Mesh Refinement for Inverse Problems with Finite Element Models

Abstract

Many inverse problems arising in experimental mechanics involve solutions to partial differential equations in the forward problem, typically using finite element methods for those solutions. Given that iterative solutions to the inverse problem then involve repeated evaluations of the finite element model, it is useful to carefully consider the mesh to be used and its effect on the trade-off between accuracy and computational cost of the solution. We show that approximation theory can be applied directly to the inverse problem, not merely to the finite element model contained in the forward problem, to give bounds for the error made by using a given mesh to approximate the solution to the partial differential equation. Adaptive mesh refinement allows to focus computational effort on goals set by the quantities of interest in the inverse problem, rather than on the overall accuracy of the solution to the forward problem.

Introduction

Inverse problems —in the sense of problems that, by Hadamard’s criteria[1], are ill-posed whereas the corresponding forward problem is well-posed— occur in many different guises in experimental mechanics, although they are not always recognized as such. For example, the whole family of full-field measurements, including methods based on photoelasticity, moire fringes, in-terferometry, digital image correlation, et cetera, represent ill-posed inverse problems of determining a continuum field variable from a large but finite set of measurements. Conversely, the name inverse problems is widely used in experimental mechanics to refer to parameter identification problems which are not necessarily ill-posed. Determining linear elastic material parameters from mechanical tests with full-field measurements is well-posed if the experiment is not too poorly designed, but using those same measurements to characterize nonlinear constitutive behaviour such as elastoplastic deformation[2] can be ill-posed.


The inherent ill-posedness of full-field measurements is typically resolved without much fuss by implicitly or explicitly taking the spatial resolution of the measurement method into account in the conceptual definition of the field that is measured. Often, the inverse problem is not even considered at all, and the device is simply assumed to provide a means for sampling the continuum field. Nevertheless, there is a growing awareness that conversions between different representations of a continuum field can be problematic[3], and that there are significant advantages to considering the entire measurement process in the inverse methods[4, 5].

Often, the forward problems involve the solution of elliptic partial differential equations, such as the equations of continuum mechanics or the Fourier heat flow equation. These give rise to a particular form of ill-posedness related to Saint Venant’s principle: with increasing distance between load and response, the effects of local details in loading become increasingly difficult to discern. Therefore it is pointless to try to extract too much detail from an inverse method, and it makes sense to search instead for a regularized inverse of the forward problem. For numerical calculations, the forward problem is typically discretized using a finite element approximation, which then raises the question, if it is not useful to extract fine details from the calculation, whether it is necessary to calculate those details at all. In other words, does refining the finite element mesh have the same effect on the accuracy of the solution to the inverse problem as on the accuracy of the solution to the forward problem?

Forward problem

Using the terminology of heat diffusion, the problem we consider is that of reconstructing the heat sources based on a number of temperature measurements. This problem might arise when using infrared thermography for non-destructive testing, or in postprocessing thermoelastic stress analysis data[6]. More importantly, the diffusion equation with different terminology describes many different physical processes. The heat equation serves here as a prototype elliptic partial differential equation; the methods used and the resulting conclusions will be more generally applicable.

Our model for the transfer of heat is Poisson’s equation

tmp11283_thumb

where f is the heat source field, a is the thermal conductivity and u is the temperature field.

With suitable additional assumptions [7, 8], there exists a unique solution to this problem for eachtmp11284_thumbbe the solution operator to the problem, such thattmp11285_thumbbe the finite element solution operator with mesh density parameter h, and let the corresponding FE solution be denotedtmp11286_thumbAssuming full regularity and a first order FE approximation, the standard finite element H1 error estimate then gives

tmp11290_thumb

and the Aubin-Nitsche lemma gives

tmp11291_thumb

The temperature measurements are modeled as linear functionals on the temperature field. That is, the measurement is a vector m such that

tmp11292_thumb

For technical reasons, the measurement functionals are assumed to be square integrable, i.e.

tmp11293_thumb

Inverse problem

In the inverse problem, a reconstruction fr of the heat source field f is determined from a measurement m. The relation in equation (4) is not uniquely invertible, but a reconstruction is possible with additional assumptions. Often in practice there is some a priori knowledge of the expectation value of f, as well as of its covariance operator. The expectation value of f is denoted by f, and the covariance operator by b(-, •). In addition, since we are interested in enforcing the smoothness of the reconstruction, it will be sought from the Sobolev space H1 (ft).

With these assumptions, then for instance by using the method of statistical inversion or generalized Tikhonov regulariza-tion, a reconstruction can obtained as the following minimization problem [9]

tmp11294_thumb

The termtmp11295_thumbin the minimization hence enforces that f is close to f and that the difference f — f is smooth. This minimization problem can equivalently be written in variational form as: findtmp11296_thumbsuch that

tmp11299_thumb

where

tmp11300_thumb

and

tmp11301_thumb

Given that the covariance operator b(-, •) is (i) continuous and (ii) coercive in

tmp11302_thumbtmp11303_thumb

it follows that the bilinear form a(-, •) is continuous and coercive. Due to the Lax-Milgram lemma, the problem in equation (6) therefore has a unique solution [7, 8]. However, because both a( , •) and /(•) contain the solution operator K—1 of the forward problem, they cannot be computed with finite resources. To obtain a computable problem close to the original problem, we modify a(-, •) and /(•) to use the FE solution operatortmp11304_thumbinstead.

