Convergence of Dziuk’s semidiscrete finite element method for mean curvature flow of closed surfaces with high-order finite elements
Abstract.
Dziuk’s surface finite element method for mean curvature flow has had significant impact on the development of parametric and evolving surface finite element methods for surface evolution equations and curvature flows. However, the convergence of Dziuk’s surface finite element method for mean curvature flow of closed surfaces still remains open since it was proposed in 1990. In this article, we prove convergence of Dziuk’s semidiscrete surface finite element method with high-order finite elements for mean curvature flow of closed surfaces. The proof utilizes the matrix-vector formulation of evolving surface finite element methods and a monotone structure of the nonlinear discrete surface Laplacian proved in this paper.
1. Introduction
We consider the evolution of a closed surface under mean curvature flow, moving with velocity , where and are the mean curvature and outward unit normal vector of the surface. The surface at time can be described by
as the image of a flow map , which is a smooth embedding at every time from a given closed initial surface into , satisfying the following geometric evolution equation:
(1.1) |
where denotes the Laplace–Beltrami operator on the surface , and is the identity function satisfying for all .
Numerical approximation to mean curvature flow by parametric finite element method was first considered by Dziuk [12] in 1990. The method determines the parametrization of the unknown surface by solving partial differential equations on a surface using the surface finite element method (FEM). The evolution of the nodes determines the approximate evolving surface. This idea has had significant influence on the development of surface FEMs for many different types of geometric evolution equations, and was systematically developed to the evolving surface FEMs in [15].
However, proving convergence of Dziuk’s method for mean curvature flow of closed surfaces remains still open. For curve shortening flow, convergence of semidiscrete FEM was proved in [13]; convergence of nonlinearly implicit and linearly implicit FEMs were proved in [27] and [24], respectively. Convergence of non-parametric FEMs for mean curvature flow of graph surfaces was proved by Deckelnick & Dziuk [5, 7], but the analysis cannot be extended to closed surfaces.
Many other techniques were also developed for approximating mean curvature flow. For example, Deckelnick & Dziuk [6] has introduced an artificial tangential velocity to reformulate curve shortening flow into a non-divergence form; Barrett, Garcke & Nürnberg introduced a parametric FEM based on a different variational formulation [3] and a parametric FEM based on choosing different test functions [4]; Elliott and Fritz [20] introduced DeTurck’s trick of re-parametrization into the computation of mean curvature flow, leading to a non-degenerate parabolic system in a non-divergence form, which generalizes the reformulation of Deckelnick & Dziuk in [6].
For all the methods mentioned above, convergence of semi- and fully discrete FEMs for mean curvature flow of closed surfaces remains open. Convergence of semidiscrete FEMs was proved for curve shortening flow in [6, 20], for anistropic curve shortening flow in [14, 25], for curve shortening flow coupled with reaction–diffusion in [2, 26], and for mean curvature flow of axisymmetric surfaces in [1] based on DeTurck’s trick. The only convergence result of surface FEMs for mean curvature flow of closed surfaces was in [22] for an equivalent system of equations governing the evolution of normal vector and mean curvature, instead of for the original equation (1.1) used by Dziuk [12] and many others.
In this paper, we prove convergence of Dziuk’s semidiscrete FEM for mean curvature flow of closed surfaces for sufficiently high-order finite elements. Our proof utilizes two ideas, i.e, the matrix-vector formulation of the evolving surface FEM and the monotone structure of the finite element discrete operator associated to . The matrix-vector formulation was used in [23] in analysis of convergence of evolving surface FEMs for solution-driven surfaces; the monotone structure of the nonlinear finite element discrete operator associated to was used in [24] for analysis of curve shortening flow.
In the following, we briefly explain the two ideas in proving convergence of Dziuk’s semidiscrete FEM for mean curvature flow of closed surfaces.
Let be the vector that collects all nodes , in a triangulation of the initial surface (with finite elements of degree ). The nodal vector defines an approximate surface that interpolates at the nodes . We evolve the vector in time and denote its position at time by , which determines the approximate surface to mean curvature flow and satisfies an ordinary differential equation in the matrix-vector form (see Section 2 for details)
(1.2) |
with initial value , where and are the mass and stiffness matrices on the surface . Equation (1.2) is the matrix-vector formulation of Dziuk’s semidiscrete FEM. Correspondingly, Dziuk’s linearly implicit parametric FEM in [12] is equivalent to the following linearly implicit Euler method for (1.2):
(1.3) |
where denotes the stepsize of time discretization.
As mentioned in [2, 24], the main difficulty of numerical analysis for mean curvature flow (1.1) is the lack of full parabolicity, namely there does not exist a positive constant satisfying
(1.4) |
even if the two flow maps and are smooth and sufficiently close to each other. Similarly, if we denote by the interpolated surface of exact surface , and denote by the discrete semi-norm on defined by
with denoting the finite element function on the surface with nodal vector , then there does not exist a positive constant satisfying
(1.5) |
even if the two vectors and are sufficiently close to each other. This is the main difficulty in analysis of Dziuk’s semidiscrete FEM for mean curvature flow of closed surfaces.
We overcome this difficulty by showing the following identity (a monotone structure):
(1.6) |
where is the finite element function with nodal vector
on the intermediate finite element surface , and is the unit normal vector on . The identity (1.6) can be used to control the semi-norm of the normal component of the error. It was known for closed curves and was used to analyze convergence of Dziuk’s linearly implicit FEM for curve shortening flow in [24]. We extend this approach to mean curvature flow of closed surfaces using the matrix-vector technique.
In addition to (1.6), we also show that
(1.7) |
where can be arbitrarily small and is the discrete norm on the surface , defined by
with being the finite element function on the surface with nodal vector . Hence, the last term in (1) can be absorbed by (1.6) in the error estimation, and Gronwall’s inequality can be applied to yield an error estimate.
To illustrate the idea clearly without complicating the problem, we focus on Dziuk’s semidiscrete FEM (instead of fully discrete FEMs). As we shall see, high-order finite elements of polynomial degree are needed to bound the nonlinear terms in the error estimation, though the computations in [12] seem to work well with lower-order finite elements.
In the next section, we present rigorous description of the matrix-vector formulation of Dziuk’s semidiscrete FEM, and present the main theorem of this paper. The proof of the main theorem is presented in Section 3.
2. The main result
2.1. Basic notions and notation
If is a function defined on the surface for , then the material derivative of with respect to the parametrization is defined as
On any regular surface , for any function we denote by the surface tangential gradient as a 3-dimensional column vector. For a vector-valued function , we define , where each is a 3-dimensional column vector. We denote by the surface divergence of a vector field on , and by the Laplace–Beltrami operator applied to ; see [8] or [19, Appendix A] for these notions.
2.2. Triangulation
The given smooth initial surface is partitioned into an admissible family of shape-regular and quasi-uniform triangulations with finite elements of degree and mesh size ; see [15, 9] for the notion of admissible family of triangulations. For a fixed triangulation with mesh size , we denote by the vector that collects all nodes , in the triangulation of by finite elements of degree . The nodal vector defines an approximate surface that interpolates at the nodes .
We consider the evolution of the nodal vector and denote its value at time by , with initial condition . By piecewise polynomial interpolation on the plane reference triangle that corresponds to every curved triangle of the triangulation, the nodal vector defines a closed surface denoted by
There exists a unique finite element function of polynomial degree defined on the surface satisfying
This is the discrete flow map, which maps the initial surface to . If is a function defined on for , then the material derivative on with respect to the discrete flow map is defined by
2.3. Finite element spaces
The globally continuous finite element basis functions on the surface are denoted by
which satisfy
The pullback of from any curved triangle on to the reference plane triangle is a polynomial of degree . It is known that the basis functions , , have the following transport property (see [15])
(2.8) |
The finite element space on the surface is defined as
where each is a -dimensional column vector.
2.4. Interpolated surface and lift onto the exact surface
In order to compare functions on the exact surface with functions on the approximate surface , we introduce the interpolated surface , where denotes the nodal vector collecting the nodes , , moving along with the exact surface.
For any point there exists a unique lifted point , which was defined for linear and higher-order surface approximations in [11] and [9], respectively. The lift operator is one-to-one and onto. As a result, any function on can be lifted to a function on , defined as.
