274
Views
2
CrossRef citations to date
0
Altmetric
Articles

Numerical reconstruction of an inhomogeneity in an elliptic equation

&
Pages 184-198 | Received 05 Jul 2013, Accepted 05 Jul 2013, Published online: 06 Aug 2013

Abstract

In this paper, the inverse problem which consists of reconstructing an unknown inner boundary of a domain from a single pair of boundary Cauchy data associated to an elliptic equation is solved numerically using the meshless method of fundamental solutions. A nonlinear minimization of the objective function is regularized. The stability of the numerical results is investigated for several test examples with respect to noise in the input data and various values of the regularization parameters.

Introduction

The determination of an inhomogeneity (anomaly) contained in a given domain from the knowledge of the imposed voltage (boundary temperature) and the measured current (heat) flux arises in many non-destructive tomography testing of materials. In particular, in this paper, we consider the inverse problem of determining an inhomogeneity Ω2 (with Lipschitz boundary Ω2) compactly contained in a bounded domain Ω (with smooth boundary Ω) entering in the modified Helmholtz elliptic equation(1) 2uk2u=0inΩ,(1) where k2=k12+(k22k12)χΩ2 is a given positive function with k2>k10 and of class C2(Ω), χΩ2 is the characteristic function of the domain Ω2, and u is the potential (temperature). In heat transfer, Equation (Equation1) stands as the governing equation for a (two-dimensional) dry fin-tube heat exchanger with k2=2hκδ, where κ is the thermal conductivity of the fin, h is the convective heat transfer coefficient and δ is the fin thickness, see [Citation1]. By defining(2) u:={u1inΩ1:=ΩΩ2,u2inΩ2,(2) Equation (Equation1) can be rewritten as the following transmission problem:(3) 2u1k12u1=0inΩ1,(3) (4) 2u2k22u2=0inΩ2,(4) (5) u1=u2,onΩ2.(5) (6) u1n=u2nonΩ2(6) where n¯ denotes the outward unit normal to the boundary.

Associated to the above problem (Equation3)–(Equation6), we also have the Cauchy boundary conditions on Ω given by(7) u1=fonΩ,(7) (8) u1n=gonΩ.(8) Such a problem arises in the determination of the contact resistivity of planar electronic devices.[Citation2] Other formulations of inverse obstacle problems can be found in the topical review by Isakov.[Citation3] Uniqueness of the inhomogeneity Ω2 entering the inverse problem (Equation3)–(Equation8) has been established in the class of balls, star-shaped domains, convex hulls of polygons and other classes of subdomains in [Citation4Citation7], respectively.

Recently, in [Citation8, Citation9] the authors determined inner rigid inclusions, cavities and absorbing obstacles in modified Helmholtz inverse geometric problems using the method of fundamental solutions (MFS). This method has proved rather versatile and easy to use for solving a wide range of inverse problems, see the recent review of Karageorghis et al. [Citation10]. In this study, we provide yet another application of the MFS for solving the inverse transmission problem (Equation3)–(Equation8).

The outline of this paper is as follows. In Section 2, we present the MFS for the modified Helmholtz equation in composite bi-materials. In Section 3, we present and discuss the numerically obtained results. In Section 4, we give some conclusions and possible future work.

Fig. 1 Sketch of the curves on which the source (--) and the boundary collocation () points are located in the MFS.

Fig. 1 Sketch of the curves on which the source (--−) and the boundary collocation (—) points are located in the MFS.

Fig. 2 The unregularized objective function for p{0,1,5,10}% noise, as a function of the number of iterations, for example 1.

Fig. 2 The unregularized objective function for p∈{0,1,5,10}% noise, as a function of the number of iterations, for example 1.