The modified problem is then: findtmp11305_thumbsuch that

tmp11308_thumb

where

tmp11309_thumb

and

tmp11310_thumb

Like , •), the bilinear form &(•, •) is also continuous and coercive in the space H1 (ft), and again by the Lax-Milgram lemma, the variational problem in equation (11) has a unique solution. This problem can then be discretized using Galerkin’s method [7], by choosing a suitable finite subspacetmp11311_thumband solving the variational equation in that space, to findtmp11312_thumbsuch that

tmp11315_thumb

For practical reasons, we choose to build our space Fh on the same mesh, using the same polynomial order in the shape functions, as in the finite element solution of the forward problem.

Error Analysis

The error of the approximate solutiontmp11316_thumb can be broken down into two components

tmp11317_thumb

The first part, the termtmp11318_thumb, is due to using the FE approximation of the forward problem instead of the exact one. We refer to this error source as the consistency error. The second part,tmp11319_thumbis the discretization error of seeking the solution only from a finite subspace.

To estimate the consistency error, we define two operators

tmp11322_thumb

and

tmp11323_thumb

The operator Ei(•) can be estimated as

tmp11324_thumb

where the last inequality is due to the Aubin-Nitsche L2 error estimate of the finite element solution (3). The operator Ea(•, •) can be estimated similarly to give

tmp11325_thumb

Using the definitions of fr and fr, we see that for anytmp11326_thumbit holds that

tmp11327_thumb

Then, due to the coerciveness of a(, •) and the estimates for Ei(•) and Ea(-, ), we can estimate the consistency error as

tmp11328_thumb

Combining the constants and estimatingtmp11329_thumbwith a stronger norm gives

tmp11330_thumb

which finally gives

tmp11331_thumb

To estimate the discretization error, we note that since a(-, •) is continuous and coercive in the space „tmp11332_thumbCea’s lemma [7, 8] holds true. It states that, up to a constant factor Cs, the discrete solutiontmp11333_thumbis the best approximation oftmp11334_thumbi.e. for any

tmp11339_thumb

Since we want to compare the error to the norm of fr rather than fr, we do the following

tmp11340_thumb

Now, since we assume that fr has full-regularity and because Fh is a first order finite element space, there exists an interpolating function g e Fh such that

tmp11341_thumb

In addition, we can estimate the termtmp11342_thumbwith the consistency error estimate in equation (23), to obtain

tmp11343_thumb

Combining the constants and assuming h < 1 gives

tmp11344_thumb

An L2 error estimate follows from the Aubin-Nitsche lemma [7] applied to the variational problem in equation (11). It then holds that

tmp11345_thumb

Combining the consistency error and the discretization error, we get the following estimates

tmp11346_thumb

and

tmp11347_thumb

Numerical experiments

Numerical tests were run in the rectangular domaintmp11348_thumbThe measurements were average temperatures over discs of radius 0.1 (the average is to ensure that they are L2 continuous), which were scattered in a uniform 16×4 grid. The measurement data vector m was simulated on a highly refined mesh using a smooth temperature source field. The bilinear form b( , ) was chosen as

tmp11350_thumb

Figure 1 shows three numerical solutions of this problem with different mesh resolutions. The larger features are already present on the coarsest mesh, whereas small details first appear on the second mesh and are further refined on the densest mesh.

The behavior of H1 and L2 errors against the mesh density parameter h are shown in figure 2. Two straight lines are fitted to the individual computations to derive the convergence rate. For the L2 error, we see convergence in the order of h187, which, is reasonably close to h2 as the analysis predicted. The H1 error also shows a convergence rate quite close to the predicted rate. These errors were computed against a highly refined numerical solution, which acted as the exact solution.

Figure 1 shows three numerical solutions of this problem with different mesh resolutions. The larger features are already present on the coarsest mesh, whereas small details first appear on the second mesh and are further refined on the densest mesh.

The behavior of H1 and L2 errors against the mesh density parameter h are shown in figure 2. Two straight lines are fitted to the individual computations to derive the convergence rate. For the L2 error, we see convergence in the order of h187, which, is reasonably close to h2 as the analysis predicted. The H1 error also shows a convergence rate quite close to the predicted rate. These errors were computed against a highly refined numerical solution, which acted as the exact solution.

Approximate solutions of the inverse problem with an increasing mesh resolution

Figure 1: Approximate solutions of the inverse problem with an increasing mesh resolution

L2 and H1 errors of the approximate solution with different values of h. The convergence rates correspond well with the rates estimated a priori.

Figure 2: L2 and H1 errors of the approximate solution with different values of h. The convergence rates correspond well with the rates estimated a priori.

Conclusions

By formulating the inverse problem as a minimization problem similar to the weak form of the finite element problem, the a priori error analysis of the inverse method can be carried out similarly to the error analysis for finite element methods. Errors arise both due to discretizing the quantity of interest and due to using a FE approximation of the forward problem.

Using the same approximation order, the consistency error —i.e., the error due to replacing the exact forward problem with a finite element approximation— is at most the same order of magnitude as the inverse problem discretization error. In fact, in the "natural" H1 norm, it is a full order of magnitude smaller. This suggests that the accuracy of the forward problem is not as critical as usually thought. One should choose the discretization with just enough resolution so that expected features of the reconstruction can be represented.

Next post:

Previous post: