Evaluation of Inner Products of Implicitly-defined Finite Element Functions on Multiply Connected Planar Mesh Cells
Abstract.
Recent advancements in finite element methods allows for the implementation of mesh cells with curved edges. In the present work, we develop the tools necessary to employ multiply connected mesh cells, i.e. cells with holes, in planar domains. Our focus is efficient evaluation the semi-inner product and inner product of implicitly-defined finite element functions of the type arising in boundary element based finite element methods (BEM-FEM) and virtual element methods (VEM). These functions may be defined by specifying a polynomial Laplacian and a continuous Dirichlet trace. We demonstrate that these volumetric integrals can be reduced to integrals along the boundaries of mesh cells, thereby avoiding the need to perform any computations in cell interiors. The dominating cost of this reduction is solving a relatively small Nyström system to obtain a Dirichlet-to-Neumann map, as well as the solution of two more Nyström systems to obtain an “anti-Laplacian” of a harmonic function, which is used for computing the inner product. We demonstrate that high-order accuracy can be achieved with several numerical examples.
1. Introduction
Let be an open, bounded, and connected planar region with a piecewise smooth boundary . Assume the boundary is partitioned into a finite number of edges, with each edge being smooth and connected. Edges are permitted to meet at interior angles strictly between and , so that has no cusps or slits. Consider the problem of computing the semi-inner product and inner product
(1) | |||
(2) |
where are implicitly defined elements of a local Poisson space , which we define as follows. Fix a natural number , and let consist of the functions such that:
-
(a)
for , is harmonic in ;
-
(b)
for , the Laplacian is a polynomial of degree at most in ;
-
(c)
the trace is continuous;
-
(d)
the trace along any edge is the trace of a polynomial of degree at most (defined over all of ).
Note, for instance, that contains all of the polynomials of degree at most . Such subspaces of arise naturally in the context of finite element methods posed over curvilinear meshes, whose mesh cells have curved edges. Our present interest is extending the application of theses spaces to curvilinear meshes with punctured (i.e. multiply connnected) mesh cells; see Figure 1.



Such spaces of implicitly-defined functions, whether arising from curvilinear, polygonal, or more conventional tetrahedral meshes, have appeared in the literature frequently in the last several years. Many readers are likely to be familiar with Virtual Element Methods (VEM), which have gained signifcant popularity in the last decade and have a large body of recent publications [7, 8, 5, 9], some of which concern employing curvilinear mesh cells [6, 1, 11, 30]. Our approach is more closely aligned with Boundary Element Based Finite Element Methods (BEM-FEM) and Trefftz methods [10, 15, 17, 16, 18, 24, 25, 26, 29, 28, 27]. In contrast to VEM, we actually construct a basis of and do not require projections and so-called stabilization terms. Futhermore, while all computations needed for forming the finite element system do not involve any calculations on the interior of mesh cells, and despite these basis functions being implicitly-defined, we have the option of obtaining information about the interior values, gradient, and higher derivatives of basis functions on the interior of each cell, and therefore those of the finite element solution. The basic framework for our approach was proposed in [2], where we proposed methods for construction of a basis that automatically preserves conformity, and proved estimates for associated interpolation operators. Subsequently, in [23], we demonstrated that practical computation of semi-inner products and inner products of functions in are feasible whenever is simply connected (i.e. has no holes). Indeed, we showed that these volumetric integrals can be reduced to boundary integrals, thereby circumventing any need to develop 2D quadratures for the unconventional geometries present in curvilinear meshes. The goal of this work is to extend these results to the case where is multiply connected.
In Section 2, we briefly summarize how (1) and (2) may be computed in the case when is simply connected. In Section 3, we address how these calculations can be modified in order to accomodate multiply connected mesh cells. We provide a handful of numerical illustrations in Section 4, and conclude in Section 5.
2. Simply Connected Mesh Cells
Let be simply connected, and suppose that . The goal of this section is to provide an overview of some of the techniques used to compute the integrals (1) and (2). In particular, we will see that each of these volumetric integrals can be feasibly reduced to boundary integrals over . These were discussed in detail in [23], although since its publication we have made some improvements that reduce computational cost, which we present here.
2.1. The Semi-inner Product
Given , note that and are given polynomials of degree at most . Let and be polynomials of degree at most satisfying
As pointed out in [19], such polynomials and can be explicitly constructed term-by-term by observing that
(3) |
is a polynomial anti-Laplacian of for a multi-index . That is, . Note that, in practice, is obtained only by manipulation of polynomial coefficients, and poses no computational barrier. The same can be said of other operations involving polynomials, such as gradients, etc.
Since the functions
are harmonic, we have the expansion
(4) |
For the first two integrals in the final expression, the normal derivatives and may be computed using the Dirichlet-to-Neumann map discussed below. Furthermore, is clearly a polynomial of degree at most . As noted in [3], a straightforward application of the Divergence Theorem shows that
(5) |
In this fashion, we reduce the volumetric integral to readily computable boundary integrals.
2.2. A Dirichlet-to-Neumann Map
Consider the problem of determining the normal derivative of a harmonic function given its Dirichlet trace . Recall that is a harmonic conjugate of a harmonic function whenever are continuously twice differentiable on and satisfy the Cauchy-Riemann equations:
Given that is harmonic on a simply connected domain , the existence of a harmonic conjugate of is guaranteed, and is unique up to an additive constant. If is smooth, for every it holds that
(6) |
where is the fundamental solution of the Laplacian in . Supposing that the boundary is traversed counterclockwise, we let denote the unit tangent vector and denote the outward unit normal vector, so that the normal and tangential derivatives of and are related by
(7) |
from which the right-hand side in (6) can be computed. Since is unique only up to an additive constant, we impose , which we add to the left-hand side above to obtain
(8) |
In practice, we solve this integral equation numerically for on using a Nyström method, where the right-hand side is computed using the tangential derivative of , which is readily accessible from its trace . Having obtained values of the harmonic conjugate on the boundary , we may obtain its tangential derivative via numerical differentiation, which then yields values of the normal derivative . Indeed, if is a sufficiently smooth parameterization of and we define , then
(9) |
Since is periodic, a natural choice to obtain is to write a Fourier expansion and obtain an approximation of by truncating the series
(10) |
In practice, a Fast Fourier Transform (FFT) may be used on a discretization of , and an inverse FFT used on the coefficients to obtain the approximate values of .
Details of such calculations, including the case where has corners, are discussed in [22].
2.3. The Inner Product
Let and be as above. We have the expansion
(11) |
Notice that the last integral can be computed with (5), whereas the two middle integrals have the form
where is a polynomial and is harmonic. Using (3), let be a polynomial such that
then applying Green’s Second Identity, we have
(12) |
The remaining integral to be computed in (11) is the inner product of the two harmonic functions and . Toward this end, suppose that is an anti-Laplacian of the harmonic function , that is,
Then using Green’s Second Identity again yields
(13) |
The problem of determining such a , in particular its trace and normal derivative , is addressed as follows.
2.4. Anti-Laplacians of Harmonic Functions
Notice that if is an anti-Laplacian of , it holds that is biharmonic, that is, . A well-known fact (see, for example, pp. 269 of [12]) is that every biharmonic function is of the form
where are some analytic functions, denotes the real part of , denotes the complex conjugate, and we use the natural identitification of the complex plane with via . Since any anti-Laplacian of will suffice for our purposes, we will take and write
where is a harmonic function and is a harmonic conjugate of . It follows from the Cauchy-Riemann equations that
and that the gradients of and must take the form
where is a harmonic conjugate of .
Remark 2.1.
Note that the gradient of is given by
from which we may obtain the normal derivative .
Remark 2.2.
For any fixed constants and , we have that
is also an anti-Laplacian of , since is harmonic.
In order to compute and , consider the following. For the sake of illustration, we assume that the boundary is smooth, but a similar approach works when is only piecewise smooth, using some minor modifications. Given the traces of and on , we have access to the tangential derivative of via
Given a sufficiently smooth parameterization of the boundary, we define by
By the Fundamental Theorem of Calculus, we may obtain an anti-derivative via
Note that is -periodic and admits a Fourier expansion
As mentioned above, in practice we truncate this series and compute the Fourier coefficients using an FFT. Integrating termwise yields
for an arbitrary constant . In light of Remark 2.2, we may pick arbitrarily; for instance, choose . Moreover, since is -periodic, we also see that . In the computational context, we apply an inverse FFT to the coefficients in order to obtain approximate values of on the boundary. We apply an analogous procedure to obtain values of on the boundary, using the tangential derivative data
and computing an anti-derivative of
2.5. Piecewise Smooth Boundaries
In our discussion so far, we have assumed the cell boundary to be smooth, but here we will briefly address how the calculations described above can be modified when has one or more corners. Elaboration on these modifications can be found in [22].
Each edge of the mesh cell is discretized into points, including the endpoints, so that the boundary is discretized into points, with redundant endpoints being neglected. When an edge is a smooth closed contour, the boundary points are assumed to be sampled according a strongly regular parameterization of (i.e. for all and for some fixed ). In the case where terminates at a corner, we employ a Kress reparameterization, defined as follows (cf. [20]). Suppose that is a strongly regular parameterization of for , then define using
where the Kress parameter is fixed. The Kress reparameterization is not regular, with vanishing at the endpoints. Indeed, has roots at and of order , which leads to heavy sampling of the boundary near corners. This effect is amplified for larger values of .
Recall that whenever has a sufficiently smooth Dirichlet trace, we can compute the weighted tangential derivative
by using, for instance, the FFT-based approach described by (10). Replacing with a Kress reparameterization leads to difficulty in recovering the values of the tangential derivative from the weighted tangential derivative. The same can be said for the weighted normal derivative
Notice, though, that for the sake of computing the boundary integral
the weighting term appears in the Jacobian anyway, so it is natural to keep the tangential and normal derivatives in their weighted form, inlcuding the case when using a Kress reparameterization.
Note that, for the sake of effectively applying an FFT, we assume that the parameter is sampled at equispaced nodes , , , and likewise for the parameter when using a Kress reparameterization.
2.6. Summary of Simply Connected Case
Thus far, we have all the necessary tools to compute the goal integrals (1) and (2) in the case where is simply connected. It is worth reiterating that both of these volumetric integrals have been successfully reduced to contour integrals along the boundary , and there is no need for 2-dimensional quadratures as all necessary computations occur only on .
We have the option, though, of obtaining interior values of as follows. Write as above, and determine a harmonic conjugate of . Then is an analytic function, and for any fixed interior point we have Cauchy’s integral formula
(14) |
Furthermore, we can obtain interior values of by observing that
Interior values of higher derivatives, such as the components of the Hessian, can be obtained in similar fashion if so desired.
We conclude this section with a few remarks about computational complexity. Assume the boundary is parameterized and then discretized using points. The Nyström system resulting from (8) is dense, though well-conditioned, and simple linear solvers come with a computational cost . Using more sophisticated methods, such as GMRES, make an improvement, but in general will never be better than . Although even more sophisticated methods, such as those based on hierarchical matrices [14] or hierarchical semiseparable matrices [31], can reduce the computational complexity even further, for relatively small problems such as those considered here, GMRES is sufficient.
The FFT calls used for numerical differentiation have a computational cost of , and integration along using (using, say, the trapezoid rule) take operations. Operations on polynomials, such as computing anti-Laplacians, can be performed by manipulation of the coefficients and do not meaningfully contribute to the computational cost. So despite the many terms we have encountered in the expansion of the integrals (1) and (2), in practice these expansions are relatively cheap in comparison to the cost of obtaining the trace of the harmonic conjugate. Note that the latter computation need only happen once for each function considered.
Additionally, we can use the notion of trigonometric interpolation to reduce computational cost even further, as was explored in [22]. With the boundary discretized into points, we can solve the the Nyström system obtained from (8) as usual to obtain the harmonic conjugate . While performing numerical differentiation with FFT as proposed, we have the Fourier coefficients at our disposal, which allows for rapid interpolation to, say, points. We then compute the boundary integrals obtained from expanding (1) and (2) using standard 1D quadratures (e.g. the trapezoid rule, Simpson’s rule, etc.) on the larger collection of points. The heuristics presented in [22] suggest that similar levels of accuracy are achieved as would be in the case where all sampled points are used for solving the Nyström system.
In the next two sections, we address how our approach can be modified to accomodate multiply connected mesh cells.
3. Punctured Cells
We now consider the case with being multiply connected. That is, we take to be simply connected, open, bounded regions, such that:
-
(a)
for each , we have that is a proper subset of —that is, ;
-
(b)
for each , the closures of and are disjoint—that is, .
Additionally, we will require that for each , the boundary is piecewise smooth without slits or cusps. We then take to be the region
We refer to as the th hole (or puncture) of . We sometimes call the outer boundary of , and the th inner boundary. The outer boundary is assumed to be oriented counterclockwise, and the inner boundaries oriented clockwise, with the unit tangential vector , wherever it is defined, oriented accordingly. The outward unit normal is therefore always a clockwise rotation of .
In the simply connected case, we made liberal use of the notion of harmonic conjugates. However, in multiply connected domains, a given harmonic function is not guaranteed to have a harmonic conjugate, e.g. on an annulus centered at the origin. The following theorem, which is proved in [4], for example, provides a very helpful characterization of which harmonic functions have a harmonic conjugate.
Theorem 3.1 (Logarithmic Conjugation Theorem).
For each of the holes of a multiply connected domain , fix a point . Suppose that is a harmonic function on . Then there are real constants such that, for each ,
where is the real part of an analytic function. In particular, has a harmonic conjugate .
To simplify the notation in what is to come, it will be convenient to define
(15) |
In a minor notational shift from Section 2, note that, in this section, we will reserve to represent a “conjugable part” of a harmonic function , rather than treat and as independent harmonic functions as we did in the previous section.
3.1. A Dirichlet-to-Neumann Map for Punctured Cells
Our present goal is to determine the coefficients , as in the statement of Theorem 3.1, given the trace of a harmonic function . We will see that we simultaneously determine by solving an integral equation similar to (8), and thereby arrive at the Dirichlet-to-Neumann map
(16) |
Assume for now that the boundary is smooth. The case with corners is handled with Kress reparameterization, as discussed in the previous section. Our current task is to generalize the technique described by (8) in the case where is multiply connected. An alternative approach to the method we discuss here is presented in [13], which is comparable to our method in terms of cost and accuracy when all boundary edges are smooth, but does not achieve similar levels of accuracy when corners are present.
Let denote a harmonic conjugate of satisfying . Just as in the simply connected case, we have
Making the replacement
and rearranging yields
(17) |
This integral equation is underdetermined due to the additional degrees of freedom in contrast to (8). To resolve this, we multiply both sides of by the normal derivative and integrate over to obtain
Invoking Green’s Second Identity and the Cauchy-Riemann equations yields
To write this in a form more conducive to computation, observe that the Fundamental Theorem of Calculus for contour integrals implies that
since consists of closed contours. From the Product Rule, we obtain
Therefore, ought to satisfy
(18) |
In summary, we have obtained a system of equations (17) and (18) for determining the trace of on . Discretizing the boundary into points, as we did for the simply connected case, this system of equations yields a square augmented Nyström system in variables, which we may solve with the same techniques used for solving (8). The case when has corners can be handled using a Kress reparameterization, just as in the simply connected case.
3.2. Anti-Laplacians of Harmonic Functions on Punctured Cells
Next, we wish to construct an anti-Laplacian of a harmonic function
as in the statement of Theorem 3.1. It is simple to verify that
is an anti-Laplacian of , so if is an anti-Laplacian of , then we have that
(19) |
is an anti-Laplacian of . The normal derivative can be computed using
Analogous to the simply connected case, we might seek potentials of the vector fields
While and both have vanishing curls, this is not sufficient to guarantee that they are both conservative on a multiply connnected domain—a simple counterexample being
taken on a circular annulus centered at the origin.
An elementary observation from complex analysis is that an analytic function has an antiderivative if and only if and are conservative vector fields. Indeed, satisfies for and . This simple fact inspires us to decompose as , where has an antiderivative and the real part of has an anti-Laplacian which can be computed a priori. The following proposition reveals such a decomposition. The proof is inspired by that given in [4] and is unlikely to surprise a reader familiar with elementary complex analysis, but we include it for the sake of completeness.
Proposition 3.2.
Let be analytic on a multiply connected domain . Let denote a point fixed in the th hole of . Then there are complex constants such that
where has an antiderivative.
Proof.
For each , let
where is the boundary of the th hole traversed clockwise, and define
We wish to show that has an antiderivative, i.e. there is an analytic function for which . It will suffice to show that for any closed contour in . To see why, fix any point and define
where is any contour starting at and terminating at . Since this integral would be path independent, it would hold that is well defined and .
Let be a closed contour in . If is homotopic to a point, then
holds because and are analytic in simply connected open subset of . If is homotopic to , then by the Deformation Theorem we have
Note that the same conclusion holds when is oriented opposite to . Finally, if is any closed contour in that is not homotopic to a point, it holds that can be decomposed as a closed chain (cf. Theorem 2.4 in [21])
with being homotopic to and being the winding number of with respect to . Then
holds by the previous conclusion. Therefore for any closed contour in . As remarked above, this is sufficient to show that has an antiderivative. ∎
Remark 3.3.
For in the proof above, we have
(20) |
are real-valued contour integrals around the boundary of the th hole. Again, note that the inner boundary is taken to be oriented clockwise, with behaving accordingly.
Using the above results, we may decompose into one part that has an antiderivative
and another part that is a linear combination of rational functions
where we use , with being chosen when Theorem 3.1 is applied. Easy calculations verify that
satisfies , and whose normal derivative can be directly obtained from
Suppose that we have computed , using Remark 3.3 and thereby obtained as in Proposition 3.2. Since has an antiderivative, we have that the vector fields
are conservative. Let be their corresponding potentials, then it follows from the Cauchy-Riemann equations that is a harmonic conjugate of , and their Neumann data is supplied with
The solution to the Neumann problem , is unique up to an additive constant, and similarly for . Given the conclusion of Remark 2.2, we may fix these constants arbitrarily, and we make the choice to impose
Using the same techniques used to solve (8), we may determine the traces of by solving
(21) | ||||
(22) |
In summary, we have the following recipe to determine an anti-Laplacian of a given harmonic function on a multiply connected domain :
- (a)
-
(b)
Determine by computing (20).
-
(c)
Set
- (d)
-
(e)
Set
3.3. Summary of Multiply Connected Case
The strategies outlined in Section 2 for expanding the semi-inner product and inner product and reducing each term to integrals along the boundary still hold in the multiply connected case. The two primary ways in which the computations have changed in the multiply connected case is (i) the Dirichlet-to-Neumann map for harmonic functions, and (ii) obtaining an anti-Laplacian of a harmonic function.
We also note that the FFT-based method (10) can still be used to obtain from , but must be used on each component of the boundary separately. An astute reader may also notice that similar concerns would thwart our attempts to obtain with an FFT-based approach when is multiply connected, in contrast to the simply connected case.
Should we wish to obtain interior values of
which we emphasize is optional, we may proceed as follows. The values of and may be obtained through direct computation. To see how to obtain interior values of , consider the complex contour integral of along the outer boundary . Let be the positively oriented boundary of a closed disk that lies in and is centered at a fixed . Then can be decomposed into the chain
with the inner boundaries oriented clockwise. So integrating and applying Cauchy’s integral formula to yields
and rearranging gives the familiar formula
provided that we orient the boundary components properly. The same argument can be applied to show that the components of the gradient can be obtained in the interior by applying the above result to , and similarly for higher derivatives.
4. Numerical Examples



For each of the following examples, we pick explicit functions of the form
where are harmonic functions and are polynomials. While we will pick to be explicitly-defined, note that only the boundary traces and the coefficients of the polynomial Laplacians are supplied as input for the computations. Using explicitly defined functions is convenient for convergence studies, but in practice the computations will work the same for implicitly-defined functions.
Unless otherwise noted, we keep the Kress parameter fixed, as we observed that this value of the Kress parameter gave satisfactory results under a wide range of circumstances. Boundary integrals are evaluated by applying the trapezoid rule.
In each example, reference values for the and (semi-)inner products were obtained with Wolfram Mathematica. The mesh cell was defined using ImplicitRegion[], and volumetric integals were computed using NIntegrate[]. We remark that our implementation, albeit far from optimized, was significantly faster than Mathematica for computing these kinds of integrals. (In fairness, we compute these integrals as boundary integrals, whereas it seems that Mathematica implements general-purpose adaptive 2D quadrature over the volume, so perhaps the comparison in performance is unjustified.) For each reference value, we also give the error estimate that was provided by Mathematica.
Each of the numerical examples in this section is presented with Jupyter Notebook in the GitHub repository
https://github.com/samreynoldsmath/PuncturedFEM
which also contains the Python source code implementing the numerical methods we have described in this work.
Example 4.1 (Punctured Square).
error | error | wnd error | error | |
---|---|---|---|---|
4 | 1.7045e-03 | 3.5785e-02 | 2.8201e-01 | 8.3234e-03 |
8 | 3.5531e-07 | 2.6597e-04 | 1.2855e-03 | 3.9429e-05 |
16 | 1.0027e-09 | 1.1884e-06 | 3.7415e-06 | 3.3785e-07 |
32 | 3.5905e-13 | 2.3095e-09 | 1.0434e-08 | 1.9430e-09 |
64 | 1.8874e-14 | 1.6313e-12 | 6.4780e-11 | 7.0728e-12 |
Let be a unit square, and let be a disk of radius centered at . The cell under consideration is the square with the disk removed, , as depicted in the left-hand side of Figure 2. Define
Notice that can be decomposed into
with being harmonic and the polynomial having the Laplacian
Furthermore, can be decomposed as
with and . In Table 1, we report the errors in the computed approximations of , , the weighted normal derivative (wnd) of , and the trace of the anti-Laplacian . Since a harmonic conjugate is unique only up to an additive constant, we compute the error as
where is a constant minimizing the distance between the traces of and , namely
In general, an anti-Laplacian is unique only up to the addition of a harmonic function, which is much less restrictive. However, we see from Remark 2.2 that two different anti-Laplacians computed using the techniques described will differ by the addition of a linear function for some constants . It follows the that difference between and
ought to be well-modeled by , and we choose to determine optimal constants via least-squares. Upon doing so, we compute the error in with
In Table 1, we list the absolute error in the logarithmic coefficient , as well as the boundary norm of the errors in the harmonic conjugate trace , the weighted normal derivative of , and the trace of the anti-Laplacian . We observe superlinear convergence in these quantities with respect to the boundary discretization parameter . In Table 2, we provide the errors in the semi-inner product and inner product of and , where
The reference values
were obtained with Mathematica, as noted above. Notice that the convergence trends in these quantities parallels those of the intermediate quantities found in Table 1.
Punctured Square | Pac-Man | Ghost | ||||
---|---|---|---|---|---|---|
error | error | error | error | error | error | |
4 | 1.5180e-02 | 3.4040e-03 | 7.2078e-02 | 2.1955e-02 | 2.4336e+00 | 5.9408e-03 |
8 | 2.6758e-04 | 8.3812e-05 | 3.3022e-02 | 5.4798e-03 | 1.0269e-02 | 1.3086e-02 |
16 | 8.4860e-07 | 3.8993e-08 | 1.2495e-03 | 1.0159e-04 | 1.5273e-03 | 1.3783e-04 |
32 | 1.0860e-09 | 2.8398e-11 | 6.5683e-06 | 4.6050e-07 | 5.3219e-07 | 8.1747e-07 |
64 | 9.5390e-13 | 1.1036e-13 | 4.6834e-08 | 2.1726e-09 | 1.5430e-11 | 4.6189e-11 |
Example 4.2 (Pac-Man).
For our next example, we consider the Pac-Man domain , where is the sector of the unit circle centered at the origin for , , and is a disk of radius centered at . (See Figure 2, center.) The function
specified in polar coordinates is harmonic everywhere except possibly the origin for any fixed . For the choice , we have that the gradient of is unbounded near the origin; indeed, . Noting that the boundary intersects the origin, it follows that normal derivative of is also unbounded near the origin. To test whether our strategy is viable for such functions, we compute the seminorm and norm of for . The results are given in Table 2, using the reference values
Although convergence is still rapid in this case, it is less so than in the previous example, as may be expected when considering more challenging integrands, as we have here.
Example 4.3 (Ghost).
Our final example demonstrates that our method works when has more than one puncture, as well as when the boundary has edges that are not line segments or circlular arcs. The lower edge of the Ghost is the sinusoid for , the sides are vertical line segments, the upper boundary is a circular arc of radius centered at , and the inner boundaries are ellipses with and as the semi-minor and semi-major axes, respectively, with one centered at and the other at . (See Figure 2, right.) The functions we choose to integrate are
Notice that these functions have singularities in the holes of , one rational and the other logarithmic. In Table 2, we compare the computed and (semi-)inner products to the reference values
We conjecture that for , the error in the semi-inner product is significantly worse than in the other two examples because this level of boundary discretization is insufficient to fully capture the oscillatory behavior of the lower edge.
Lastly, we demonstrate the ability to obtain interior values of and in the interior of in Figure 3. All computations used to generate these values used the boundary discretization parameter . Due to the factor(s) of in the denominator of the integrand in Cauchy’s integral formula, where is a point on the boundary and is the point in the interior where we wish to evaluate , notice that the error in evaluation is considerably greater when is near the boundary. For this reason, we choose to not to perform the evaluation if is found to hold for some boundary point and a fixed . For this example, we chose .

5. Conclusion
We have seen that, given implicitly-defined functions of the type that arise in a finite element setting, we can efficiently compute the semi-inner product and inner product of and over multiply connected curvilinear mesh cells. All of the necessary computations occur only on the boundary of mesh cells, although we have the option of obtaining interior values of these functions and their derivatives using quantities obtained in the course of these calculations. Two key computations needed for our approach are (i) a Dirichlet-to-Neumann map for harmonic functions, and (ii) finding the trace and normal derivative of an anti-Laplacian of a harmonic function. We have described how both of these computations may be feasiblely accomplished on planar curvilinear mesh cells with holes. Numerical examples demonstrate superlinear convergence with respect to the number of sampled boundary points.
Funding
This work was partially supported by the National Science Foundation through NSF grant DMS-2012285 and NSF RTG grant DMS-2136228.
Code availability
Python code used for the experiments in this manuscript is publicly available at https://github.com/samreynoldsmath/PuncturedFEM
References
- [1] F. Aldakheel, B. Hudobivnik, E. Artioli, L. Beirão da Veiga, and P. Wriggers. Curvilinear virtual elements for contact mechanics. Comput. Methods Appl. Mech. Engrg., 372:113394, 19, 2020.
- [2] A. Anand, J. S. Ovall, S. E. Reynolds, and S. Weißer. Trefftz Finite Elements on Curvilinear Polygons. SIAM J. Sci. Comput., 42(2):A1289–A1316, 2020.
- [3] P. F. Antonietti, P. Houston, and G. Pennesi. Fast numerical integration on polytopic meshes with applications to discontinuous Galerkin finite element methods. J. Sci. Comput., 77(3):1339–1370, 2018.
- [4] S. Axler. Harmonic functions from a complex analysis viewpoint. Amer. Math. Monthly, 93(4):246–258, 1986.
- [5] L. Beirão da Veiga, F. Brezzi, L. D. Marini, and A. Russo. Polynomial preserving virtual elements with curved edges. Math. Models Methods Appl. Sci., 30(8):1555–1590, 2020.
- [6] L. Beirão da Veiga, A. Russo, and G. Vacca. The virtual element method with curved edges. ESAIM Math. Model. Numer. Anal., 53(2):375–404, 2019.
- [7] L. Beirão da Veiga, F. Brezzi, A. Cangiani, G. Manzini, L. D. Marini, and A. Russo. Basic principles of virtual element methods. Math. Models Methods Appl. Sci., 23(1):199–214, 2013.
- [8] L. Beirão da Veiga, F. Brezzi, L. D. Marini, and A. Russo. The hitchhiker’s guide to the virtual element method. Math. Models Methods Appl. Sci., 24(8):1541–1573, 2014.
- [9] F. Brezzi and L. D. Marini. Virtual element and discontinuous Galerkin methods. In Recent developments in discontinuous Galerkin finite element methods for partial differential equations, volume 157 of IMA Vol. Math. Appl., pages 209–221. Springer, Cham, 2014.
- [10] D. Copeland, U. Langer, and D. Pusch. From the boundary element domain decomposition methods to local Trefftz finite element methods on polyhedral meshes. In Domain decomposition methods in science and engineering XVIII, volume 70 of Lect. Notes Comput. Sci. Eng., pages 315–322. Springer, Berlin, 2009.
- [11] F. Dassi, A. Fumagalli, D. Losapio, S. Scialò, A. Scotti, and G. Vacca. The mixed virtual element method on curved edges in two dimensions. Comput. Methods Appl. Mech. Engrg., 386:Paper No. 114098, 25, 2021.
- [12] P. R. Garabedian. Partial Differential Equations. John Wiley & Sons, New York, 1964.
- [13] A. Greenbaum, L. Greengard, and G. B. McFadden. Laplace’s equation and the Dirichlet-to-Neumann map in multiply connected domains. Journal of Computational Physics, 105(1):267–278, 1993.
- [14] W. Hackbusch. Hierarchical matrices: algorithms and analysis, volume 49 of Springer Series in Computational Mathematics. Springer, Heidelberg, 2015.
- [15] H. Hakula. Resolving boundary layers with harmonic extension finite elements. Mathematical and Computational Applications, 27(4), 2022.
- [16] C. Hofreither. error estimates for a nonstandard finite element method on polyhedral meshes. J. Numer. Math., 19(1):27–39, 2011.
- [17] C. Hofreither, U. Langer, and C. Pechstein. Analysis of a non-standard finite element method based on boundary integral operators. Electron. Trans. Numer. Anal., 37:413–436, 2010.
- [18] C. Hofreither, U. Langer, and S. Weißer. Convection-adapted BEM-based FEM. ZAMM Z. Angew. Math. Mech., 96(12):1467–1481, 2016.
- [19] V. V. Karachik and N. A. Antropova. On the solution of a nonhomogeneous polyharmonic equation and the nonhomogeneous Helmholtz equation. Differ. Uravn., 46(3):384–395, 2010.
- [20] R. Kress. A Nyström method for boundary integral equations in domains with corners. Numer. Math., 58(2):145–161, 1990.
- [21] S. Lang. Complex analysis, volume 103 of Graduate Texts in Mathematics. Springer-Verlag, New York, fourth edition, 1999.
- [22] J. S. Ovall and S. E. Reynolds. A high-order method for evaluating derivatives of harmonic functions in planar domains. SIAM J. Sci. Comput., 40(3):A1915–A1935, 2018.
- [23] J. S. Ovall and S. E. Reynolds. Quadrature for implicitly-defined finite element functions on curvilinear polygons. Computers & Mathematics with Applications, 107:1–16, 2022.
- [24] S. Weißer. Residual error estimate for bem-based fem on polygonal meshes. Numer. Math., 118:765–788, 2011. 10.1007/s00211-011-0371-6.
- [25] S. Weißer. Arbitrary order Trefftz-like basis functions on polygonal meshes and realization in BEM-based FEM. Comput. Math. Appl., 67(7):1390–1406, 2014.
- [26] S. Weißer. Residual based error estimate and quasi-interpolation on polygonal meshes for high order BEM-based FEM. Comput. Math. Appl., 73(2):187–202, 2017.
- [27] S. Weißer. Anisotropic polygonal and polyhedral discretizations in finite element analysis. ESAIM Math. Model. Numer. Anal., 53(2):475–501, 2019.
- [28] S. Weißer. BEM-based Finite Element Approaches on Polytopal Meshes, volume 130 of Lecture Notes in Computational Science and Engineering. Springer International Publishing, 1 edition, 2019.
- [29] S. Weißer and T. Wick. The dual-weighted residual estimator realized on polygonal meshes. Comput. Methods Appl. Math., 18(4):753–776, 2018.
- [30] P. Wriggers, B. Hudobivnik, and F. Aldakheel. A virtual element formulation for general element shapes. Comput. Mech., 66(4):963–977, 2020.
- [31] J. Xia, S. Chandrasekaran, M. Gu, and X. S. Li. Fast algorithms for hierarchically semiseparable matrices. Numer. Linear Algebra Appl., 17(6):953–976, 2010.