The method of fundamental solutions (MFS)

In the MFS for a composite bi-material Ω=Ω1Ω2, Ω1Ω2=, we approximate the solutions u1 and u2 of the modified Helmholtz Equations (Equation3) and (Equation4) by a linear combination of fundamental solutions in the form [Citation11],(9) u1,2N(X¯)=j=12NajGHM(X¯,ξ¯1j;k1),X¯Ω¯1,(9) (10) u2,N(X¯)=j=1NbjGMH(X¯,ξ¯2j;k2),X¯Ω¯2,(10) where GMH is the fundamental solution for the modified Helmholtz equations which in two dimensions, for example, is given by,(11) GMH(X¯,ξ¯;ki)=K0(kiX¯ξ¯),i=1,2,(11) where K0 is the modified Bessel function of the second kind of order zero. For simplicity, the constant 12π, which does not appear in (Equation11), has been embedded in the unknown coefficients (bj)j=1,N¯ in (Equation10). The modified Bessel function K0 is computed using the NAG routine S18ACF. In the case that the background medium Ω1 is harmonic, i.e. k1=0, we should replace GMH(X¯,ξ¯1;k1) in (Equation9) by the fundamental solution for the Laplace equation(12) 2u1=0inΩ1(12) which in two dimensions is given by(13) GL(X¯,ξ¯)=12πlnX¯ξ¯.(13) In this case, the MFS approximation for (Equation12) is(14) u1,2N(X¯)=j=12NajGL(X¯,ξ¯1j),X¯Ω¯1,(14) The source points (ξ¯1j)j=1,2N¯ in (Equation9) or (Equation14) are located outside the domain Ω¯ and inside the domain Ω2, whilst the source points (ξ¯2j)j=1,N¯ in (Equation10) are located outside Ω¯2. More precisely, (ξ¯1j)j=N+1,2N¯Ω2 and (ξ¯2j)j=1,N¯Ω2 are placed on (moving) pseudo-boundaries Ω2 and Ω2 similar to Ω2 at a distance δ>0 inwards and outwards, respectively. The rest of source points (ξ¯1j)j=1,N¯R2Ω¯ are placed on a (fixed) pseudo-boundary Ω similar to Ω. A sketch of the fictitions curves Ω2, Ω2 and Ω on which the source points are located is shown in Figure .

Fig. 3 The reconstructed inner boundary with no regularization for p{0,1,5,10}% noise, for Example 1.

Fig. 3 The reconstructed inner boundary with no regularization for p∈{0,1,5,10}% noise, for Example 1.

Fig. 4 The regularized objective function with various regularization parameters λ1=0,λ2{108,105,103,101} for p=10% noise, as a function of the number of iterations, for Example 1.

Fig. 4 The regularized objective function with various regularization parameters λ1=0,λ2∈{10−8,10−5,10−3,10−1} for p=10% noise, as a function of the number of iterations, for Example 1.

For simplicity, we assume that Ω is the unit circle B(0¯;1) and that the unknown domain Ω2 is star shaped with respect to the origin, i.e. Ω2={(r(θ)cos(θ),r(θ)sin(θ))|θ[0,2π)}, where r is a 2π-periodic smooth function with values in the interval (0,1). We takeX¯i=(cos(θi),sin(θi)),i=1,N¯to be the outer boundary collocation points uniformly distributed on Ω=B(0¯;1), where θi=2πi/N for i=1,N¯, and the source pointsξ¯1j=(Rcos(θj),Rsin(θj)),j=1,N¯,where R>1 is fixed. We also take(15) X¯i=(riNcos(ϑiN),riNsin(ϑiN)),i=N+1,2N¯(15) to be the inner unknown boundary collocation points on Ω2, and the inner and outer source pointsξ¯1j=(1δ)X¯j,ξ¯2jN=(1+δ)X¯j,j=N+1,2N¯,where δ(0,1). In Equation (Equation15), ri:=r(θi) for i=1,N¯.