Let be the quotient between the continuous and interpolated surface measures, i.e., . Then the following inequality holds (cf. [21, Lemma 5.2]):
(2.9) |
If we denote by the standard Lagrange interpolation operator, then the lifted Lagrange interpolation approximates a function on with optimal-order accuracy (cf. [9, Proposition 2.7]), i.e.,
(2.10) |
We denote by the normal vector on and denote by its lift onto . Then approximates the normal vector on with the following accuracy (cf. [9, Propositions 2.3]):
(2.11) |
2.5. The main result
The mean curvature flow equation (1.1) can be equivalently written as
(2.12) |
Correspondingly, the semidiscrete evolving surface FEM for mean curvature flow is to find a nodal vector , , such that the corresponding approximate surface satisfies the following weak form:
(2.13) |
The mass matrix and stiffness matrix on the surface consist of block components
for , where is the identity matrix. Substituting
into (2.13) and using the transport property (2.8), we obtain the following matrix-vector form of the semidiscrete FEM:
(2.14) |
The main result of this paper is the following theorem.
Theorem 2.1.
Consider the semidiscrete FEM (2.14) with finite elements of degree . Suppose that the mean curvature flow problem (1.1) admits an exact solution that is sufficiently smooth on the time interval , and that the flow map is non-degenerate so that is a regular surface for every . Then, there exists a constant such that for all mesh sizes the following error bound holds when :
(2.15) | |||
(2.16) |
where denotes the lift of the approximate flow map from onto , and the constant is independent of .
3. Proof of Theorem 2.1
Throughout, we denote by a generic positive constant that takes different values on different occurrences.
3.1. Preliminaries
We denote by the vector consisting of the errors of numerical solutions at the nodes, and denote by
the finite element error function on surface .
Let be the maximal time such that the solution of (2.14) exists and the following inequality hold (with coefficient ):
(3.17) |
Since , it follows that , as the solution of the ordinary differential equation (2.14) exists locally in time and is continuous in time. In the following, we prove the stated error bounds for . Then we show that actually coincides with .
The smoothness and non-degeneracy of the flow map guarantees that it is locally close to an invertible linear transformation with bounded gradient uniformly with respect to . Hence, it preserves the admissibility of grids with sufficiently small mesh width . This guarantees that the triangulations determined by the nodes remain admissible uniformly for and , and the interpolated flow map and its inverse are bounded in (uniformly in ). Then (3.17) implies, through inverse inequality,
(3.18) |
Remark 3.1.
The powers of in (3.17) and (3.18) are needed to bound the nonlinear terms in the error estimation. For example, (3.17) is used to prove the boundedness of the numerical velocity in (3.3), which is used in bounding the nonlinear term in (3.3); inequality (3.18) is used in estimating the nonlinear terms in (3.4), (3.4) and (3.4). The powers of in (3.17) and (3.18) require high-order finite elements of degree in view of our error estimate (2.15) — the power of in (2.15) should be strictly bigger than in order to absorb the constant in the derivation of (3.17). This is done in (3.61) for sufficiently small .
Since and is bounded in (uniformly in ), the estimate above guarantees that the approximate flow map and its inverse are bounded in uniformly with respect to . Since deformation is the gradient of position, the boundedness of in (uniformly with respect to ) guarantees that the mesh on the approximate surface is not degenerate. Moreover, we can define an intermediate surface
(3.19) |
The estimate (3.18) also guarantees that the intermediate surface is well defined with non-degenerate mesh, with
The argument above is standard and was used in [22].
For any nodal vector with , we define a finite element function
on the intermediate surface . In particular,
is the finite element error function on the surface . As changes from to , the surface moves with velocity (with respect to ). When we simply denote
(3.20) |
which is a function on . The lift of onto is denoted by .
On the intermediate surface we define the following discrete norm and semi-norm:
(3.21) | ||||
(3.22) |
Lemma 3.1.
In the above setting, the following identities hold:
(3.23) | ||||
(3.24) |
where for two matrices and , and
Proof.
