288
Views
3
CrossRef citations to date
0
Altmetric
Articles

To solve the inverse Cauchy problem in linear elasticity by a novel Lie-group integrator

Pages 641-671 | Received 06 Jun 2012, Accepted 22 Sep 2013, Published online: 24 Oct 2013

Abstract

In this paper, we propose a simple, iteration free and easy-to-implement numerical algorithm for the solution of inverse Cauchy problem in linear or nonlinear elasticity. The bottom of a finite rectangular plate is imposed by overspecified boundary data, and we seek unknown data on the top side. A spring-damping transform method (SDTM) is introduced to the Navier equations, such that after a discretization by the differential quadrature method, we can apply a novel Lie-group integrator, namely the mixed group-preserving scheme (MGPS), to solve them as an initial value problem. Several numerical examples including nonlinear ones are examined to show that the MGPS can overcome the ill-posed behaviour of the inverse Cauchy problem in elasticity, which has good efficiency and stability against the noisy disturbance, even with an intensity large up to 10% and 20%.

1 Introduction

There are many important issues in the reconstruction of deformed shape for the purposes of structural health monitoring (SHM), damage detection, real-time shape control and shape modification. The shape reconstruction techniques from discrete position data have been under development for numerous years and there is a multitude of publications in the literature, especially with regard to the polynomial and spline reconstructions. Research on the reconstruction of displaced shape from strain measurements is less frequently encountered.[Citation1] Indeed, a real-time reconstruction of the deformed shape of structure is a key technology for SHM. This sort problem is commonly referred to as a shape sensing problem, which is one class of the general inverse Cauchy problem in elasticity, if the deformation is restricted in the elastic range. For a real-time reconstruction of the deformed shape, it is necessary to develop a fast and stable algorithm for solving the resultant inverse Cauchy problems.

During the past several years, the science and engineering communities have paid much attention to the inverse Cauchy problem in linear elasticity, which is a noncharacteristic initial value problem of the elliptic type partial differential equations (PDEs). According to the Cauchy–Kowalewski theorem, the solution of an analytic Cauchy problem for PDEs exists and is unique. However, the inverse Cauchy problem is quite difficult to be solved numerically, since the solution does not depend continuously on the given data, that is, a small error in the specified data may result in an incorrect solution. Therefore, we must treat this type inverse problem with a suitable numerical algorithm, which compromises stability and accuracy.

The inverse Cauchy problem in linear elasticity was studied theoretically by Yeih et al. [Citation2], who analyzed its existence, uniqueness and continuous dependence on the data, and proposed an alternative regularization procedure, namely the fictitious boundary indirect method, based on simple and double layer potential theory. Its numerical implementation was undertaken by Koya et al. [Citation3], who employed the boundary element method (BEM) and the Nystro¨m method for discretizing the boundary integrals.

The inverse Cauchy problem in linear elasticity is to solve the boundary value problem of Navier partial differential equations given by some overspecified Cauchy data on a partial portion of the boundary and other portion given no data, which is proved to have a unique solution if the solution exists. However, the problem of numerical instability lends the researchers a great challenge, because the inverse of the original forward operator is not obtainable. It means that the continuous dependence of the solution on the given data is not satisfied in the Hardmard sense. To treat this kind inverse problem, many techniques were proposed. The most famous one is the Tikhonov regularization method, which perturbs the original ill-posed problem into a constrained minimization problem. The Lagrangian multiplier, i.e. the regularization parameter, is not known in advance, which can be determined by the L-curve concept or some discrepancy principles. Except for this method, the truncated singular value decomposition method (SVD) had also been used. All of those methods try to improve the ill-conditioning behaviour of the resultant linear equations system.

The iterative algorithm of Kozlov et al. [Citation4], which reduces the inverse Cauchy problem to solve a sequence of well-posed boundary value problems, was implemented using the BEM for linear elastic materials by Marin et al. [Citation5], Marin and Lesnic,[Citation6] and Comino et al. [Citation7]. Ellabib and Nachaoui [Citation8] have numerically investigated the relaxation of the alternating iterative algorithm of Kozlov et al. [Citation4], while further investigations were carried out by Marin and Johansson,[Citation9] who also proposed an alternative way of relaxation of both the prescribed displacements and tractions on the overspecified boundary. Moreover, Marin and Johansson [Citation9] have proved the convergence of these schemes, and introduced appropriate optimal stopping rules to implement these algorithms with relaxation using the BEM. Andrieux and Baranger [Citation10], and Baranger and Andrieux [Citation11] have reformulated the inverse Cauchy problem for elastic material as an energy error minimization problem, with the unknowns given by the surface displacement and traction vectors on the underspecified boundary of the solid.

Four regularization methods for stably solving the inverse Cauchy problem in linear elasticity, namely the Tikhonov regularization method, the SVD, the conjugate gradient method (CGM) and the alternating iterative algorithm of Kozlov et al. [Citation4], were compared by Marin and Lesnic,[Citation12] and they found that the truncated SVD outperforms the Tikhonov regularization method, whilst the latter outperforms the CGM. Marin [Citation13] has used the above alternating iterative algorithm of Kozlov et al. [Citation4] together with the method of fundamental solutions (MFS) to solve the two-dimensional inverse Cauchy problem in linear elasticity. The inverse Cauchy problem in linear elasticity with the L2-boundary data was approached by Marin and Lesnic [Citation14] and Marin [Citation15] by combining the BEM with the Landweber–Fridman method and the minimal error method.

Recently, Chen and his coworkers have used several alternating regularization techniques in conjunction with the singular boundary method (SBM) and the MFS to solve the inverse Cauchy problems.[Citation16Citation19]

In contrast to those methods, the present paper aims to provide a simple numerical computation method based on a novel Lie-group integrator, directly integrating the inverse Cauchy problem as an initial value problem without needing of any iteration. In order to express the new method simpler and clearer, we first restrict ourself to the inverse Cauchy problem for a rectangular plate, because in this domain a numerical method of lines can be easily developed for the discretization of the Navier equations into a linear system of ODEs. An extension to arbitrary plane domain is possible. As mentioned above, the inverse Cauchy problem is a noncharacteristic problem, and there were only rare papers to develop a direct numerical integration method to solve it.[Citation20Citation22] After a suitable and mathematically equivalent transformation technique by introducing the spring/damping terms into the Navier equations for an enhancement of the numerical stability in Section2, we develop a numerical method in Section 3 to solve the inverse Cauchy problem in linear elasticity for a rectangular plate, which is discretized into a linear ODEs system by using the numerical method of lines and differential quadrature method. In Section 4, we introduce three different group-preserving schemes, of which the last, as a combination of the previous two methods, is a novel Lie-group integrator named the mixed group-preserving scheme (MGPS), which is used in the numerical integration of the Cauchy problem in linear elasticity. The numerical results are given in Section 5. While Section 6 devotes to extend the Cauchy problem for a linear elliptic equation into that defined in an arbitrary plane domain, Section 7 extends the MGPS to nonlinear inverse Cauchy problems. Finally, the conclusions are drawn in Section 8.

2 A spring-damping transform method for the Cauchy problem of elasticity

2.1 Mathematical formulation

Consider an isotropic linear elastic material which occupies a bounded domain ΩR2 and assume that Ω=[0,a]×[0,b]. Γ1=[0,a]×{y=0} denotes the bottom of the plate Ω, and Γ2=[0,a]×{y=b} denotes the top side. In the absence of body force, the equilibrium equations written with respect to the displacement vector u(x)=(u1(x),u2(x)), known as the Navier equations, are given by [Citation23]:1 G2ui(x)xjxj+G1-2ν2uj(x)xixj=0,xΩ,i=1,2,1 where the repeated index is summed from 1 to 2, and G and ν are, respectively, the shear modulus and the Poisson ratio of elastic material.

The strains εij(x),i,j=1,2, are related to the displacement gradients by the small deformation kinematic relations:2 εij(x)=12ui(x)xj+uj(x)xi,xΩ¯,i,j=1,2,2 while the stresses σij(x),i,j=1,2, are related to the strains through the Hooke constitutive law:3 σij(x)=2Gεij(x)+ν1-2νεkk(x)δij,xΩ¯,i,j=1,2,3 where δij is the Kronecker delta tensor. Let n(x) denote the outward normal vector on Γ, and t(x) be the traction vector at a point xΓ whose components are given by4 ti(x)=σij(x)nj(x),i=1,2.4 As shown by Fung and Tong [Citation23], in the direct problem of linear elasticity, the knowledge of the displacement and/or traction vectors on the whole boundary Γ gives the corresponding Dirichlet, Neumann or mixed boundary condition which enables us to determine the displacement vector in the whole domain Ω. Then, the strain tensor εij can be calculated from the kinematic relations in Equation (Equation2) and the stress tensor σij is determined by using the constitutive law in Equation (Equation3).

If it is possible to measure both the displacement and the displacement gradient vectors on a part of the boundary Γ, say Γ1, and on another portion Γ2 it gives no data, then this leads to a mathematical formulation of an inverse Cauchy problem, which consists of Equation (Equation1) and the following boundary conditions:5 ui(x)=u~i(x),ui,j(x)=u~i,j(x),xΓ1,i=1,2,j=2,5 where u~ and u~,j are prescribed vector valued functions. In the above formulation of the boundary conditions in Equation (Equation5), it can be seen that the boundary Γ1 is overspecified by prescribing both the displacement u|Γ1=u~ and the j-th gradient u,j|Γ1=u~,j vector, whilst the boundary Γ2 is underspecified since both the displacement u|Γ2 and the traction t|Γ2 vectors are unknown and have to be determined for a completion of the boundary data. On the portion of Γ-(Γ1Γ2) we also have one type, not overspecified, boundary data for the uniqueness of solution. This problem, termed the inverse Cauchy problem, is much more difficult to be solved both analytically and numerically than the direct problem, since the solution does not satisfy the general conditions of well-posedness. Although the inverse Cauchy problem may have a unique solution, it is well known that this solution is unstable with respect to a small perturbation of the data on Γ1. Thus, the problem is ill-posed, and we cannot use a direct approach to solve the system of linear equations which arises from the discretization of Equation (Equation1) and the boundary conditions in Equation (Equation5). Therefore, as mentioned above, some regularization methods are required in order to stably and accurately solve the inverse Cauchy problem in linear elasticity.

2.2 A spring-damping transform method

Before the introduction of our stabilization technique, we introduce the super-dashpot time-dependent stabilization technique developed by Essers.[Citation24] For simplicity we only consider a one-dimensional steady n equations system of fluid dynamics:6 A(u)ux=0,6 where A is an n×n matrix function and u is an n-dimensional unknown vector of variables.

Essers [Citation24] has introduced his first DASH1 method by considering the 2n systems:7 qt=A(u)ux,7 8 ut+ku=ATqx,8 where k>0 is an internal damping factor, and q is a supplemental vector function. Eliminating q between Equations (Equation7) and (Equation8) one obtains9 2ut2+kut=D2ux2,9 where D=ATA is obviously a symmetric matrix with real positive eigenvalues dj2,j=1,,n. The above equations are actually similar to those of a vibrating string damped by dashpots. Essers [Citation24] used this technique to stabilize and accelerate the iterative convergence of computational fluid algorithms.

Next, we consider a Cauchy problem of Laplace equation defined in a rectangle being studied by Liu and Kuo [Citation20]:10 2ux2+2uy2=0,0<x<,0<y<b,10 11 u(x,0)=f(x),0x,11 12 uy(x,0)=g(x),0x,12 13 u(0,y)=u0(y),u(,y)=u(y),0yb,13 where f(x),g(x),u0(y) and u(y) are given functions. Here, we give a motivation to introduce a spring-damping method by using the following transformation:14 u(x,y)=eαyU(x,y).14 Such that Equations (Equation10)–(Equation13) become15 2Ux2+2Uy2+2αUy+α2U=0,15 16 U(x,0)=f(x),0x,16 17 Uy(x,0)=g(x)-αf(x),0x,17 18 U(0,y)=e-αyu0(y),U(,y)=e-αyu(y),0yb.18 Liu and Kuo [Citation20] have proven that the introduction of the factor α can enhance the stability in the numerical integration of the above equations in the y-direction. For the purposes of comparison with Equation (Equation9), in Equation (Equation15) we replace y by t:19 2Ut2+2αUt+α2U=-2Ux2.19 Again, the left-hand side bears certain similarity to a vibrating string damped by dashpot and stabilized by spring. Unfortunately, the above equation is not a wave equation because a minus sign stands before 2U/x2.

Now we can comment the methods proposed by Essers [Citation24] and the one introduced in the above.

(i)

The method proposed by Essers [Citation24] makes a perturbation from Equation (Equation6) to Equations (Equation7) and (Equation8) and then to Equation (Equation9). The present method makes a mathematical transformation from Equation (Equation10) to Equation (Equation19), which is not a perturbation.

(ii)

In the method proposed by Essers [Citation24] there is only a dashpot; however, Equation (Equation19) possesses a dashpot and a spring.

(iii)

The method proposed by Essers [Citation24] modifies the type of the governing equations. The present method does not alter the type of the governing equations.

(iv)

Both methods posses the effect of ‘stabilization’ in the numerical algorithms.

The above process is called a spring-damping transform method (SDTM) in this paper. Although it is quite simple, as to be shown below the spring-damping method is effective for our inverse Cauchy problem of linear elasticity to increase the accuracy of numerical solutions. However, the essential ‘unstability’ of the inverse Cauchy problem must be further treated by a supplemental stabilization technique in Section 4.3.

2.3 The transformed equations

For our convenience, we rewrite Equation (Equation1) into a componential form with u(x,y):=u1(x,y) and v(x,y):=u2(x,y):20 ν12ux2+2uy2+11-2ν2vxy=0,0<x<a,0<y<b,20 21 2vx2+ν12vy2+11-2ν2uxy=0,0<x<a,0<y<b,21 22 u(x,0)=f1(x),0xa,22 23 v(x,0)=f2(x),0xa,23 24 uy(x,0)=g1(x),0xa,24 25 vy(x,0)=g2(x),0xa,25 26 u(0,y)=h1(y),u(a,y)=h2(y),0yb,26 27 v(0,y)=h3(y),v(a,y)=h4(y),0yb,27 where f1(x),f2(x),g1(x),g2(x), hk(y),k=1,,4 are given functions, and28 ν1=2(1-ν)1-2ν.28 We must emphasize that it is difficult to directly apply the numerical integration method to integrate the above equations as an initial value problem to obtain the solution u(x,y) and v(x,y).

In the present paper, we will develop a novel Lie-group integration method to directly solve the above inverse Cauchy problem by recovering the data on the top side y=b. However, before that we need to make the above equations ‘more stable’. The integration direction is along the y-axis, and thus as that done in Section 2.2 we consider the following transformations:29 u(x,y)=eαyU(x,y),v(x,y)=eαyV(x,y).29 From Equations (Equation20)–(Equation27) and (Equation29) it follows that30 ν12Ux2+2Uy2+2αUy+α2U+11-2ν2Vxy+αVx=0,30 31 2Vx2+ν12Vy2+2αVy+α2V+11-2ν2Uxy+αUx=0,31 32 U(x,0)=f1(x),0xa,32 33 V(x,0)=f2(x),0xa,33 34 Uy(x,0)=g1(x)-αf1(x),0xa,34 35 Vy(x,0)=g2(x)-αf2(x),0xa,35 36 U(0,y)=e-αyh1(y),U(a,y)=e-αyh2(y),0yb,36 37 V(0,y)=e-αyh3(y),V(a,y)=e-αyh4(y),0yb.37 It can been that in Equations (Equation30) and (Equation31) we have introduced four extra terms: two spring terms α2U and α2V and two damping terms 2αU/y and 2αV/y. In this regard, the present ‘stabilization technique’ is quite different from that of the Tikhonov regularization method.[Citation25, Citation26] As mentioned in Section 1, the Tikhonov regularization method perturbs the original ill-posed problem into a ‘less ill-posed one’, not necessarily a well-posed one; however, the SDTM does not perturb the original ill-posed problem, which is merely a mathematical transformation. The ill-posedness still exists in the above equations. In Section 4.3, a new technique will be used to overcome the ill-posedness of the inverse Cauchy problem.

A similar regularization technique in the time domain has been first used by Liu [Citation27] to stabilize the backward heat conduction problem, and then Liu and Kuo,[Citation20] Liu et al. [Citation21], and Liu and Chang [Citation22] extended it to the spatial domain, and developed the spring and damping stabilization techniques to the inverse Cauchy problems of elliptic type PDEs.

Here we extend the spring-damping transform method (SDTM) as shown in Section2.2 for the Laplace equation to the inverse Cauchy problem in linear elasticity. However, we must emphasize that the above equations are also of the elliptic type systems, and the associated Cauchy problems are also ill-posed in nature, because the transformations in Equation (Equation29) do not alter the type of the governing equations. After the numerical of lines is used to discretize the governing equations, in Section 4.3 we will introduce a mixed-group preserving scheme to overcome the ill-posed behaviour of the inverse Cauchy problem.

Because the transformations invoked in Equation (Equation29) are purely mathematical, which do not change the inverse Cauchy problems which exhibit boundary singularities, e.g. an elastic solid with a crack/notch or jump/discontinuity in the available boundary conditions. However, those inverse Cauchy problems are much more complex than the present ones, and we do not treat them in this paper.

3 Numerical method of lines

The numerical method of lines is simple in concept that for a given system of PDEs discretize all but one of the independent variables. The semi-discrete procedure yields a coupled system of ODEs which are then being numerically integrated. For Equations (Equation30) and (Equation31), we adopt the numerical method of lines to discretize the spatial coordinate x and use the Differential Quadrature method (see the Appendix) to approximate the following differential terms:U(x,y)xx=iΔx=bijUj(y),38 2U(x,y)x2x=iΔx=aijUj(y),38 39 V(x,y)xx=iΔx=bijVj(y),39 40 2V(x,y)x2x=iΔx=aijVj(y),40 where Δx=a/(n+1) is a uniform discretization spacing length, Ui(y)=U(iΔx,y), Vi(y)=V(iΔx,y), and the coefficient matrices bij and aij=bikbkj are introduced in the Appendix. Now Equations (Equation30) and (Equation31) can be approximated by41 Ui(y)y=Wi(y),41 42 Wi(y)y=-ν1aijUj(y)-2αWi(y)-α2Ui(y)-11-2ν[bijZj(y)+αbijVj(y)],42 43 Vi(y)y=Zi(y),43 44 Zi(y)y=-1ν1aijVj(y)-2αZi(y)-α2Vi(y)-12(1-ν)[bijWj(y)+αbijUj(y)].44 The next step is to advance the solution from the initial conditions given at y=0 to a terminal position y=b. Really, Equations (Equation38)–(Equation41) have totally 4n coupled linear differential equations for the 4n variables Ui(y),Wi(y),Vi(y),Zi(y),i=1,2,,n, which can be numerically integrated to obtain the solutions. However, we cannot directly apply the conventional numerical integration methods to integrate the above ODEs; the essentially numerical instability may cause a terribly incorrect solution. A numerical example will be given in Section 5 to show that the above equations are still highly ill-posed, which do not allow a direct numerical integration to obtain the solution, and thus we will develop a mixed group-preserving scheme to solve this ill-posed problem.

4 Group-preserving schemes

Liu [Citation28] has derived a Lie-group transformation for the augmented dynamics on the future cone, and developed a group-preserving scheme for an effective numerical solution of nonlinear ordinary differential equations (ODEs).