The MFS coefficient vectors a¯=(aj)j=1,2N¯, b¯=(bj)j=1,N¯ in (Equation9) or (Equation14), and (Equation10), and the radii vector r¯=(ri)i=1,N¯ characterizing the star-shaped inner boundary Ω2 are determined by imposing the transmission conditions (Equation5), (Equation6) and the Cauchy data (Equation7), (Equation8) at the boundary collocation points (X¯i)i=1,2N¯ in a least-squares sense which recasts into minimizing the non-linear objective function(16) T(a¯,b¯,r¯):=u1fL2(Ω)2+u1ngL2(Ω)2+u1u2L2(Ω2)2+u1nu2nL2(Ω2)2+λ1{a¯2+b¯2}+λ2r¯2,(16) where λ1,λ20 are regularization parameters to be prescribed. Introducing the MFS approximations (Equation10), and say (Equation14) (with obvious modifications if (Equation9) is used) into (Equation16) yields(17) T(a¯,b¯,r¯)=i=1N[j=12NajGL(X¯i,ξ¯1j)f(X¯i)]2+i=N+12N[j=12NajGLn(X¯iN,ξ¯1j)g(X¯iN)]2+i=2N+13N[j=12NajGL(X¯iN,ξ¯1j)j=1NbjGMH(X¯iN,ξ¯2j;k2)]2+i=3N+14N[j=12NajGLn(X¯i2N,ξ¯1j)j=1NbjGMHn(X¯i2N,ξ¯2j;k2)]2+λ1{j=12Naj2+j=1Nbj2}+λ2j=1N1(rj+1rj)2.(17) The minimization of (Equation17) imposes 4N non-linear equations in the 4N unknowns (a¯,b¯,r¯). We can obviously have more equations than the unknowns if we take more boundary collocation points than sources. If there is noise in the measured data (Equation8), we replace the exact g in (Equation17) by the noisy gε given by(18) gϵ(X¯i)=g(X¯i)+ϵi,i=1,N¯,(18) where ϵi are random variables generated (using the NAG routine D05DDF) from a Gaussian normal distribution with mean zero and standard deviation(19) σ=p×maxΩ|g|,(19) where p represents the percentage of noise. In Equation (Equation17), the normal derivatives of GL and GMH, via (Equation12) and (Equation11), are given by(20) GLn(X¯,ξ¯)=(X¯ξ¯)·n¯2πX¯ξ¯2,GMHn(X¯,ξ¯;k2)=k2(X¯ξ¯)·n¯X¯ξ¯K1(k2X¯ξ¯),(20) where K1 is the modified Bessel function of the second kind of order one, and(21) n¯(X¯)={cos(θ)i¯+sin(θ)j¯,X¯Ω,1r2(θ)+r2(θ)[(r(θ)sin(θ)+r(θ)cos(θ))i¯+(r(θ)cos(θ)r(θ)sin(θ))j¯],X¯Ω2,(21) where i¯=(1,0) and j¯=(0,1). The modified Bessel function K1 is computed using the NAG routine S18ADF. In Equation (Equation21), the derivative r is approximated using backward finite differences as(22) r(θ)riri1θiθi1,i=1,N¯,(22) with the convention that r0=rN and θ0=0.

Fig. 5 The reconstructed inner boundary with various regularisation parameters λ1=0,λ2{108,105,103,101} for p=10% noise, for Example 1.

Fig. 5 The reconstructed inner boundary with various regularisation parameters λ1=0,λ2∈{10−8,10−5,10−3,10−1} for p=10% noise, for Example 1.

Fig. 6 The numerical solutions for the normal derivative u1/n(1,θ), obtained for various values of M=N{20,40,80} with (a) no regularisation, and (b) regularisation parameter λ=106 for the direct problem associated to Example 2.

Fig. 6 The numerical solutions for the normal derivative ∂u1/∂n(1,θ), obtained for various values of M=N∈{20,40,80} with (a) no regularisation, and (b) regularisation parameter λ=10−6 for the direct problem associated to Example 2.

Fig. 7 The unregularized objective function for p{0,1,3,5}% noise, as a function of the number of iterations, for Example 2.

Fig. 7 The unregularized objective function for p∈{0,1,3,5}% noise, as a function of the number of iterations, for Example 2.

Fig. 8 The reconstructed inner boundary with no regularization for p{0,1,3,5}% noise, for Example 2.