Let and denote to be the finite element function on the surface with nodal vector , where is defined in (3.19). As changes from to , the surface moves with velocity with respect to and . By using the fundamental theorem of calculus and the Leibniz formula, we have
where the last equality was essentially proved in [15, eq (2.11)]. By using the notation and in Lemma 3.1, and using the identities
we obtain (3.1). ∎
Lemma 3.2.
In the above setting, if
then for and the finite element function
satisfies the following norm equivalence:
where is an constant independent of and , with .
For sufficiently small , (3.18) guarantees that
Then Lemma 3.2 implies that
(3.25) |
By using this result in Lemma 3.1, together with the definition of the discrete and norms in (3.21)–(3.22), we obtain the following result (as in (7.7) of [22]):
(3.26) | The norms are -uniformly equivalent for , | |||
and so are the norms . |
3.2. The monotone structure
Note that the interpolated nodal vector satisfies equation (2.14) up to some defect , i.e.,
(3.27) |
where the defect satisfies the following estimate (to be proved in Section 5):
(3.28) |
Subtracting (3.27) from (2.14), we obtain the error equation
(3.29) |
By using Lemma 3.1, we have
(3.30) |
where is defined in Lemma 3.1, and we have used the identity
with denoting the unit normal vector on (thus ).
Note that is a symmetric projection matrix satisfying
By using the properties above and the expression of , we furthermore reduce (3.2) to
(3.31) |
Then we use the following lemma, of which the proof is presented in Section 4.
Lemma 3.3.
In the above setting, the following identity holds:
(3.32) |
3.3. Error estimation
Let be the velocity of the exact surface , and let be the velocity of the exact surface at the th interpolation node. We define
which is the interpolation of onto . Let be the lift of onto the exact surface , and denote
which is a finite element function on the surface .
Let be the velocity of the approximate surface , and let
Then the nodal vector associated to the finite element function is , and by using the norm equivalence in Lemma 3.2,
(3.36) |
where the last inequality uses (3.17) and , and the second to last inequality can be proved as follows. Testing (3.29) with , we obtain
(3.37) |
By using Lemma 3.1, we have
(3.38) |
By denoting , we have
(3.39) |
By using the estimate (3.28) for the defect , we have
(3.40) |
Substituting (3.3)–(3.40) into (3.37) and choosing , we obtain
This proves the second to last inequality of (3.3).
Recall that the finite element function on with the nodal vector is denoted by . By using (3.3), the first term on the right-hand side of (3.3) can be estimated as follows:
(3.41) |
where the norm equivalence in (3.26) is used.
The third term on the right-hand side of (3.3) satisfies
(3.42) |
We decompose the second term on the right-hand side of (3.3) into several terms as follows:
(3.43) |
The purpose of transforming from to (namely to be able to replace with ) is to perform integration by parts on the last term of (3.3). This would yield , which is the only term that contains the partial derivative of on the right-hand side. The term can be furthermore converted to (which can be absorbed by the left-hand side of (3.3)) after transforming back to , as shown in the following estimates.
The last term in (3.3) can be estimated as follows. Using the integration by parts formula (cf. [16, Section 2.3])
we have
Recall that denotes the normal vector on and is the lift of onto . By introducing to be the finite element interpolation of , and denoting by the lift of to the surface , the inequality above furthermore implies that
where the last inequality uses the interpolation error estimate (2.10)–(2.11). By using the norm equivalence and in Lemma 3.2, and using the inverse inequality of finite element functions, we obtain from the above inequality
where is defined as the finite element function on with the same nodal vector as . Substituting this into (3.3) yields
(3.44) |
3.4. Estimation of ,
(3.45) |
where the last inequality uses the properties , and the fact that the surface moves with velocity with respect to . By using the identity (cf. [18, Lemma 2.6])
(3.46) |
we find that
(3.47) |
Let denote the lift of onto . By using (2.9) we have
(3.48) |
where we have used inverse inequality in the second to last inequality.
For the exact surface , we denote by the signed distance from to , defined by
Let be the Weingarten matrix on . Then the following identity holds (for example, see [17, Remark 4.1]):
where , with denoting the normal vector on . Hence, denoting , we have
(3.49) |
where the second to last inequality uses estimate (2.11) in estimating , and uses (see [21, Lemma 5.2]). For sufficiently small , the inequality above furthermore implies, via using the triangle inequality,
(3.50) |
where we have used the norm equivalence between and as shown in Lemma 3.2. By using the two results above, we have
where we have used inverse inequality in the second to last inequality.