4.1 Forward group-preserving scheme

Liu [Citation28] has embedded the differential equations system into an augmented dynamical system, which concerns with not only the evolution of state variables but also the evolution of the magnitude of state variables vector. That is, for an n ODEs system:45 x˙=f(x,t),xRn,tR,45 we can embed it into the following n+1-dimensional augmented dynamical system:46 ddtxx=0n×nf(x,t)xfT(x,t)x0xx.46 Here we assume x>0, and hence the above system is well defined.

It is obvious that the first row in Equation (Equation43) is the same as the original Equation (Equation42), but the inclusion of the second row in Equation (Equation43) gives us a Minkowskian structure of the augmented state variables of X:=(xT,x)T, satisfying a future cone condition:47 XTgX=0,47 where48 g=In0n×101×n-148 is a Minkowski metric, In is the identity matrix of order n and the superscript T stands for the transpose. In terms of (x,x), Equation (Equation44) becomes49 XTgX=x·x-x2=x2-x2=0,49 where the dot between two n-dimensional vectors denotes their Euclidean inner product. The cone condition is thus the most natural constraint that we can impose on the dynamical system (Equation43).

Consequently, we have an n+1-dimensional augmented differential equations system:50 X˙=AX50 with a constraint in Equation (Equation44), where51 A:=0n×nf(x,t)xfT(x,t)x0,51 satisfying52 ATg+gA=0,52 is a Lie algebra so(n,1) of the proper orthochronous Lorentz group SOo(n,1). This fact prompts us to devise the so-called group-preserving scheme, whose discretized mapping G exactly preserves the following properties:53 GTgG=g,53 54 detG=1,54 55 G00>0,55 where G00 is the 00th component of G.

We assume that the value of A at the k-th time step is denoted by A(k), which is viewed as a constant matrix. An exponential mapping of A(k) admits a closed-form representation:56 exp[hA(k)]=In+(ak-1)fk2fkfkTbkfkfkbkfkTfkak,56 where57 ak:=coshhfkxk,bk:=sinhhfkxk.57 Consequently, we can derive58 xk+1=xk+η1(k)fk,58 59 η1(k):=(ak-1)fk·xk+bkxkfkfk2.59 This scheme preserves all the group properties in Equations (Equation50)–(Equation52) for all h>0, which is called a forward group-preserving scheme.

4.2 Backward group-preserving scheme

We can also embed Equation (Equation42) into the following n+1-dimensional augmented dynamical system:60 ddtx-x=0n×n-f(x,t)x-fT(x,t)x0x-x.60 It is obvious that the first equation in Equation (Equation57) is the same as the original Equation (Equation42), but the inclusion of the second equation gives us a Minkowskian structure of the augmented state variables of X:=(xT,-x)T, satisfying a past cone condition:61 XTgX=x·x-(-x)2=x2-x2=0.61 Here, we should stress that the cone condition being imposed on the dynamical system (Equation43) is a future cone, and that for the dynamical system (Equation57) the imposed cone condition (Equation58) is a past cone.

Similarly, we have an n+1-dimensional augmented differential equations system:62 X˙=BX62 with a constraint in Equation (Equation58), where63 B:=0n×n-f(x,t)x-fT(x,t)x063 satisfying64 BTg+gB=0,64 is a Lie algebra so(n,1) of the proper orthochronous Lorentz group SOo(n,1).

Similarly, we assume that the value of B at the k-th time step is denoted by B(k), which is viewed as a constant matrix. Accordingly, an exponential mapping of B(k) admits a closed-form representation:65 exp[-hB(k)]=In+(ak-1)fk2fkfkTbkfkfkbkfkTfkak,65 where ak and bk were defined by Equation (Equation54), and we have66 xk-1=xk+η2(k)fk,66 67 η2(k):=(ak-1)fk·xk-bkxkfkfk2.67 This scheme preserves the group properties in Equations (Equation50)–(Equation52) for all h>0, which is called a backward group-preserving scheme by Liu et al. [Citation29].

Comparing Equations (Equation56) and (Equation64) it is interesting to note that these two numerical schemes have the same form in addition that the sign before bkxkfk in the numerators.

4.3 Mixed group-preserving scheme

Although the above two schemes, forward and backward group-preserving schemes, are very accurate, but we cannot directly apply them to obtain the solution, because the inverse Cauchy problems are sensitive to noisy disturbance, which is added on the data by68 f^i=f(xi)[1+sR(i)],g^i=g(xi)[1+sR(i)],68 where R(i) are random numbers in [-1,1]. In order to overcome this instability we apply the following novel Lie-group integrator to solve Equations (Equation38)–(Equation41):69 Ui(yj+1)=Ui(yj)+η1(j)Wi(yj),69 70 Wi(yj+1)=Wi(yj)+η2(j)H1i(yj),70 71 Vi(yj+1)=Vi(yj)+η1(j)Zi(yj),71 72 Zi(yj+1)=Zi(yj)+η2(j)H2i(yj),72 where73 H1i(y):=-ν1aijUj(y)-2αWi(y)-α2Ui(y)-11-2ν[bijZj(y)+αbijVj(y)],73 74 H2i(y):=-1ν1aijVj(y)-2αZi(y)-α2Vi(y)-12(1-ν)[bijWj(y)+αbijUj(y)],74 and η1(j) and η2(j) are defined in Equations (Equation56) and (Equation64), respectively. Here, instead of η1(j) used in the forward group-preserving scheme (Equation55), the term η2(j) used in Equations (Equation67) and (Equation69) can impress the instability of inverse Cauchy problem. The above scheme is a mixture of the forward group-preserving scheme (Equation55), which uses η1, and the backward group-preserving scheme (Equation63), of which η2 is used; hence, we may name the present novel Lie-group integrator in Equations (Equation66)–(Equation69) as a mixed group-preserving scheme (MGPS). Equations (Equation66) and (Equation68) are the integrations of Equations (Equation38) and (Equation40), which are just the equations for the definitions of U/y and V/y; they are unimportant, and in their integrations we use η1(j).

5 Numerical examples

In order to present the performance of the MGPS, we consider an isotropic linear elastic plate characterized by the material constants G=3.35×1010 N/m2 and ν=0.34 corresponding to a copper alloy, and we solve the inverse Cauchy problem in Equations (Equation1) and (Equation5) for four typical examples.

5.1 Example 1

We consider the following analytical solution for displacements:75 u(x,y)=σ02G(1+ν)xy,75 76 v(x,y)=-σ04G(1+ν)1+2ν21-2νx2-2+νy2,76 where -1<x<1,0<y<1. Here we fix σ0=1.5×1010 N/m2. It should be noted that for this example, the overspecified Cauchy data is available on a portion Γ1 of the boundary Γ such that measure(Γ1) = measure(Γ)/3.