Fig. 8 The reconstructed inner boundary with no regularization for p∈{0,1,3,5}% noise, for Example 2.

Fig. 9 The regularized objective function with various regularization parameters λ2=0, λ1{108,105,103,101} for p=5% noise, as a function of the number of iterations, for Example 2.

Fig. 9 The regularized objective function with various regularization parameters λ2=0, λ1∈{10−8,10−5,10−3,10−1} for p=5% noise, as a function of the number of iterations, for Example 2.

Fig. 10 The reconstructed inner boundary with various regularization parameters λ2=0, λ1{108,105,104,103} for p=5% noise, for Example 2.

Fig. 10 The reconstructed inner boundary with various regularization parameters λ2=0, λ1∈{10−8,10−5,10−4,10−3} for p=5% noise, for Example 2.

Fig. 11 The reconstructed inner boundary with various regularization parameters λ2=0, λ1{108,105,103,101} for p=5% noise, for Example 3.

Fig. 11 The reconstructed inner boundary with various regularization parameters λ2=0, λ1∈{10−8,10−5,10−3,10−1} for p=5% noise, for Example 3.

Fig. 12 The reconstructed inner boundary for (a) full angle data for λ1=101,λ2=0, and (b) limited angle data for λ1=λ2=109, for no noise, for Example 3.

Fig. 12 The reconstructed inner boundary for (a) full angle data for λ1=10−1,λ2=0, and (b) limited angle data for λ1=λ2=10−9, for no noise, for Example 3.

The minimization of the objective function (Equation17) is accomplished computationally using the NAG routine E04FCF, which is a comprehensive algorithm for minimizing an unconstrained sum of squares of nonlinear functions. Note that the NAG routine E04FCF does not require the user to supply the gradient of (Equation17). If required, the physical constraints 0<ri<1 for i=1,N¯, that the inhomogeneity Ω2 stays within the host domain Ω during the iteration proccess can be imposed manually during the iterative procedure by adjustment at each iteration. The minimization process usually terminates when either a user-specified tolerance is achieved, or when a user-specified maximum number of iterations is reached.

Numerical results and discussion

In this section, three examples in two dimensions are presented in order to show the accuracy and stability of the MFS described in the previous section.

We take k1=0 and k2=1. In all numerical experiments, the initial guess for the unknown vectors a¯, b¯ are 0¯ and the initial guess for the inner boundary is taken to be a circle located at the origin with radius 0.7, i.e. the initial guess for r¯ is 0.7¯. Moreover, as required by the NAG routine E04FCF used, the tolerance XTOL was set to 106 and the maximum number of function evaluations, MAXCAL, was set to 1000. However, it is noted that by increasing the MAXCAL to be large, say 400 × (number of unknowns), as suggested by the NAG Fortran library manual, the computational time increases significantly and moreover, it does not produce more accurate numerical results. We also take R=2, M=N=20 for examples 1 and 2, and M=N=40 for example 3. Clearly, the rigorous choice of the regularization two-parameter family λ1 and λ2 in the nonlinear Tikhonov functional (Equation16) is very challenging. One can attempt the usual discrepancy principle or the more recent L-surface criterion [Citation12] but the calculations are very expensive and close to being prohibitive. In our study, we have investigated by trial and error several values for λ1 and λ2, and although we are not optimal in their difficult choice, at least we can discard the unstable solutions and provide some nearly optimal plausible stable candidates. Nevertheless, more work should be undertaken on the subject of multiple regularization in the future.

Example 1

We first consider an example for which the analytical solution is available and given by, see [Citation5],(23) u1(r,θ)=1+a0ln(r),R0<r<1,(23) where a0=R0I1(R0)I0(R0)R0I1(R0)ln(R0),(24) u2(r,θ)=A0I0(r),0<r<R0,(24) where A0=1I0(R0)R0I1(R0)ln(R0). The modified Bessel functions of the first kind of order 0 and 1, namely I0 and I1 are computed using the NAG routines S18AEF and S18AFF, respectively.