Since the lifted Lagrange interpolation has optimal-order accuracy in approximating , as shown in (2.10), it follows that
Recall that is the finite element interpolation of onto and is the lift of onto the surface . By using (2.9) and (3.4), we can estimate similarly as , i.e.,
(3.51) |
where we have used inverse inequality in the second to last inequality.
Recall that is finite element function on with the same nodal vector as the interpolated finite element function . Since , as defined in (3.20), it follows that
where the last equality uses the facts that and the surface moves with velocity with respect to . By substituting the identity (cf. [18, Lemma 2.6]),
(3.52) |
into the above expression of , we obtain
(3.53) |
Let be the restriction of the identity function to the surface , and note that the surface has parametrization defined on . Hence, has parametrization defined on . Let be a local parametrization of the surface in a chart, and let . Then
Since the exact surface is non-degenerate, we have . Hence,
This implies that
(3.54) |
where the second to last inequality is obtained from (3.50). By using the estimate above, we have
(3.55) |
Finally,
(3.56) |
Substituting the estimates of , , into (3.44), we have
(3.57) |
Remark 3.2.
Then, substituting (3.3)–(3.42) and (3.57) into (3.3), we obtain
(3.58) |
where can be an arbitrary positive number between and . By choosing and integrating the inequality above in time, we have
(3.59) |
which holds for all . Since is equivalent to , as explained in (3.26), applying Gronwall’s inequality yields
Hence,
(3.60) |
When and sufficiently small, this implies
(3.61) |
If then the inequality above furthermore implies that the solution can be extended to time for some sufficiently small such that (3.17) holds. The maximality of for (3.17) implies that .
4. Proof of Lemma 3.3
In this section we prove Lemma 3.3, which is used in the proof of Theorem 2.1. Note that is the boundary of a bounded Lipschitz domain. We first prove the result for a smooth surface and then extend it to a general Lipschitz surface through approximating it by smooth surfaces.
Proposition 4.1.
Let be a bounded, closed and smooth surface and let . Then
(4.62) |
Proof.
Proposition 4.2.
If is the boundary of a bounded Lipschitz domain , then for the following identity holds:
(4.64) |
Proof.
In the following, we show that there exists a sequence of smooth functions such that converges to in as , and a sequence of smooth domains with smooth boundary such that as . By using the result of Proposition 4.1, we have
(4.65) |
By taking in the equality above, we shall prove the following result:
(4.66) |
This would prove the desired result for the smooth function . Since in , letting in (4.66) yields the desired result (4.64).
First, we consider a partition of unity , , such that in a neighborhood of and each has compact support in an open ball in which the surface can be represented by a Lipschitz graph after a rotation :
(4.67) | |||
(4.68) | |||
(4.69) |
where is a Lipschitz continuous function on , which is a bounded domain in . Hence,
For the Lipschitz domain , there exists a sequence of domains , , with smooth boundary such that as in the following sense (see [10, Theorem 5.1]):
(4.70) |
where , , is a sequence of functions converging to strongly in both and for all , and converges to weakly∗ in ( is bounded in as ).
Next, on the two-dimensional region , we define and
(4.71) |
Then is a parametrization of and . We can approximate in by a sequence of smooth functions with compact supports inside . These functions have natural extensions to , i.e.,
(4.72) |
where is a one-dimensional smooth cut-off function which satisfies
(4.73) |
Then we can define a smooth function (with compact support in ) that approximates in , i.e.,
(4.74) |
By choosing a sufficiently small , the extended functions have compact supports in . Since is a parametrization of , it follows that “ converges to in ” if and only if “ converges to in ”. In view of the definitions in (4.71)–(4.72) and (4.74), we have
(4.75) |
Since converges to in as for arbitrary (see the statement below (4.70)), and converges to in for all (this is how is defined), from (4) it is straightforward to verify that converges to in . As a result, converges to in . Therefore,
is a sequence of functions in that converges to in as .