First we use this simple example to demonstrate that although we have added a factor α=-2 into the numerical method of lines Equations (Equation38)–(Equation41), we cannot obtain stable solutions by using the GPS in Section 4.1 to integrate the above equations. Under a zero noise and with Δx=2/50, Δy=1/2000 used in the GPS, the errors of u and v at y=1 are shown in Figures (a) and (b), respectively. The maximum errors of u and v are, respectively, 7.49×1019 and 1.56×1019. It shows that the governing equations are highly ill-posed, which do not allow a direct numerical integration to obtain the solution, and thus we need to use the mixed group-preserving scheme (MGPS) developed in Section 4.3 to solve this ill-posed problem.

Figure 1. For example 1 by applying the GPS to integrate the governing equations with α=-2 leads to unstable solutions.

Figure 1. For example 1 by applying the GPS to integrate the governing equations with α=-2 leads to unstable solutions.

The above computation shows that if we only consider the stabilizing effect by spring and dashpot, the resulting equations are also very ill-posed, which cannot be integrated by using the usual integration method, like as the GPS used in the above case. Under a large noise with s=10%, and with Δx=2/50, Δy=1/2000 used in the MGPS, instead of GPS, the numerical solutions of displacements and tractions at y=1 are compared with the exact ones in Figure , where α=-2 is used. In order to further prove that the algorithm is robust, we raise the noise intensity to s=20%, and from Figure we can observe that the two numerical solutions have little difference. It can be seen that the present algorithm is quite robust, which can recover the data on the top side very well, even under a large noise up to s=10% and s=20%. The computational results are rather promising that one can apply the MGPS developed in Section 4.3 to solve the inverse Cauchy problems. Below we will employ the MGPS to solve all examples.

Figure 2. For example 1 comparing the recovered displacements and tractions with exact ones on the top side.

Figure 2. For example 1 comparing the recovered displacements and tractions with exact ones on the top side.

In Figure , we compare the numerical errors obtained by the MGPS with α=-1 and α=0 in the governing equations. The effect of α can be seen, which can largely increase the accuracy in the recovered solutions of u and v. Although α=0 is used, the MGPS itself can make a stable solution, however its accuracy is not so good.

Figure 3. For example 1 comparing the numerical errors of recovered displacements on the top side with α=-1, and α=0.

Figure 3. For example 1 comparing the numerical errors of recovered displacements on the top side with α=-1, and α=0.

Figure 4. For example 1 comparing the numerical errors of displacements and the Neumann data on the top side with different α.

Figure 4. For example 1 comparing the numerical errors of displacements and the Neumann data on the top side with different α.

In Figure we compare the numerical errors of displacements and the Neumann data of uy and vy on the top side obtained by the MGPS with different values of α in the range of α[-2,-0.5]. It can be seen that all the α in the range of α[-2,-1] can cause rather small errors; however, when α is larger than -1 the errors of uy and vy increase rapidly. In the range of α[-2,-1] the algorithm is insensitive to the parameter. If we alternate the roles of η1 and η2 in Equations (Equation66)–(Equation69), it leads to an incorrect result as shown in Figure.

Figure 5. For example 1 with η1 and η2 alternating, comparing the numerical solutions with exact ones.

Figure 5. For example 1 with η1 and η2 alternating, comparing the numerical solutions with exact ones.

5.2 Example 2

We consider the following analytical solution for displacements:77 u(x,y)=σ0(1-ν)2G(1+ν)x,v(x,y)=σ0(1-ν)2G(1+ν)(2-y+2x),0<x<1,0<y<1.77 Under a large noise with s=10%, and with Δx=1/20, Δy=1/1000 used in the MGPS, the numerical solutions of displacements and tractions at y=1 are compared with the exact ones in Figure , where α=-0.9 is used. It can be seen that the present algorithm is quite robust, which can recover the data on the top side very well, even under a large noise.

Figure 6. For example 2 comparing the recovered displacements and tractions with exact ones on the top side.

Figure 6. For example 2 comparing the recovered displacements and tractions with exact ones on the top side.

5.3 Example 3

In this example, we consider a simply support beam with the following analytical solution for displacements:78 u(x,y)=σ02G(1+ν)(x-a/2)y,v(x,y)=σ04G(1+ν)[(a-x)x-νy2],0<x<a,0<y<0.5,78 where a=10, and σ0=M/I with I being the inertial moment and M the applied moment. Here we fix σ0=1.5×1010 N/m2. For this example, a body force 2ν2/(1-2ν) should be added on the right-hand side in Equation (Equation21).

Under a large noise with s=10%, and with Δx=10/200, Δy=0.5/1000 used in the MGPS, the numerical solutions at y=0.5 are compared with the exact ones in Figure , where α=-0.2 is used. It can be seen that the present algorithm is quite robust, which can recover the data on the top side very well, even under a large noise. In order to test the influence of α, we compute the numerical solution under a different α=-0.5, of which the recovered Cauchy data on the top side as shown in Figure is less well. The errors of uy and vy are also acceptable. However, the accuracy of the recovered Neumann data of uy and vy was worse than that for the displacements.

Figure 7. For example 3 comparing the recovered displacements and the Neumann data with exact ones on the top side.

Figure 7. For example 3 comparing the recovered displacements and the Neumann data with exact ones on the top side.

5.4 Example 4

In this example, we consider a cantilever beam with the following analytical solution for displacements:79 u(x,y)=-σ02G(1+ν)(a-x)y,v(x,y)=-σ04G(1+ν)[(a-x)2+νy2],0<x<a,0<y<1,79 where a=10. Similarly, a body force 2ν2/(1-2ν) should be added on the right-hand side in Equation (Equation21) for this example.

Under a large noise with s=10%, and with Δx=10/200, Δy=1/1000, the numerical solutions at y=0.5 are compared with the exact ones in Figure , where α=-0.01 is used. It can be seen that the numerical solutions are quite close to the exact ones.

Figure 8. For example 4 comparing the recovered displacements and tractions with exact ones on the top side.

Figure 8. For example 4 comparing the recovered displacements and tractions with exact ones on the top side.

Under a large noise with s=10%, and with Δx=10/100, Δy=1/200, in Figure we plot the maximum numerical errors of displacements and the Neumann data of uy and vy on the top side with different values of α in the range of α[-0.1,0.1]. It can be seen that all the α in the range of α[-0.1,0.1] can cause rather small errors.

