Universität Stuttgart

Similar documents
AND BARBARA I. WOHLMUTH

An Iterative Substructuring Method for Mortar Nonconforming Discretization of a Fourth-Order Elliptic Problem in two dimensions

Universität Stuttgart

ETNA Kent State University

Universität Stuttgart

Discontinuous Galerkin Methods

Simple Examples on Rectangular Domains

Numerical Solutions to Partial Differential Equations

Non-Conforming Finite Element Methods for Nonmatching Grids in Three Dimensions

PhD dissertation defense

Multigrid Methods for Saddle Point Problems

INTRODUCTION TO FINITE ELEMENT METHODS

Numerical Solutions to Partial Differential Equations

A note on accurate and efficient higher order Galerkin time stepping schemes for the nonstationary Stokes equations

Scientific Computing WS 2017/2018. Lecture 18. Jürgen Fuhrmann Lecture 18 Slide 1

Scientific Computing WS 2018/2019. Lecture 15. Jürgen Fuhrmann Lecture 15 Slide 1

A FETI-DP Method for Mortar Finite Element Discretization of a Fourth Order Problem

1 Discretizing BVP with Finite Element Methods.

Domain Decomposition Algorithms for an Indefinite Hypersingular Integral Equation in Three Dimensions

Chapter 5 A priori error estimates for nonconforming finite element approximations 5.1 Strang s first lemma

A local-structure-preserving local discontinuous Galerkin method for the Laplace equation

b i (x) u + c(x)u = f in Ω,

A Framework for Analyzing and Constructing Hierarchical-Type A Posteriori Error Estimators

PREPRINT 2010:25. Fictitious domain finite element methods using cut elements: II. A stabilized Nitsche method ERIK BURMAN PETER HANSBO

Basic Concepts of Adaptive Finite Element Methods for Elliptic Boundary Value Problems

Chapter 3 Conforming Finite Element Methods 3.1 Foundations Ritz-Galerkin Method

The mortar element method for quasilinear elliptic boundary value problems

Yongdeok Kim and Seki Kim

A gradient recovery method based on an oblique projection and boundary modification

Hamburger Beiträge zur Angewandten Mathematik

Two-scale Dirichlet-Neumann preconditioners for boundary refinements

PREPRINT 2010:23. A nonconforming rotated Q 1 approximation on tetrahedra PETER HANSBO

arxiv: v1 [math.na] 1 May 2013

Technische Universität Graz

A multipoint flux mixed finite element method on hexahedra

A Mixed Nonconforming Finite Element for Linear Elasticity

Lecture Note III: Least-Squares Method

R T (u H )v + (2.1) J S (u H )v v V, T (2.2) (2.3) H S J S (u H ) 2 L 2 (S). S T

A posteriori error estimation for elliptic problems

The All-floating BETI Method: Numerical Results

Optimal multilevel preconditioning of strongly anisotropic problems.part II: non-conforming FEM. p. 1/36

The Mortar Wavelet Method Silvia Bertoluzza Valerie Perrier y October 29, 1999 Abstract This paper deals with the construction of wavelet approximatio

Finite Elements. Colin Cotter. February 22, Colin Cotter FEM

Domain Decomposition Methods for Mortar Finite Elements

Adaptive Finite Element Methods Lecture Notes Winter Term 2017/18. R. Verfürth. Fakultät für Mathematik, Ruhr-Universität Bochum

A LAGRANGE MULTIPLIER METHOD FOR ELLIPTIC INTERFACE PROBLEMS USING NON-MATCHING MESHES

Multigrid Methods for Elliptic Obstacle Problems on 2D Bisection Grids

Overlapping Schwarz preconditioners for Fekete spectral elements

A Robust Preconditioner for the Hessian System in Elliptic Optimal Control Problems

A Balancing Algorithm for Mortar Methods

A BIVARIATE SPLINE METHOD FOR SECOND ORDER ELLIPTIC EQUATIONS IN NON-DIVERGENCE FORM

2 Two-Point Boundary Value Problems

Spectral element agglomerate AMGe

OVERLAPPING SCHWARZ ALGORITHMS FOR ALMOST INCOMPRESSIBLE LINEAR ELASTICITY TR

Basics and some applications of the mortar element method

On an Approximation Result for Piecewise Polynomial Functions. O. Karakashian

arxiv: v1 [math.na] 29 Feb 2016

PARTITION OF UNITY FOR THE STOKES PROBLEM ON NONMATCHING GRIDS

A Fast Fourier transform based direct solver for the Helmholtz problem. AANMPDE 2017 Palaiochora, Crete, Oct 2, 2017

c 2008 Society for Industrial and Applied Mathematics

Comparison of V-cycle Multigrid Method for Cell-centered Finite Difference on Triangular Meshes