In this example, the unknown inner boundary is the disk(25) Ω2=B(0¯;R0)={(x,y)R2|x2+y2<R02}(25) of radius R0=0.5 and δ is taken to be 0.5. This analytical solution satisfies problem (Equation4)–(Equation8) and (Equation12), with(26) f(θ)=1,θ(0,2π](26) and(27) u1n(1,θ)=g(θ)=a0,θ(0,2π].(27) Figure shows the objective function (Equation17) with no regularization, i.e. λ1=λ2=0, when p{0,1,5,10}% noise is added in the input data (Equation27), as a function of the number of iterations. From this figure, it can be seen that the unregularized objective cost functional decreases rapidly within 6–8 iterations to either a very low value of O(1029) for p=0, or to a stationary level for p>0. Moreover, as expected, this level of stationarity decreases with decreasing the level of noise p.

In Figure , we present the reconstructed inner boundary and the exact shape (Equation25) with no regularization for p{0,1,5,10}%. As expected, as the level of noise increases, since no regularization is imposed, the reconstructed inner boundary becomes unstable.

Next, regularization is used in (Equation17) in order to stablize the numerical solutions. Figures and show the regularized objective function (Equation17) and the reconstructed inner boundary, respectively, for p=10% noise and various regularization parameters λ1=0, λ2{108,105,103,101}. From Figure , it can be seen that accurate and stable numerical solutions are achieved for λ1=0 and λ2 in the range 103 to 101, whilst clearly for λ2105 the obtained reconstructions become unstable.

Example 2

In the second example, we consider a more complicated bean-shaped inclusion Ω2 given by the radial parameterization(28) r(θ)=0.5+0.4cos(θ)+0.1sin(2θ)1+0.7cos(θ),θ(0,2π](28) within the unit circle Ω=B(0¯,1). The Dirichlet data (Equation7) on Ω is taken to be the same as in Example 1 and given by Equation (Equation26). We also choose δ=0.3.

Since in this case no analytical solution is available, the Neumann flux data (Equation8) on Ω is simulated numerically by solving, using the MFS, the direct problem given by Equations (Equation3)–(Equation6) and (Equation26), when Ω2 is known and given by (Equation28). In this case, the numerical solutions for the normal derivative u1/n(1,θ) on Ω, obtained for various values of M=N{20,40,80} with no regularization λ=0, and with regularization λ=106 are shown in Figure . From Figure it can be seen that the numerical results are convergent as the number of degrees of freedom increases. Furthermore, a small regularization with λ=106 tends to improve the independance of the mesh between M=N=40 and 80. Twenty evenly spread points out of the curve M=N=80 with λ=106 of Figure are chosen as input Neumann numerically simulated data (Equation8) in the inverse problem. Next, in order to avoid committing an inverse crime we solve the inverse problem with M=N=20.

Figures and for Example 2 are anologous to Figures and for Example 1 and the same conclusions about the unstable nature of the unregularized solution, as the level of noise p increases, can be drawn. Figures and for Example 2 present similar characteristics to Figures and for Example 1. In order to investigate a different situation to that in Example 1 we regularize with λ1>0 instead of λ2>0. From Figure , it can be seen that a stable and reasonably accurate numerical solution is obtained for λ1=103 (and λ2=0), whilst clearly for λ1104 the obtained reconstructions become unstable.

Example 3

We finally consider reconstructing a complicated pear-shaped inclusion Ω2 given by the radial parameterization(29) r(θ)=0.6+0.125cos(3θ),θ(0,2π],(29) within the unit circle Ω=B(0¯,1). The Dirichlet voltage (Equation2) on Ω was taken to be the same as in Example 1 and given by Equation (Equation26). We also take δ=0.3. Since no analytical solution of this example is available, the Neumann flux data (Equation8) on Ω is simulated numerically by solving the direct problem (Equation5)–(Equation8) and (Equation26), when Ω2 is known and given by (Equation29), using the MFS with M=N=80 and λ=106. In order to avoid committing an inverse crime, we use a different number M=N=40 in the inverse problem. Next, two cases are considered for this example, as follows.

Recovery from full angle data

We consider the first case when the full data of measurment of the flux g on the outer boundary Ω is available. Figure shows reconstructed inner boundary with p=5%, for various regularization parameters λ1{108,105,103,101}, λ2=0. From this figure it can be seen that a stable and reasonably accurate numerical solution is obtained for λ1 between 103 and 101 (and λ2=0).

Recovery from limited angle data

Finally, we consider the case of limited flux measurment data g prescribed on the subportion {(1,θ)|θ[0,π]} of the full outer boundary Ω=B(0¯,1). Figures and show the reconstructed inner boundary for both full and limited flux data with or without noise. From these figures it can be seen that, as expected, the reconstructed inner boundary from full angle data is much more accurate than that obtained from limited angle data.

Fig. 13 The reconstructed inner boundary for (a) full angle data for λ1=101,λ2=0, and (b) limited angle data for λ1=λ2=109, for p=5% noise, for Example 3.

Fig. 13 The reconstructed inner boundary for (a) full angle data for λ1=10−1,λ2=0, and (b) limited angle data for λ1=λ2=10−9, for p=5% noise, for Example 3.

Conclusions

In this paper, a novel numerical method based on a regularized iterative MFS has been developed for the reconstruction of an inhomogeneity in an inverse problem for the elliptic (modified Helmholtz) equation. The choice of the two-family multiple regularization parameters was based on trial and error. More rigorous choices of these parameters should be investigated in any future work. Several examples have been investigated showing that the MFS is accurate (for exact data) and stable (for noisy data). Future work will concern extending the MFS developed in this paper to the reconstruction of a source domain from Cauchy data.[Citation13]

Acknowledgments

B. Bin-Mohsin would like to thank King Saud University for their financial support in this research.

References

  • Lin C-N, Jang J-Y. A two-dimensional fin efficiency analysis of combined heat and mass transfer in elliptic fins. Int. J. Heat Mass Transfer. 2002;45:3839–3847.
  • Fang W, Cumberbatch E. Inverse problems for metal oxide semiconductor field-effect transistor contact resistivity. SIAM J. Appl. Math. 1992;52:699–709.
  • Isakov V, Inverse obstacle problems, Inverse Probl. 2009;25:123002 (pp. 18).
  • Kang H, Kwon K, Yun K. Recovery of an inhomogeneity in an elliptic equation. Inverse Probl. 2001;17:25–44.
  • Hettlich F, Rundell W. Recovery of the support of a source term in an elliptic differential equation. Inverse Probl. 1997;13:959–976.
  • Kim S, Yamamoto M. Uniqueness in identification of the support of a source term in an elliptic equation. SIAM J. Math. Anal. 2003;35:148–159.
  • Kim S. Uniqueness determination of inhomogeneity in an elliptic equation. Inverse Probl. 2002;18:1325–1332.
  • Bin-Mohsin B, Lesnic D. Determination of inner boundaries in modified Helmholtz inverse geometric problems using the method of fundamental solutions. Math. Comput. Simulation. 2012;82:1445–1458.
  • Lesnic D, Bin-Mohsin B. Inverse shape and surface heat transfer coefficient identification. J. Comput. Appl. Math. 2012;236:1876–1891.
  • Karageorghis A, Lesnic D, Marin L. Survey of applications of the MFS to inverse problems. Inverse Probl. Sci. Eng. 2011;19:309–336.
  • Bin-Mohsin B, Lesnic D. The method of fundamental solutions for Helmholtz-type equations in composite materials. Comput. Math. Appl. 2011;62:4377–4390.
  • Belge M, Kilmer ME, Miller EL. Efficient determination of multiple regularisation parameters in a generalized L-curve framework. Inverse Probl. 2002;18:1161–1183.
  • Ikehata M. Reconstruction of a source domain from the Cauchy data. Inverse Probl. 1999;15:637–645.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.