Finally, we prove that taking in (4.65) would yield (4.66). This would complete the proof of Proposition 4.2. To this end, we consider the decomposition
(4.76) |
and prove the following two results:
(4.77) | ||||
(4.78) |
Let . Then is a parametrization of the surface after a rotation by . By using this parametrization, the left-hand side of (4.77) can be written as
(4.79) | ||||
where is the inverse matrix of the Riemannian metric tensor , i.e.,
Since converges to in as for all , it follows that converges to in for all . Furthermore, since
is bounded from both below and above (because is bounded in as ), it follows that the inverse matrix also converges, i.e.,
(4.80) | converges to in for all as . |
Note that
where denotes the th row of . Since for fixed and converges to in for all as , it follows that
(4.81) | converges to in for all as . |
Since is smooth and converges to in as , it follows that
(4.82) | converges to in as . |
Then, substituting (4.80), (4.81) and (4.82) into the right-hand side of (4.79) and taking limit , we obtain (4.77). The proof of (4.78) is similar and omitted.
5. Proof of the defect’s estimate (3.28)
In this section we prove (3.28), which is used in the proof of Theorem 2.1. We rewrite equation (1.1) into
(5.83) |
Let be a finite element function on the interpolated surface , and let be the lift of onto the exact surface . Then, testing (5.83) by , we obtain
(5.84) |
This can be furthermore written as
(5.85) |
where is the unique finite element function determined by the relation
In the matrix-vector form, (5.85) can be equivalently written as
(5.86) |
with being the nodal vector of the finite element function .
Note that on and on , where and are the velocity of the surfaces and , respectively. In particular, is the Lagrange interpolation of . Hence, by using (2.10) and (2.9),
Let and be the identity function restricted to and , respectively, and let be the lifted function on . Then
where the second to last inequality again uses [21, Lemma 5.2]. This proves that
In the matrix-vector form, this can be equivalently written as
Hence, by choosing in the inequality above, we obtain
This proves the defect’s estimate (3.28).
6. Concluding remarks
The main contribution of this paper is the discovery of the structure (1.6) and its application to proving the convergence of Dziuk’s semidiscrete FEM for mean curvature flow of closed surfaces with sufficiently high-order finite elements.
The following additional difficulty would appear in the analysis of linearly implicit time discretisation:
(6.87) |
is no longer in the form of the left-hand side of (1.6) due to the shift of superscript indices. Hence, additional terms would appear in converting (6.87) to the form of the left-hand side of (1.6). Those additional terms may be bounded by using the approach in [24] under a certain grid-ratio condition.
It is straightforward to verify that both (3.2) and Proposition 4.1 can be extended to higher dimensions, i.e., for mean curvature flow of -dimensional hypersurfaces in with . As a result, the monotone structure and the convergence proof can be generalised to this case. However, the monotone structure of mean curvature flow of two-dimensional surfaces in higher codimension is not obvious from the current proof, and therefore the convergence of evolving surface FEMs in this case still remains open.
Convergence of Dziuk’s semidiscrete FEM with low-order finite elements, as well as the parametric FEMs of Barrett, Garcke & Nürnberg [3, 4], remain open for mean curvature flow of closed surfaces. Efficient numerical methods for the non-divergence parabolic system constructed from DeTurck’s trick in [20], allowing singularity to appear in the numerical simulation of closed surfaces, is still challenging.
Acknowledgement
I would like to thank Prof. Christian Lubich for reading the manuscript and providing many valuable comments and suggestions.
References
- [1] J. W. Barrett, K. Deckelnick, and R. Nürnberg. A finite element error analysis for axisymmetric mean curvature flow. arXiv:1911.05398, 2019.
- [2] J. W. Barrett, K. Deckelnick, and V. Styles. Numerical analysis for a system coupling curve evolution to reaction diffusion on the curve. SIAM J. Numer. Anal., 55(2):1080–1100, 2017.
- [3] J. W. Barrett, H. Garcke, and R. Nürnberg. On the parametric finite element approximation of evolving hypersurfaces in . J. Comput. Phys., 227(9):4281–4307, 2008.