SPECTRAL PROPERTIES OF THE LAPLACIAN ON BOUNDED DOMAINS

Remarks on the analysis of finite element methods on a Shishkin mesh: are Scott-Zhang interpolants applicable?

arxiv: v1 [math.na] 11 Jul 2011

Overlapping Schwarz Preconditioners for Spectral. Problem in H(curl)

Numerical Solutions to Partial Differential Equations

Thomas Apel 1, Ariel L. Lombardi 2 and Max Winkler 1

Technische Universität Graz

Variational Formulations

LECTURE 1: SOURCES OF ERRORS MATHEMATICAL TOOLS A PRIORI ERROR ESTIMATES. Sergey Korotov,

The Plane Stress Problem

Maximum norm estimates for energy-corrected finite element method

The Mortar Boundary Element Method

Mixed Finite Element Methods. Douglas N. Arnold, University of Minnesota The 41st Woudschoten Conference 5 October 2016

A u + b u + cu = f in Ω, (1.1)

Hybridized Discontinuous Galerkin Methods

Institut für Mathematik

A Finite Element Method for the Surface Stokes Problem

Finite Elements. Colin Cotter. January 15, Colin Cotter FEM

Multispace and Multilevel BDDC. Jan Mandel University of Colorado at Denver and Health Sciences Center

Subdiffusion in a nonconvex polygon

Numerische Mathematik

Chapter 6 A posteriori error estimates for finite element approximations 6.1 Introduction

Technische Universität Graz

Chapter Two: Numerical Methods for Elliptic PDEs. 1 Finite Difference Methods for Elliptic PDEs

Introduction. J.M. Burgers Center Graduate Course CFD I January Least-Squares Spectral Element Methods

arxiv: v1 [math.na] 5 Jun 2018

Iterative methods for positive definite linear systems with a complex shift

Analysis of a high order trace finite element method for PDEs on level set surfaces

1. Introduction. We consider the model problem that seeks an unknown function u = u(x) satisfying

Optimal Left and Right Additive Schwarz Preconditioning for Minimal Residual Methods with Euclidean and Energy Norms

An Introduction to Algebraic Multigrid (AMG) Algorithms Derrick Cerwinsky and Craig C. Douglas 1/84

Adaptive Coarse Space Selection in BDDC and FETI-DP Iterative Substructuring Methods: Towards Fast and Robust Solvers

arxiv: v3 [math.na] 8 Sep 2015

Solutions of Selected Problems

A Balancing Algorithm for Mortar Methods

arxiv: v2 [math.na] 16 Nov 2016

BETI for acoustic and electromagnetic scattering

Local Mesh Refinement with the PCD Method

NONSTANDARD NONCONFORMING APPROXIMATION OF THE STOKES PROBLEM, I: PERIODIC BOUNDARY CONDITIONS

Transcription:

Universität Stuttgart Second Order Lagrange Multiplier Spaces for Mortar Finite Elements in 3D Bishnu P. Lamichhane, Barbara I. Wohlmuth Berichte aus dem Institut für Angewandte Analysis und Numerische Simulation Preprint 2003/007

Universität Stuttgart Second Order Lagrange Multiplier Spaces for Mortar Finite Elements in 3D Bishnu P. Lamichhane, Barbara I. Wohlmuth Berichte aus dem Institut für Angewandte Analysis und Numerische Simulation Preprint 2003/007

Institut für Angewandte Analysis und Numerische Simulation (IANS) Fakultät Mathematik und Physik Fachbereich Mathematik Pfaffenwaldring 57 D-70 569 Stuttgart E-Mail: WWW: ians-preprints@mathematik.uni-stuttgart.de http://preprints.ians.uni-stuttgart.de ISSN 6-476 c Alle Rechte vorbehalten. Nachdruck nur mit Genehmigung des Autors. IANS-Logo: Andreas Klimke. L A T E X-Style: Winfried Geis, Thomas Merkle.