Figure 9. For example 4 comparing the numerical errors of displacements and the Neumann data on the top side with different α.

Figure 9. For example 4 comparing the numerical errors of displacements and the Neumann data on the top side with different α.

6 Extension to arbitrary plane domain

The extension of the MGPS to the inverse Cauchy problem defined in an arbitrary plane domain is possible. However, for the linear elasticity equations the process to derive the governing equations in terms of the new variables would be too lengthy, and it would be reported in other place. Here, we only consider the inverse Cauchy problem of a linear elliptic equation defined in an arbitrary plane domain as shown in Figure . We consider the following inverse Cauchy problem of the elliptic type equation:80 Δu+pu=f(x,y),(x,y)Ω,80 81 un=ϕ~(x,y),(x,y)Γ1,81 82 u=φ~(x,y),(x,y)Γ1,82 where Ω={(x,y)|0<x2+y2<ρ(θ)}, and Γ1={(x,y)|x2+y2=ρ(θ),0θπ}. We suppose that in an arbitrary plane domain as shown in Figure , the boundary can be described by a contour curve with ρ(θ) as its radius function. p is a real constant, f, φ~ and ϕ~ are given functions. If the boundary value u|Γ2, Γ2={(x,y)|x2+y2=ρ(θ),πθ2π} can be made available, then the data are completed in the whole boundary, and the solution of elliptic equation can be obtained.

Figure 10. The inverse Cauchy problem to recover data on Γ2.

Figure 10. The inverse Cauchy problem to recover data on Γ2.

Here, it is convenient to consider the problem in the polar coordinates:83 x=rcosθy=rsinθr[0,ρ],θ[0,2π],83 and write84 u(r,θ):=u(x=rcosθ,y=rsinθ).84 We consider the following transformation from r to R:85 R=rρ.85 It is obvious that when r=ρ we have R=1. So we can transform the arbitrary plane domain in terms of a small polar-coordinate (r,θ) into a unit disk in terms of a large polar-coordinate (R,θ)

Next we transform the governing equation into the large polar-coordinate (R,θ). Let86 v(R,θ)=u(r,θ).86 Through some operations we can derive87 F12vR2+F2vR+F32vθ2+F42vRθ+pv=f(R,θ),(R,θ)Ω0,87 88 v=φ(R,θ),(R,θ)Γ10,88 89 vn=ϕ(R,θ),(R,θ)Γ10,89 where Ω0={(R,θ)|R<1,0θ2π}, Γ10={(R,θ)|R=1,0θπ}, and90 F1:=1ρ21+ρρ2,90 91 F2:=1ρ2R1-ρρ-2(ρ)2ρ2,91 92 F3:=1ρ2R2,92 93 F4:=-2ρρ3R.93 The integration direction will be in the R-axis from R=1 to R=-1, and thus we consider the following variable transformation:94 v(R,θ)=eαRU(R,θ).94 From Equations (Equation83)–(Equation85) it follows that95 F12UR2+[2αF1+F2+αF9]UR+[α2F1+αF2+α2F9+p]U+F32Uθ295 96 +αF4Uθ+F42URθ=e-αRf(R,θ),96 97 U(1,θ)=e-αφ(θ),θ[0,π],97 98 UR(1,θ)=e-α[ϕ(θ)-αφ(θ)],θ[0,π],98 where99 F9:=-Rρρ.99 For simple notations we can write Equation (Equation91) as100 2UR2+F5UR+F6U+F72Uθ2+αF8Uθ+F82URθ=e-αRf(R,θ)F1,100 where101 F5:=2αF1+F2+αF9F1,F6:=α2F1+αF2+α2F9+pF1,F7:=F3F1,F8:=F4F1.101 For Equation (Equation95) we adopt the numerical method of lines to discretize the angular coordinate θ, and thus we have102 URi(R)=Vi(R),i=1,2,,m,102 103 VR1(R)=-F615U1(R)-11U2(R)+7U3(R)-U4(R)(2Δθ)2-F41V1(R)-F51U1(R)+e-αRf1(R)F11,i=1,Vri(R)=-F6iUi+1(R)-2Ui(R)+Ui-1(R)(Δθ)2-F4iVi(R)-F5iUi(R)+e-αRfi(R)F1i,i=2,,n,Vrm(R)=-F6m5Um(R)-11Um-1(R)+7Um-2(R)-Um-3(R)(2Δθ)2-F4mVm(R)-F5mUm(R)+e-αRfm(R)F1m,i=m,103 where m=n+1, and for simplicity, we consider Δθ=πn to be a uniform increment of polar angle, and θi=(i-1)Δθ. Ui(R)=U(R,θi), Vi(R)=V(R,θi), fi(R)=f(R,θi), Fji are the discretized quantities at the node of θi. The above finite differences can maintain the same second-order accuracy.

The next step is to advance the solution from the initial conditions (at the upper boundary) to the desired position R=-1. Really, Equations (Equation97) and (Equation98) have totally 2m coupled differential equations for the 2m variables Ui(R),Vi(R),i=1,2,,m, which can be numerically integrated to obtain the solutions.

6.1 Example 5

In this example we apply the MGPS to104 uxx+uyy=0,104 105 u(x,y)=x2-y2,105 where the contour of the problem domain isρ(θ)=2(0.6cos(θ)+0.3cos(2θ)-0.2)2+(0.6sin(θ))2.

Figure 11. For example 5 comparing numerical and exact solutions of the inverse Cauchy problem for Laplace equation.

Figure 11. For example 5 comparing numerical and exact solutions of the inverse Cauchy problem for Laplace equation.

The required data can be obtained from the exact solution. By using the following parameters: Δθ=π/17, h=0.5/2000, and α=0.075 for the noised case with s=0.01 we compare the numerical result obtained by the present Lie-group integration method with the exact one in Figure . The solution obtained is acceptable.

7 Extension to nonlinear inverse Cauchy problem

In order to extend the MGPS to the inverse Cauchy problem of nonlinear elasticity, we consider106 ν12ux2+2uy2+11-2ν2vxy+G1(ux,uy,vx,vy)=F1(x,y),2vx2+ν12vy2+11-2ν2uxy+G2(ux,uy,vx,vy)=F2(x,y),106 and other conditions are the same as those in Equations (Equation22)–(Equation27).

Figure 12. For example 6 of a nonlinear inverse Cauchy problem of elasticity, comparing the recovered displacements and the Neumann data with exact ones on the top side.

Figure 12. For example 6 of a nonlinear inverse Cauchy problem of elasticity, comparing the recovered displacements and the Neumann data with exact ones on the top side.

Figure 13. For example 7 of a nonlinear inverse Cauchy problem of elasticity, comparing the recovered displacements with exact ones on the top side.

