.0

η = 2.0

SM = 9a

Fig. 12. Model of a rigid cigar-shaped scatterer in a half-plane surrounded by various elliptical absorbing boundaries.

Table 2Computational cost in terms of the total number of degrees-of-freedom per case for the cigar-shaped scatterer

Semi-axes ratio Total DOF= sM=sm

2.0 23 7981.5 32 9141.0 52 174Reference 321 008

computational gains that are attainable, at no or minimal loss of accuracy, when elliptically-shapedboundaries are adopted over their circular counterparts for the truncation of the semi-in nite extentof the physical domain. Borrowing from full-plane developments of earlier works, we have shownthat the ease by which second-order conditions can be incorporated into existing nite element codesextends to the half-plane case as well.

Acknowledgements

We wish to thank the reviewers for their helpful suggestions.

Appendix A. The exact solution

In order to compare the numerical results obtained using the outlined approach against exactsolutions to canonical problems, we provide next the solution to the scattering problem arising whena time-harmonic plane wave insoni es a near-surface rigid circular cylindrical scatterer. While theexact solution to this problem appears not to be readily available in the literature, a closely relatedsolution that follows similar lines, albeit for the more complex case of an elastic cylinder, can befound in [31].

Page 17: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2079

ka = 1 ka = 3

ka = 1 ka = 3

-1.5-180 -150 -120 -90 -60 -30 0 30 60 90 120 150 180

-1.0

-0.5

0.0

0.5

1.0

-1.5

-1.0

-0.5

0.0

0.5

1.0

1.5

Reference

-5

-4

-3

-2

-1

0

1

2

3

4

-5

-4

-3

-2

-1

0

1

2

3

4

5

Re

[Sca

ttere

d Pr

essu

re]

Im [

Scat

tere

d Pr

essu

re]

Re

[Sca

ttere

d Pr

essu

re]

Im [

Scat

tere

d Pr

essu

re]

η = 1.0η = 1.5η = 2.0

θ-180 -150 -120 -90 -60 -30 0 30 60 90 120 150 180

θ

Fig. 13. Real and imaginary parts of the scattered pressure on the cigar-shaped scatterer (Fig. 12) for ka= 1 and ka= 3; = 45.

At Point 3 At Point 4

At Point 2At Point 1

-2

-1

1

2

-30 10 20 30 40 50 60 70 80 90 100 10 20 30 40 50 60 70 80 90 100

0

-2

-1

1

2

-3

0

3

Reference

Scat

tere

d Pr

essu

re

Time tc/a

η = 1.0η = 1.5η = 2.0

Fig. 14. Scattered pressure trace at various points on the surface of the cigar-shaped scatterer (Fig. 12); transient responseto a traveling plane wave; various boundaries; = 45.

Page 18: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2080 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

d/2

d/2

r′

r

Free surfacexy

Real cylinder

Image cylinder

αp

inc

prefl

α′

a

x

y

y′

θ′

x′

O

O′

Fig. 15. Geometry of a rigid circular scatterer in a half-plane; real and image scatterers.

Accordingly, we resort again to the method of images in conjunction with an addition theoremfor cylindrical wave functions. We consider the geometry shown in Fig. 15. The image and realscatterers are separated by a distance d between their centers O′ and O, respectively.We consider an incident time-harmonic plane wave pinc impinging at an angle of incidence with

respect to the positive x-axis.

pinc(x; y; t) = eik(x cos +y sin )e−i!t: (A.1)

Similarly, the image incident, or free- eld reLected wave, preL, is

preL(x; y; t) = −eik(x cos −y sin )e−i!t; (A.2)

where k denotes wavenumber (k=!=c). Using polar coordinates (r; ) for the system that has originat O (Fig. 15), and the following transformations:

x = Sx = r cos ; y = Sy − d2= r sin − d

2; (A.3)

the real and image incident waves (A.1) and (A.2), respectively, can be written as (we drop thee−i!t dependence from all subsequent expressions):

pinc(r; ) = e−ik(d=2) sin eikr cos(−); (A.4)

preL(r; ) = −eik (d=2) sin eikr cos(+): (A.5)