Second Order Lagrange Multiplier Spaces for Mortar Finite Elements in 3D Bishnu P. Lamichhane and Barbara I. Wohlmuth Abstract Domain decomposition techniques provide a flexible tool for the numerical approximation of partial differential equations. Here, we consider mortar techniques for quadratic finite elements in 3D with different Lagrange multiplier spaces. In particular, we focus on Lagrange multiplier spaces which yield optimal discretization schemes and a locally supported basis for the associated constrained mortar spaces. As a result, standard efficient iterative solvers as multigrid methods can be easily adapted to the nonconforming situation. We present the discretization errors in different norms for linear and quadratic mortar finite elements with different Lagrange multiplier spaces. Numerical results illustrate the performance of our approach. Key words. mortar finite elements, Lagrange multiplier, dual space, domain decomposition, nonmatching triangulation. AMS subject classification. 65N30, 65N55. Introduction The coupling of different discretization schemes or of nonmatching triangulations can be analyzed within the framework of mortar methods. These nonconforming domain decomposition techniques provide a more flexible approach than standard conforming approaches. The nonconforming approach is of particular interest in many situations, for example, in problems with discontinuous diffusion coefficients and local anisotropies, when different parameters dominate different parts of the simulation domain or different discretization schemes are used in different subdomains. A complex global domain can be decomposed into several small subdomains of simple structure, and these subdomains can be meshed independently. To obtain a stable and optimal discretization scheme for the global problem, the information transfer among the subdomains has to be analyzed. Mortar methods were originally introduced to couple spectral and finite element approximations, see [BDM90, BMP93]. An optimal a priori estimate in the H -norm for the mortar finite element method has been established in [BMP93, BMP94, Ben99, BM97]. The analysis of three-dimensional mortar finite elements is given in [BM97, KLPV0, BD98], and a hp version is studied in [SS00]. In the linear 3D case, the stability of the mortar projection for the standard Lagrange multiplier space is established in [BD98], where also a multigrid This work was supported in part by the Deutsche Forschungsgemeinschaft, SFB 404, C2. 5

method for the saddle point problem is discussed. The main idea of the mortar technique is to replace the strong continuity condition of the solution across the interface by a weak one. Here, we consider mortar methods for second order finite elements in 3D. We focus on Lagrange multiplier spaces which yield locally supported basis functions of the constrained mortar space. The paper is organized as follows: In the rest of this section, we present our model problem and briefly review the mortar method. In Section 2, we give some sufficient conditions on Lagrange multiplier spaces for quadratic finite elements and show the optimality of the approach. In Section 3, we present some examples of Lagrange multiplier spaces for quadratic finite elements for hexahedral triangulations. In contrast to earlier approaches, we use Lagrange multiplier spaces yielding a sparse inverse of the mass matrix. We also use the so-called dual Lagrange multiplier spaces which are biorthogonal to the trace of the finite element space at the interface. Unfortunately, a locally defined dual Lagrange multiplier space containing the bilinear hat functions does not exist for the serendipity elements. In that case, we augment the space by face bubble functions. Finally in Section 4, we present some numerical results in 3D for different Lagrange multiplier spaces illustrating the flexibility and performance of our approach. In particular, we consider the discretization errors in the L 2 -norm, the energy norm and in a weighted L 2 -norm for the Lagrange multiplier. We consider the following elliptic second order boundary value problem div(a u) + cu = f, in Ω, u = 0, on Ω, () where 0 < a 0 a L (Ω), f L 2 (Ω), 0 c L (Ω), and Ω 3 is a bounded polyhedral domain. The domain Ω is decomposed into K non-overlapping polyhedral subdomains Ω k, k =,, K, such that Ω = K Ω k with Ω i Ω j = for i j. k= Here, we consider only geometrically conforming situations where the intersection between the boundaries of any two different subdomains Ω l Ω k, k l, is either empty, a common edge or a face. We define Γ kl := Ω k Ω l, k, l K, the intersection of the boundaries of two subdomains and select only disjoint and non-empty interfaces γ k, k N. Moreover, each γ k can be associated with a couple k < k 2 K such that γ k = Ω k Ω k2. On each subdomain, we define H (Ω k ) := {v H (Ω k ), v Ω Ωk = 0}, k =,, K and consider the unconstrained product space X := K H (Ω k). k= To impose a weak matching condition on the interface, a Lagrange multiplier space M := N k= H /2 (γ k ) is introduced on the skeleton Γ = N γ k. k= 6