Figure 13. For example 7 of a nonlinear inverse Cauchy problem of elasticity, comparing the recovered displacements with exact ones on the top side.

Upon letting107 u(x,y)=eα1yU(x,y),v(x,y)=eα2yV(x,y),107 we can derive108 ν12Ux2+2Uy2+2α1Uy+α12U+e(α2-α1)y1-2ν2Vxy+α2Vx=e-α1yF1(x,y)-G1(eα1yUx,α1eα1yU+eα1yUy,eα2yVx,α2eα2yV+eα2yVy),108 109 2Vx2+ν12Vy2+2α2Vy+α22V+e(α1-α2)y1-2ν2Uxy+α1Ux=e-α2yF2(x,y)-G2(eα1yUx,α1eα1yU+eα1yUy,eα2yVx,α2eα2yV+eα2yVy),U(x,0)=f1(x),0xa,V(x,0)=f2(x),0xa,Uy(x,0)=g1(x)-α1f1(x),0xa,109 110 Vy(x,0)=g2(x)-α2f2(x),0xa,110 111 U(0,y)=e-α1yh1(y),U(a,y)=e-α1yh2(y),0yb,111 112 V(0,y)=e-α2yh3(y),V(a,y)=e-α2yh4(y),0yb.112 Now we can apply the DQ to discretize the above equations and then apply the MGPS to integrate them.

7.1 Example 6

We consider the following analytical solutions of displacements:113 u(x,y)=σ02G(1+ν)cos(πx)y,113 114 v(x,y)=-σ04G(1+ν)1+2ν21-2νx2-2+νy2,114 and the nonlinear terms are115 G1=ux2(x,y),G2=0.115 The body forces F1(x,y) and F2(x,y) can be obtained by inserting the exact u and v into Equations (Equation101) and (Equation102).

The parameters α1=-0.1 and α2=-1 are used. Under a large noise with s=10%, and with Δx=2/50, Δy=1/2000 used in the MGPS, the numerical solutions at y=1 are compared with the exact ones in Figure . It can be seen that the present algorithm is quite robust, which can recover the data on the top side very well, even under a large noise, where the maximum errors of u and v are, respectively, 0.01926 and 0.02526. The errors of the Neumann data uy and vy are also acceptable. On the other hand, we also compute the solutions with s=0 as shown in Figure by the dashed-dotted lines, which, as compared with the numerical results under a large noise with s=10%, are more smooth without having highly oscillatory behaviour. The main reason to cause the highly oscillatory phenomenon of the noisy examples is the random noise with a large intensity we imposed in the numerical simulations.

7.2 Example 7

We use the same exact solutions as that in the above example; however, the nonlinear terms are more complex:116 G1=ux2(x,y),G2=vx2(x,y)+vy2(x,y).116 The body forces F1(x,y) and F2(x,y) can be obtained by inserting the exact u and v into Equations (Equation101) and (Equation102).

The parameters α1=-0.1 and α2=-1 are used. Under a large noise with s=10%, and with Δx=2/50, Δy=1/2000, the numerical solutions at y=1 are compared with the exact ones in Figure . It can be seen that the present algorithm is quite robust, which can recover the data on the top side very well, even under a large noise, where the maximum errors of u and v are, respectively, 0.01899 and 0.02526. For the purpose of comparison, we raise the noise intensity to s=20%, and from Figure we can observe that the numerical solutions also close to the exact ones, of which the maximum errors of u and v are, respectively, 0.034 and 0.036.

8 Conclusions

In this paper, we first transformed the two-dimensional Navier equations by using a spring-damping transform method (SDTM) through a parameter α. Then, by using a novel Lie-group integrator of the mixed group-preserving scheme (MGPS) we can recover the missing data on the top side very well for the inverse Cauchy problem in linear elasticity. Indeed, α is not a regularization parameter and it plays both a spring and a damping constant, which can slightly stabilize the governing equations and hence increase the accuracy of the numerical solutions. In general, some tests can help us to choose the suitable values of α in a certain range. Several numerical examples of the inverse Cauchy problem in linear elasticity and in the Laplace equation for arbitrary plane domain were worked out, which showed that the new numerical integration methods are applicable to the inverse Cauchy problem in linear elasticity, even for the very severely ill-posed ones. Under the overspecified data with a quite large noise the MGPS together with the SDTM was also robust enough to recover other unknown boundary data. We have extended the present algorithm to nonlinear inverse Cauchy problems of elasticity, which is insensitive to the noise perturbation. The present approach of the SDTM plus the MGPS can be extended to solve the inverse Cauchy problem in large deformation elasticity defined in an arbitrary plane domain. These issues however will be pursued in the near future.

Acknowledgments

The author highly appreciates the constructive comments from anonymous referees, which improve the quality of this paper. Taiwan’s National Science Council project NSC-102-2221-E-002-125-MY3, granted to the author, is highly appreciated. The 2011 Outstanding Research Award from NSC, the 2011 Taiwan Research Front Award from Thomson Reuters, and the 2013 Lifetime Distinguished Professor Award from NTU cause a great encouragement to the author.