- [4] J. W. Barrett, H. Garcke, and R. Nürnberg. Numerical approximation of gradient flows for closed curves in . IMA J. Numer. Anal., 30(1):4–60, 2010.
- [5] K. Deckelnick and G. Dziuk. Convergence of a finite element method for non-parametric mean curvature flow. Numer. Math., 72(2):197–222, 1995.
- [6] K. Deckelnick and G. Dziuk. On the approximation of the curve shortening flow. In Calculus of variations, applications and computations (Pont-à-Mousson, 1994), volume 326 of Pitman Res. Notes Math. Ser., pages 100–108. Longman Sci. Tech., Harlow, 1995.
- [7] K. Deckelnick and G. Dziuk. Error estimates for a semi-implicit fully discrete finite element scheme for the mean curvature flow of graphs. Interfaces Free Bound., 2(4):341–359, 2000.
- [8] K. Deckelnick, G. Dziuk, and C. M. Elliott. Computation of geometric partial differential equations and mean curvature flow. Acta Numerica, 14:139–232, 2005.
- [9] A. Demlow. Higher–order finite element methods and pointwise error estimates for elliptic problems on surfaces. SIAM J. Numer. Anal., 47(2):805–807, 2009.
- [10] P. Doktor. Approximation of domains with Lipschitzian boundary. Časopis pro pěstování matematiky, 101(3):237–255, 1976.
- [11] G. Dziuk. Finite elements for the Beltrami operator on arbitrary surfaces. Partial differential equations and calculus of variations, Lecture Notes in Math., 1357, Springer, Berlin, pages 142–155, 1988.
- [12] G. Dziuk. An algorithm for evolutionary surfaces. Numer. Math., 58(1):603–611, 1990.
- [13] G. Dziuk. Convergence of a semi-discrete scheme for the curve shortening flow. Math. Models Methods Appl. Sci., 4(4):589–606, 1994.
- [14] G. Dziuk. Discrete anisotropic curve shortening flow. SIAM J. Numer. Anal., 36(6):1808–1830, 1999.
- [15] G. Dziuk and C. M. Elliott. Finite elements on evolving surfaces. IMA J. Numer. Anal., 27(2):262–292, 2007.
- [16] G. Dziuk and C. M. Elliott. Finite element methods for surface PDEs. Acta Numerica, 22:289–396, 2013.
- [17] G. Dziuk and C. M. Elliott. -estimates for the evolving surface finite element method. Math. Comp., 82(281):1–24, 2013.
- [18] G. Dziuk, D. Kröner, and T. Müller. Scalar conservation laws on moving hypersurfaces. Interfaces Free Bound., 15(2):203–236, 2013.
- [19] K. Ecker. Regularity theory for mean curvature flow. Springer, 2012.
- [20] C. M. Elliott and H. Fritz. On approximations of the curve shortening flow and of the mean curvature flow based on the DeTurck trick. IMA J. Numer. Anal., 37(2):543–603, 2017.
- [21] B. Kovács. High-order evolving surface finite element method for parabolic problems on evolving surfaces. IMA J. Numer. Anal., 38(1):430–459, 2018.
- [22] B. Kovács, B. Li, and C. Lubich. A convergent evolving finite element algorithm for mean curvature flow of closed surfaces. Numer. Math., 143:797–853, 2019.
- [23] B. Kovács, B. Li, C. Lubich, and C. Power Guerra. Convergence of finite elements on an evolving surface driven by diffusion on the surface. Numer. Math., 137(3):643–689, 2017.
- [24] B. Li. Convergence of Dziuk’s linearly implicit parametric finite element method for curve shortening flow. SIAM J. Numer. Anal., 58(4):2315–2333, 2020.
- [25] P. Pozzi. Anisotropic curve shortening flow in higher codimension. Math. Meth. Appl. Sci., 30(11):1243–1281, 2007.
- [26] P. Pozzi and B. Stinner. Curve shortening flow coupled to lateral diffusion. Numer. Math., 135:1171–1205, 2017.
- [27] R. Rusu. Numerische analysis für den Krümmungsfluß und den Willmorefluß. PhD Thesis, University of Freiburg, Freiburg, 2006.