Then, the weak matching condition on the skeleton Γ is realized in terms of the L 2 -orthogonality of the jump [u] of the solution across the interface and the Lagrange multiplier space M [u] µ dσ = 0, µ M. Γ Each subdomain Ω k is associated with a shape regular family of hexahedral triangulations T k;hk, the meshsize of which is bounded by h k. We denote the discrete space of conforming piecewise triquadratic finite elements or of serendipity elements on Ω k associated with T k;hk by X hk H (Ω k ). Now, the unconstrained finite element space X h can be written as X h := K X hk. k= Each interface γ k inherits a two-dimensional triangulation S k;hk either from T k;h k or T k2;h k2. The subdomain from which the interface inherits its triangulation is called slave or nonmortar side, the opposite one master or mortar side. In the following, we denote the index of the slave side by s(k) and the one of the master side by m(k). Hence, the elements of S k;hk are boundary faces of T s(k);hs(k) with a meshsize bounded by h s(k). Furthermore, we assume that the mesh on Γ is globally quasi-uniform, and each element in S k;hk, k =,, N, can be affinely mapped to the reference element ˆT := (0, ) (0, ). The discrete Lagrange multiplier space M h on Γ is defined as M h := N k= M h(γ k ). Then, the discrete weak matching condition for v h X h can be written as [v h ] µ i dσ = 0, i n k, k N, (2) γ k where n k := dim M h (γ k ) and {µ i } i nk forms a basis of M h (γ k ). Here, [v h ] is the jump of the function v h on γ k from the master side to the slave side. As usual, s,ωk and (, ) s,ωk denote the norm and the corresponding inner product on H s (Ω k ), respectively, and s,ωk stands for the seminorm. The norm on H /2 00 (γ k) and its dual space H /2 (γ k ) will be denoted by /2 H 00 (γ k) and /2,γ k, respectively. We define the broken norm s on X and the broken dual norm M on M by K u 2 s := N u 2 s,ω k, and µ 2 M := µ 2 /2,γ k, k= respectively. There are two main approaches to obtain the mortar solution u h X h of a discrete variational problem. The first one is based on the positive definite variational problem on the constrained finite element space which is given by means of the global Lagrange multiplier space M h V h := {v h X h b(v h, µ h ) = 0, µ h M h }, where b(v h, µ h ) := N k= γ k [v h ] µ h dσ. We remark that the elements of the space V h satisfy a weak continuity condition on the skeleton Γ in terms of discrete Lagrange multiplier space M h. However, V h is, in general, not a subspace of H0 (Ω). Then, the variational formulation of the mortar method can be given in terms of the constrained space V h : find u h V h such that a(u h, v h ) = (f, v h ) 0, v h V h. (3) k= 7

Here, the bilinear form a(, ) is defined as K a(v, w) := k= Ω k a v w + cv w dx. The second approach is based on enforcing the weak continuity condition on the skeleton Γ as an additional variational equation which leads to a saddle point problem on the unconstrained product space X h, see [Ben99]: find (u h, λ h ) X h M h such that a(u h, v h )+ b(v h, λ h ) = (f, v h ) 0, v h X h, b(u h, µ h ) = 0, µ h M h. (4) It is clear that the choice of the discrete Lagrange multiplier space M h plays an essential role for the stability of the saddle point problem and the optimality of the discretization scheme. The Lagrange multiplier space has to be large enough to obtain an optimal consistency error, and it has to be small enough to get an optimal best approximation error and a suitable discrete inf-sup condition. In the next section, we state sufficient conditions on the Lagrange multiplier space for quadratic finite elements to get optimal a priori estimates. Here, the nodal Lagrange multiplier basis functions are defined locally and are associated with the interior nodes of the mesh on γ k, k =,, N. Now, we group the degrees of freedom of X h associated with the skeleton Γ into two groups u h Γ := (u m, u s ), where u m contains all nodal values of u h on the master sides and all nodal values at the nodes on the boundary of the interface γ k on the slave sides, and u s consists of all nodal values of u h at the interior nodes of γ k on the slave sides, k N, see Figure. The associated sets of nodes are called N m and N s, respectively. Furthermore, we denote by N h the set of all nodes in X h and we set N i := N h \(N m N s ). The corresponding nodal values of u h in N i will be denoted by a block vector u i. Then (2) can be written in its algebraic form as M s u s + M m u m = 0. (5) The entries of the mass matrices are given by m ij := γ k [φ j ] µ i dσ, where φ j are the finite u m u s PSfrag replacements master side slave side Figure : Decomposition into u m and u s for the serendipity elements element basis functions corresponding to the different groups of nodes, and µ i denote the basis functions of M h. Since the basis functions have a local support, the mass matrices are sparse. Formally, we can obtain the values on the slave side as u s = Ms M m u m. Although M s is a sparse matrix, the inversion of M s is, in general, expensive, and Ms is 8