References

  • Tessler A, Spangler JL. A least-square variational method for full-field reconstruction of elastic deformations in shear-deformable plates and shells. Comput. Meth. Appl. Mech. Eng. 2005;194:327–339.
  • Yeih WC, Koya T, Mura T. An inverse problem in elasticity with partially overspecified boundary conditions. I. Theoretical approach. ASME. J. Appl. Mech. 1993;60:595–600.
  • Koya T, Yeih WC, Mura T. An inverse problem in elasticity with partially overspecified boundary conditions. II. Numerical details. ASME. J. Appl. Mech. 1993;60:601–606.
  • Kozlov VA, Maz’ya VG, Fomin AV. An iterative method for solving the Cauchy problem for elliptic equations. Comput. Math. Math. Phys. 1991;31:45–52.
  • Marin L, Elliott L, Ingham DB, Lesnic D. Boundary element method for the Cauchy problem in linear elasticity. Eng. Anal. Bound. Elem. 2001;25:783–793.
  • Marin L, Lesnic D. Regularized boundary element solution for an inverse boundary value problem in linear elasticity. Commun. Numer. Meth. Eng. 2002;18:817–825.
  • Comino L, Marin L, Gallego R. An alternating iterative algorithm for the Cauchy problem in anisotropic elasticity. Eng. Anal. Bound. Elem. 2007;31:667–682.
  • Ellabib A, Nachaoui A. An iterative approach to the solution of an inverse problem in linear elasticity. Math. Comput. Simul. 2008;77:189–201.
  • Marin L, Johansson BT. A relaxation method of an alternating iterative algorithm for the Cauchy problem in linear isotropic elasticity. Comput. Meth. Appl. Mech. Eng. 2010;199:3179–3196.
  • Andrieux S, Baranger TN. An energy error-based method for the resolution of the Cauchy problem in 3D linear elasticity. Comput. Meth. Appl. Mech. Eng. 2008;197:902–920.
  • Baranger TN, Andrieux S. An optimization approach for the Cauchy problem in linear elasticity. Struct. Multidisc. Optim. 2008;35:141–152.
  • Marin L, Lesnic D. Boundary element solution for the Cauchy problem in linear elasticity using singular value decomposition. Comput. Meth. Appl. Mech. Eng. 2002;191:3257–3270.
  • Marin L. Reconstruction of boundary data in two-dimensional isotropic linear elasticity from Cauchy data using an iterative MFS algorithm. CMES: Comput. Model. Eng. Sci. 2010;60:221–245.
  • Marin L, Lesnic D. Boundary element-Landweber method for the Cauchy problem in linear elasticity. IMA J. Appl. Math. 2005;18:817–825.
  • Marin L. The minimal error method for the Cauchy problem in linear elasticity. Numerical implementation for two-dimensional homogeneous isotropic linear elasticity. Int. J. Solids Struct. 2009;46:957–974.
  • Fu Z, Chen W, Zhang C. Boundary particle method for Cauchy inhomogeneous potential problems. Inverse. Probl. Sci. Eng. 2011;20:189–207.
  • Lin J, Chen W, Wang F. A new investigation into regularization techniques for the method of fundamental solutions. Math. Compu. Simul. 2011;81:1144–1152.
  • Chen W, Gu Y. An improved formulation of singular boundary method. Adv. Appl. Math. Mech. 2012;4:543–558.
  • Gu Y, Chen W, He XQ. Singular boundary method for steady-state heat conduction in three dimensional general anisotropic media. Int. J. Heat. Mass. Transfer. 2012;55:4837–4848.
  • Liu C-S, Kuo CL. A spring-damping regularization and a novel Lie-group integration method for nonlinear inverse Cauchy problems. CMES: Comput. Model. Eng. Sci. 2011;77:57–80.
  • Liu C-S, Kuo CL, Liu D. The spring-damping regularization method and the Lie-group shooting method for inverse Cauchy problems. CMC: Comput. Mater. Continua. 2011;24:105–123.
  • Liu C-S, Chang CW. A novel mixed group preserving scheme for the inverse Cauchy problem of elliptic equations in annular domains. Eng. Anal. Bound. Elem. 2012;36:211–219.
  • Fung YC, Tong P. Classical and computational solid mechanics. Singapore: World Scientific; 2001.
  • Essers JA. New fast super-dashpot time-dependent techniques for the numerical simulation of steady flows-I. Compu. Fluids. 1980;8:351–368.
  • Engl HW, Hanke M, Neubauer A. Regularization of inverse problems. Dordrechet: Kluwer Academic Publishers; 1996.
  • Liu C-S. A dynamical Tikhonov regularization for solving ill-posed linear algebraic systems. Acta Appl. Math. 2013;123:285–307.
  • Liu C-S. Group preserving scheme for backward heat conduction problems. Int. J. Heat Mass Transfer. 2004;47:2567–2576.
  • Liu C-S. Cone of non-linear dynamical system and group preserving schemes. Int. J. Non-Linear Mech. 2001;36:1047–1068.
  • Liu C-S, Chang CW, Chang JR. Past cone dynamics and backward group preserving schemes for backward heat conduction problems. CMES: Comput. Model. Eng. Sci. 2006;12:67–81.
  • Bellman RE, Casti J. Differential quadrature and long-term integration. J. Math. Anal. Appl. 1971;34:235–238.
  • Bellman RE, Kashef BG, Casti J. Differential quadrature: a technique for the rapid solution of nonlinear partial differential equations. J. Comp. Phys. 1972;10:40–52.
  • Liu C-S, Atluri SN. A highly accurate technique for interpolations using very high-order polynomials, and its applications to some ill-posed linear problems. CMES: Comput. Model. Eng. Sci. 2009;43:253–276.
  • Shen YH, Liu C-S. A new insight into the differential quadrature method in solving 2-D elliptic PDEs. CMES: Comput. Model. Eng. Sci. 2010;71:157–178.

Appendix A

In this appendix, we provide some backgrounds of the Differential Quadrature (DQ) and the Integral Quadrature (IQ).

Bellman and Casti [Citation30], and Bellman et al. [Citation31] first proposed the Differential Quadrature (DQ) as an approximation of the derivatives of differentiable function to mimic the integral quadrature for integrable function. Here, we consider a scalar function f(x) defined in a closed interval x[a,b]. It is supposed that there are n grid points with coordinates x1=a,x2,,xn=b. The function f(x) is assumed to be differentiable at any grid point, so that its first-order derivative f(x) at any grid point xi can be approximated byA.1 f(xi)=j=1nbijf(xj).A.1 In the first approach by Bellman et al. [Citation31], the test functions are chosen as:A.2 gk(x)=xk,k=0,1,,n-1,A.2 such that we have the following algebraic equations to determine the weighting coefficients bij:A.3 j=1nbij=0,j=1nbijxj=1,j=1nbijxjk=kxik-1,k=2,,n-1.A.3 Similarly, for the integral quadrature:A.4 abf(x)dx=i=1nbif(xi),A.4 we can deriveA.5 i=1nbi=b-a,i=1nbixik=bk+1-ak+1k+1,k=1,,n-1.A.5 By inspection, we can see that the above systems are with the Vandermonde matrix as the coefficient matrix. Therefore, we can apply the technique developed by Liu and Atluri [Citation32] to solve the above system, i.e. we solveA.6 1111x1R0x2R0xn-1R0xnR0x1R02x2R02xn-1R02xnR02x1R0n-2x2R0n-2xn-1R0n-2xnR0n-2x1R0n-1x2R0n-1xn-1R0n-1xnR0n-1b1b2bkbn=b-ab2-a22R0bk+1-ak+1(k+1)R0kbn-annR0n-1,A.6 where R0>0 is a characteristic length with [a,b][-R0,R0]. Refer also Shen and Liu [Citation33]. Indeed, the above formulation is equivalent to employ the followings as the new test functions:A.7 gk(x)=xR0k,k=0,1,,n-1.A.7 By applying Equation (A1) twice we can obtain the DQ formula for the second-order differential:A.8 f(xi)=j=1naijf(xj),A.8 where aij=bikbkj.

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.