Page 19: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2081

Using a Jacobi–Auger expansion [32], Eqs. (A.4)–(A.5) can be further re-written in terms of seriesof cylindrical Bessel functions, as

pinc(r; ) = e−ik(d=2) sin eikr cos (−)

=e−ik(d=2) sin ∞∑

n=−∞inJn(kr) eine−in; (A.6)

preL(r; ) =−eik(d=2) sin eikr cos (+)

= − eik(d=2) sin ∞∑

n=−∞inJn(kr) einein: (A.7)

We next de ne two additional wave elds, namely, the scattered eld due to the real cylinder prealsc

(if the real cylinder alone were occupying the full-plane and were insoni ed by pinc), and thescattered eld due to the image cylinder pimage

sc (again, if the image cylinder alone were occupyingthe full-plane and were insoni ed by preL). Accordingly

prealsc (r; ) =

∞∑n=−∞

anH (1)n (kr)ein; (A.8)

pimagesc (r′; ′) =

∞∑n=−∞

bnH (1)n (kr′)ein

′: (A.9)

Both scattered elds above have been obtained as solutions to the Helmholtz equation written in thepolar coordinate systems with origins at O and O′, respectively. Clearly, the sum of all four eldsis equal to the total pressure eld for the half-plane problem, i.e.,

ptot = pinc + preL + prealsc + pimage

sc : (A.10)

Since on the free surface f , the total eld (A.10) must vanish (see (21)), and, by construction, thesum of the real and image incident elds also vanishes on f (see (23)), there results

prealsc (r; ) + pimage

sc (r′; ′) = 0 on f : (A.11)

Using r′ ≡ r and ′ ≡ 2( − on f , (A.11) becomes

prealsc (r; ) + pimage

sc (r; 2( − )

=∞∑

n=−∞anH (1)

n (kr)ein +∞∑

n=−∞bnH (1)

n (kr)ein(2(−)

=∞∑

n=−∞anH (1)

n (kr)ein +−∞∑m=∞

b−mH(1)−m(kr)e

im

=∞∑

n=−∞anH (1)

n (kr)ein +∞∑

m=−∞b−mem(iH (1)

m (kr)eim

Page 20: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2082 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

=∞∑

n=−∞anH (1)

n (kr)ein +∞∑

n=−∞b−n(−1)nH (1)

n (kr)ein

=∞∑

n=−∞[an + (−1)nb−n]H (1)

n (kr)ein = 0: (A.12)

Therefore,

b−n = (−1)n+1an: (A.13)

Next, we use an addition theorem for cylindrical waves [32] to transform pimagesc (A.9) to the same

coordinate system (origin at O) as prealsc (A.8). Accordingly

H (1)n (kr′)ein

′=

∞∑m=−∞

H (1)n−m(kd) Jm(kr) e

i[(n−m)3(=2+m]; r6d

Jm(kd)H(1)n−m(kr) e

i[m3(=2+(n−m)]; r¿d

: (A.14)

Using (A.13), (A.14), after manipulation of the series indices, becomes:

pimagesc (r; ) =

∞∑n=−∞

(−1)n+1 in

Jn(kr)

H (1)n (kr)

ein[ ∞∑m=−∞

amimH (1)m+n(kd)

Jm+n(kd)

];r6d

r¿d

: (A.15)

Substituting (A.6)–(A.8) and (A.15) into (A.10), the total pressure ptot becomes

ptot(r; ) =pinc(r; ) + preL(r; ) + prealsc (r; ) + pimage

sc (r; )

=∞∑

n=−∞

[e−ik(d=2) sin ine−in Jn(kr)

−eik (d=2) sin inein Jn(kr) + anH (1)n (kr) + (−1)n+1 in

×Jn(kr)

H (1)n (kr)

ein[ ∞∑m=−∞

amimH (1)m+n(kd)

Jm+n(kd)

]];r6d

r¿d:(A.16)

Eq. (A.16) provides the solution for the total pressure anywhere within the half-plane (y6 0). Allthat remains is to compute the unknown series coe:cients an. To this end, we consider the boundarycondition on the surface of the sound-hard real cylinder, i.e., we set 9ptot=9r|r=a = 0. There results:

−H′(1)n (ka)inJ ′

n(ka)an + (−1)n

∞∑m=−∞

amimH(1)m+n(kd)

= − 2i sin(kd2sin + n

); ∀n: (A.17)

Eq. (A.17) represents an in nite system of algebraic equations for the determination of the unknowncoe:cients an involved in the series (A.16). Once the coe:cients are determined, the total pressure

Page 21: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084 2083

is given by (A.16), and the total scattered eld by

psc(r; ) = ptot(r; ) − pinc(r; ): (A.18)

Appendix B. Implementational details

In implementing the discrete problem discussed in Section 3 we used standard isoparametricelements. We discuss the details pertaining to the various integrals of the absorbing element matricesin (38) and (39) using, without loss of generality, linear approximations. Accordingly, for anyelement e of the discretized arti cial boundary, we used the standard map:

x = T()xe; y = T()ye; (B.1)

where xe and ye denote the nodal coordinates of element e in a global Cartesian coordinate system(x; y), and denotes the local coordinate, with ∈ (−1; 1). For a linear approximation:

() = 12

[1 −

1 +

]; () =

12

[ −1

1

]: (B.2)

Then,

d e =√x2 + y2 d=

√( T

xe)2 + ( Tye)2 d= j d; (B.3)

() =1j

−1

212

: (B.4)

Finally, in all expressions where the curvature appears, we used its exact value, which for anellipse with semi-axes sM and sm reduces to:

(x; y) = −sMsm=(s2m=s2Mx2 + s2M=s2my

2)3=2: (B.5)

References

[1] J.B. Keller, D. Givoli, Exact non-reLecting boundary conditions, J. Comput. Phys. 82 (1989) 172–192.[2] D. Givoli, Non-reLecting boundary conditions: a review, J. Comput. Phys. 94 (1) (1991) 1–29.[3] E. Kausel, J.L. Tassoulas, Transmitting boundaries: a closed-form comparison, Bull. Seismol. Soc. Amer. 71 (1981)

143–159.[4] J.P. Wolf, A comparison of time-domain transmitting boundaries, Earthquake Eng. Struc. Dyn. 14 (1986) 655–673.[5] E. Kausel, Local transmitting boundaries, ASCE J. Eng. Mech. 114 (6) (1988) 1011–1027.[6] S.V. Tsynkov, Numerical solution of problems on unbounded domains. A review, Appl. Numer. Math. 27 (1998)

465–532.[7] B. Engquist, A. Majda, Absorbing boundary conditions for the numerical simulation of waves, Math. Comp. 31

(1977) 629–651.[8] B. Engquist, A. Majda, Radiation boundary conditions for acoustic and elastic wave calculations, Comm. Pure Appl.

Math. 32 (1979) 313–357.[9] A. Bayliss, E. Turkel, Radiation boundary conditions for wave-like equations, Comm. Pure Appl. Math. 33 (1980)

707–725.

Page 22: Local absorbing boundaries of elliptical shape for scalar wave propagation in a half-plane

2084 S. Lee, L.F. Kallivokas / Finite Elements in Analysis and Design 40 (2004) 2063–2084

[10] A. Bayliss, E. Turkel, Far eld boundary conditions for compressible Lows, J. Comput. Phys. 48 (1982) 182–199.[11] A. Bayliss, M. Gunzburger, E. Turkel, Boundary conditions for the numerical solution of elliptic equations in exterior

regions, SIAM J. Appl. Math. 42 (2) (1982) 430–451.[12] G.A. Kriegsmann, A. TaLove, K.R. Umashankar, A new formulation of electromagnetic wave scattering using an

on-surface radiation boundary condition approach, IEEE Trans. Antennas Propagation AP-35 (2) (1987) 153–160.[13] D.S. Jones, An approximate boundary condition in acoustics, J. Sound Vibration 121 (1) (1988) 37–45.[14] A. Khebir, O. Ramahi, R. Mittra, An e:cient partial diJerential equation technique for solving the problem of

scattering by objects of arbitrary shape, Microwave Opt. Tech. Lett. 2 (1989) 229–233.[15] C.F. Lee, R.T. Shin, J.A. Kong, B.J. McCartin, Absorbing boundary conditions on circular and elliptical boundaries,

J. Electromagn. Waves Appl. 4 (10) (1990) 945–962.[16] Y. Li, Z.J. Cendes, Modal expansion absorbing boundary conditions for two-dimensional electromagnetic scattering,

IEEE Trans. Magn. 29 (1993) 1835–1838.[17] B. Stupfel, Absorbing boundary conditions on arbitrary boundaries for the scalar and wave equations, IEEE Trans.