dense. This observation motivates our interest in Lagrange multiplier spaces which yield a sparse inverse of the mass matrix M s. A natural choice is a dual Lagrange multiplier space, see, e.g., [Woh0], having a diagonal mass matrix M s. Then, the basis functions {µ i } i nk of M h (γ k ) and {ϕ i } i nk of the trace space W 0,h (γ k ) having the zero boundary condition on γ k satisfy the biorthogonality relation µ i ϕ j dσ = δ ij ϕ j dσ, i, j n k. (6) γ k γ k If M s is a sparse triangular matrix, the inverse of M s is also a sparse triangular matrix. Hence, to get a sparse inverse of mass matrix M s, it is enough to work with a Lagrange multiplier space which yields a sparse triangular mass matrix M s. We define the product space W 0,h as N W 0,h := W 0,h (γ k ), k= and the broken H /2 00 -norm on W 0,h is denoted by v 2 W := N k= v 2 H /2 00 (γ k). 2 A priori estimates In this section, we give some assumptions on quadratic Lagrange multiplier spaces which guarantee optimal a priori estimates. Following a similar approach as in [KLPV0], we impose the following assumptions on the discrete Lagrange multiplier spaces for quadratic finite elements [P0 ] dim W 0,h (γ k ) = dim M h (γ k ), k N. [P ] There is a constant C independent of the triangulation such that inf v µ 0,γ k Ch 2 s(k) v 2,γ k, v H 2 (γ k ), k N. µ M h (γ k ) [P2 ] There is a constant C independent of the triangulation such that (θ, µ) 0,γk θ 0,γk C sup, θ W 0,h (γ k ), k N. µ M h (γ k )\{0} µ 0,γk It follows from assumption [P] that P (γ k ) M h (γ k ) for all k =,, N, where P (γ k ) is the space of linear functions on γ k. For each γ k, the mortar projection Π k : L 2 (γ k ) W 0,h (γ k ) is defined as Π k v µ dσ = v µ dσ, µ M h (γ k ). (7) γ k γ k The stability of the mortar projection is essential for the optimality of the best approximation error. 9

Lemma. Under the assumptions [P0] and [P2], the mortar projection (7) is stable in the L 2 -norm. Furthermore, if w H 0 (γ k ) Π k w,γk C w,γk. Proof: By assumption [P2], we find that if v W 0,h (γ k ) satisfies (v, µ) 0,γk = 0 for all µ M h (γ k ), then v = 0. Hence, the mortar projection is well-defined by the assumptions [P0] and [P2]. The L 2 -stability of Π k follows from (Π k w, µ) 0,γk (w, µ) 0,γk Π k w 0,γk C sup = C sup C w 0,γk. µ M h (γ k )\{0} µ 0,γk µ M h (γ k )\{0} µ 0,γk Now, using the L 2 -stability and an inverse estimate, we find for w H0 (γ k) ( ) Π k w,γk Π k w P w,γk + P w,γk C Π k (w P w) 0,γk + w,γk h s(k) ( ) C w P w 0,γk + w,γk C w,γk, h s(k) where P denotes the L 2 -projection onto W 0,h (γ k ). Using Lemma and an interpolation argument, we obtain for w H /2 00 (γ k), Π k w H /2 00 (γ k) C w H /2 00 (γ k). In a next step, we provide the best approximation property of the space V h. We use the ideas and techniques introduced in [BM97, BMP93]. Lemma 2. Assume that the assumptions [P0] [P2] hold. If u H0 (Ω) and u Ω k H 3 (Ω k ) for all k =,, K, then there exists a constant C depending only on the ratio of the meshsizes of the master and slave sides such that inf u u h 2 C u h V h K h 4 k u 2 3,Ω k. k= Proof: Since W 0,h (γ k ) H /2 00 (γ k), each v W 0,h (γ k ) can trivially be extended to a function ṽ H /2 ( Ω s(k) ). Let H h ṽ H (Ω s(k) ) be the discrete harmonic extension of ṽ on Ω s(k). Then, H h ṽ,ωs(k) C ṽ H /2 ( Ω s(k) ) C v /2 H. By means of this discrete 00 (γ k) harmonic extension, we define a discrete extension operator E k : W 0,h (γ k ) X h for each γ k as E k v := H h ṽ on Ω s(k), and E k v := 0 elsewhere. Then E k v C v H /2 00 (γ k), v W 0,h(γ k ). (8) Let l h u X h be the Lagrange interpolant of u in X h. It is easy to see that v := l h u + N k= E kπ k [l h u] is an element of V h. Then, we find N u v u l h u + E k Π k [l h u]. 0 k=

By using (8) and a coloring argument, we have N K E k Π k [l h u] 2 = N N E k Π k [l h u] 2,Ω l C Π k [l h u] 2. H /2 00 (γ k) k= l= k= We note that the constant C does not depend on the number of subdomains. Applying the L 2 (γ k )-stability of Π k and an inverse estimate, we get Π k [l h u] 2 H /2 00 (γ k) C h s(k) [l h u] 2 0,γ k k= h mr k= k= C ( ) (u l h u) Ωm(k) 2 0,γ h k + (u l h u) Ωs(k) 2 0,γ k s(k) C ( ) h 5 m(k) h u 2 3,Ω m(k) + h 5 s(k) u 2 3,Ω s(k). s(k) Summing over all k =,, N, we obtain ( ) N N N E k Π k [l h u] 2 C h 4 m(k) u 2 3,Ω m(k) + h 4 s(k) u 2 3,Ω s(k), where { hm(k) h mr := max, h s(k) k= } k N. Finally, the lemma follows by using the interpolation property of l h u. We remark that in contrast to a convergence theory of mortar finite elements in 2D, the constant in the right hand side depends on the ratio of the meshsizes of master and slave sides. However, if the mesh on the wirebasket is conforming and u is continuous, we find [l h u] H /2 00 (γ k) and thus the H /2 00 -stability of Π k can be directly applied. In that case, the ratio does not enter in the upper bound, see [KLPV0]. Working with mesh dependent norms and a trivial extension shows that the global ratio h mr can be replaced by a local one. Theorem 3. Let u and u h be the solutions of the problems () and (3), respectively. Assume that u H0 (Ω), u Ωk H 3 (Ω k ) for k =,, K, and [a u n ] = 0 on Γ. Under the assumptions [P0] [P2], there exists a constant C depending only on the ratio of the meshsizes of the master and the slave sides such that Proof: K u u h 2 C h 4 k u 2 3,Ω k. The bilinear form a(, ) is continuous on X, and it is coercive on { } B := v v H (Ω k ), k K, and [v] dσ = 0, k N, γ k k= see [BMP93]. Hence, assumption [P] assures that V h B. Thus, Strang s Lemma [BS94] can be applied, and we get ( ) a(u u h, v h ) u u h C inf u v h + sup. (9) v h V h v h v h V h \{0}