Antennas Propagation 42 (6) (1994) 773–780.[18] M.J. Grote, J.B. Keller, On nonreLecting boundary conditions, J. Comput. Phys. 122 (1995) 231–243.[19] X. Antoine, H. Barucq, A. Bendali, Bayliss-Turkel-like radiation conditions on surfaces of arbitrary shape, J. Math.

Anal. Appl. 229 (1999) 184–211.[20] A. Barry, J. Bielak, R.C. MacCamy, On absorbing boundary conditions for wave propagation, J. Comput. Phys. 79

(1988) 449–468.[21] L.F. Kallivokas, J. Bielak, R.C. MacCamy, Symmetric local absorbing boundaries in time and space, ASCE J. Eng.

Mech. 117 (9) (1991) 2027–2048.[22] L.F. Kallivokas, J. Bielak, R.C. MacCamy, Absorbing boundary conditions of arbitrary shape for the

three-dimensional wave equation, in: IUTAM Symposium on Computational Methods for Unbounded Domains,Kluwer Academic Publishers, Boulder, CO, July 1997, pp. 221–228.

[23] L.F. Kallivokas, J. Bielak, A simple impedance-in nite element for the nite element solution of the three-dimensionalwave equation in unbounded domains, Comput. Methods Appl. Mech. Engng. 147 (1997) 235–262.

[24] R. Djellouli, C. Farhat, A. Macedo, R. Tezaur, Finite element solution of two-dimensional acoustic scattering problemsusing arbitrarily shaped convex arti cial boundaries, J. Comput. Acoust. 8 (1) (2000) 81–99.

[25] S.K. Bhattacharyya, S. Sathyapal, C.P. Vendhan, Absorbing boundary condition on elliptic boundary for the niteelement analysis of water wave diJraction by large elongated bodies, Internat. J. Numer. Meth. Fluids 37 (2001)249–277.

[26] D.B. Meade, G.W. Slade, A.F. Peterson, K.J. Webb, Comparison of local radiation boundary conditions for thescalar Helmholtz equation with general boundary shape, IEEE Trans. Antennas and Propagation 43 (1995) 1–5.

[27] B. Lichtenberg, K.J. Webb, D.B. Meade, A.F. Peterson, Comparison of two-dimensional conformal local radiationboundary conditions, Electromagnetics 16 (1996) 359–384.

[28] L.F. Kallivokas, A. Tsikas, J. Bielak, On transient three-dimensional absorbing boundary conditions for the modelingof acoustic scattering from near-surface obstacles, J. Comput. Acoust. 5 (1) (1997) 117–136.

[29] J. Bielak, L.F. Kallivokas, J. Xu, R. Monopoli, Finite element absorbing boundary for the wave equation in ahalf-plane with an application to engineering seismology, Proceedings of the Third International Conference onMathematical and Numerical Aspects of Wave Propagation (INRIA-SIAM), 489–498, Juan-les-Pins, France, April1995.

[30] L.F. Kallivokas, S. Lee, Local high-order absorbing boundaries of elliptical shape for scalar waves, Comput. Meth.Appl. Mech. Eng. (2004), accepted.

[31] V.W. Lee, M.D. Trifunac, Response of tunnels to incident SH-waves, J. Eng. Mech. Div. ASCE 105 (1979)643–659.

[32] A. Ivanov Ye, DiJraction of electromagnetic waves on two bodies, NASA translation of Difraktsiyaelektromagnitnykh voln na dvukh telakh, Nauka i Tekhnika Press, Minsk, 1968.