The first term in the right side of (9) denotes the best approximation error and the second one stands for the consistency error. Lemma 2 guarantees the required order for the best approximation error. Thus it is sufficient to consider the consistency error in more detail. Now, a(u u h, v h ) can be written as a(u u h, v h ) = Γ a u n [v h] dσ = N k= ( a u ), [v h ], v h V h. n k 0,γ k Here, u u n is the outward normal derivative of u on Γ from the master side, and n = u n k on γ k. We take µ M h (γ k ), then the definition of V h yields ( a u ) (, [v h ] = a u ) µ, [v h ]. n k 0,γ k n k 0,γ k Since µ M h (γ k ) is arbitrary, we can bound the consistency error by ( a u ) u, [v h ] inf a µ n k 0,γ µ M h (γ k ) k n (H /2 (γ k k )) [v h] /2,γk. Let Q : L 2 (γ k ) M h (γ k ) be the L 2 -projection onto M h (γ k ). Using assumption [P2] and an interpolation between L 2 (γ k ) and H 2 (γ k ), we obtain w Qw 0,γk Ch 3/2 s(k) w 3/2,γ k, w H 3/2 (γ k ). In terms of a standard Aubin Nitsche trick and the previous estimate, we get w Qw (H /2 (γ k )) Ch/2 s(k) w Qw 0,γ k Ch 2 s(k) w 3/2,γ k. Finally, the trace theorem yields ( a u ), [v h ] Ch 2 u s(k) n k 0,γ k n k [v h ] /2,γk 3/2,γk Ch 2 s(k) u 3,Ω s(k) ( vh,ωm(k) + v h,ωs(k) ). Now, using the Cauchy Schwarz inequality and summing over all k =,, N, we obtain ( K ) /2 a(u u h, v h ) C v h h 4 k u 2 3,Ω k. k= To obtain an a priori estimate for the Lagrange multipliers, we follow exactly the same procedures as in [Ben99]. Lemma 4. Assume that the Lagrange multiplier space M h satisfies assumptions [P0] [P2]. Then, for every µ M h, there exists a v µ X h such that v µ C µ M, µ 2 M Cb(v µ, µ) and [v µ ] W C µ M. 2

Proof: By means of the stability of the mortar projection, we get for µ M h (γ k ) µ /2,γk = sup ϕ H /2 00 (γ k)\{0} = C sup ϕ W 0,h (γ k )\{0} (µ, ϕ) 0,γk ϕ /2 H 00 (γ k) (µ, ϕ) 0,γk ϕ H /2 00 (γ k) C sup ϕ H /2 00 (γ k)\{0} (µ, Π k ϕ) 0,γk Π k ϕ H /2 00 (γ k) C(µ, ϕ k ) 0,γk (0) for some ϕ k W 0,h (γ k ) with ϕ k H /2 00 (γ k) =. Now, we extend ϕ k W 0,h (γ k ) to X h by using the extension operator E k as defined in Lemma 2 to get E k ϕ k =: v k X h. Then, we have v k C ϕ k H /2 00 (γ k) and 0 (µ, ϕ k ) 0,γk = b(v k, µ). Setting v µ := N k= b(v k, µ)v k and using the fact that v k C, we get N K N v µ 2 = b(v k, µ)v k 2 = b(v k, µ)v k 2,Ω l k= k= l= k= N N C b(v k, µ) 2 C µ 2 /2,γ k = C µ 2 M. This proves the first assertion. Using a coloring argument, the constant C does not depend on the number of subdomains. Furthermore, summing the equation (0) over all interfaces γ k, k =,, N, we obtain the second assertion µ 2 M C k= N b(v k, µ) 2 = Cb(v µ, µ). k= Finally, the third assertion follows from N [v µ ] 2 W = b(v k, µ) 2 [v k ] 2 = N N b(v H /2 00 (γ k, µ) 2 µ 2 k) /2,γ k [v k ] 2 H /2 00 (γ k) = µ 2 M. k= k= k= The best approximation property, Lemma 4, assumption [P2] and the first equation of the saddle point problem give an a priori estimate for the Lagrange multipliers. Lemma 5. Under the assumptions of Theorem 3, we have λ λ h 2 M C K h 4 k u 2 3,Ω k. k= 3 Quadratic Lagrange multiplier spaces in 3D In this section, we consider different possibilities for Lagrange multiplier spaces in 3D for quadratic finite elements with suppϕ i = suppµ i. In particular, we focus on the standard finite elements and the serendipity elements and restrict ourselves to hexahedral triangulations. These two finite elements have different degrees of freedom on the interface and therefore, the Lagrange multiplier spaces should be considered separately. 3

3. A dual Lagrange multiplier space for the standard triquadratic finite elements In the case of a hexahedral triangulation, a dual Lagrange multiplier space in 3D for trilinear and triquadratic finite elements can be formed by taking the tensor product of the dual Lagrange multiplier space in 2D. Let ˆϕ 0, ˆϕ and ˆϕ 2 be the nodal quadratic finite element basis functions on the reference element (0, ) in one dimension, where ˆϕ 0 and ˆϕ are the basis functions corresponding to the left and the right vertices of the reference element, and ˆϕ 2 is the basis function corresponding to the midpoint of the reference element. Then, the quadratic dual Lagrange multiplier basis functions on the reference element are defined by ˆλ 0 (t) := ˆϕ 0 (t) 3 4 ˆϕ 2(t) + 2, ˆλ (t) := ˆϕ (t) 3 4 ˆϕ 2(t) + 2 and ˆλ 2 (t) := 5 2 ˆϕ 2(t). The Lagrange multiplier basis functions for the element touching a crosspoint should be modified. In particular, if t = 0 is a crosspoint, we have and if t = is a crosspoint, we set ˆλ 2 (t) := 2t + 2, ˆλ (t) := 2t, ˆλ 2 (t) := 2t, ˆλ0 (t) := 2t. Furthermore, for a hat function φ l p at an interior vertex p, we find φ l p(t) = µ p (t) + 2 (µ e(t) + µ e2 (t)), () where µ p is a Lagrange multiplier corresponding to the vertex p and µ e and µ e2 are the basis functions associated with the midpoints of the two adjacent edges. If p is a crosspoint, we have φ l p(t) = 2 µ e(t), where µ e is the Lagrange multiplier basis function corresponding to the midpoint of the edge containing the crosspoint. Then, the Lagrange multiplier basis functions on the reference element ˆT = (0, ) (0, ) having a tensor product structure are defined as ˆλ ij (x, y) := ˆλ i (x)ˆλ j (y). Here, ˆλ 00 (x, y), ˆλ 0 (x, y), ˆλ (x, y) and ˆλ 0 (x, y) are the Lagrange multipliers corresponding to the four vertices (0, 0), (, 0), (, ) and (0, ), and ˆλ 20 (x, y), ˆλ2 (x, y), ˆλ2 (x, y) and ˆλ 02 (x, y) are the ones corresponding to the midpoints (0.5, 0), (, 0.5), (0.5, ) and (0, 0.5) of the four edges, respectively, and finally ˆλ 22 (x, y) is the one corresponding to the center of gravity (0.5, 0.5) of the reference element. The Lagrange multiplier basis functions are associated with the vertices, midpoints of the edges and the center of gravity of elements in S k;hk, k N. The global basis functions µ i are obtained by using an affine mapping and gluing the local ones together. All nodes on the boundary γ k of γ k are crosspoints and do not carry a degree of freedom for the Lagrange multiplier space. We note that we have to use the modification at the crosspoints to compute the tensor product for the Lagrange multipliers corresponding to the elements touching γ k. Observing (), we find that the bilinear hat function at each vertex is contained in the Lagrange multiplier space M h (γ k ). We point out that this is also valid on γ k, although there are no degrees of freedom. Hence, assumption [P] is satisfied. Assumption [P0] is trivially satisfied by construction. Now, 4

we verify assumption [P2]. Let ϕ := n k k= a kϕ k be in W 0,hk (γ k ) and set µ := n k k= a kµ k. In the following, we assume that ˆϕ i and ˆµ i are obtained from ϕ i and µ i by an affine mapping from the element T to the reference element ˆT. Now, by using the biorthogonality relation (6) and the quasi-uniformity assumption, we get (ϕ, µ) 0,γk = n k i,j= n k a i a j (ϕ i, µ j ) 0,γk = i= a 2 i n k ϕ i dσ C a 2 i h2 s(k) C ϕ 2 0,γ k. γ k Taking into account the fact that ϕ 2 0,γ k µ 2 0,γ k n k i= a2 i h2 s(k), we find that assumption [P2] is satisfied. Figure 2 shows the three different types of Lagrange multipliers on the reference element. i= 2.5 2.5 2.5 2 2 2.5.5.5 0.5 0.5 0.5 0 0 0.5 0.5 0 0.5.5.5 0.8 0.6 0.4 0.2 0 0 0.2 0.4 0.6 0.8 0.8 0.6 0.4 0.2 0 0 0.2 0.4 0.6 0.8 0.8 0.6 0.4 0.2 0 0 0.2 0.4 0.6 0.8 Figure 2: The Lagrange multipliers corresponding to a vertex (left), to an edge (middle) and to the center of gravity (left) 3.2 Serendipity elements and a dual Lagrange multiplier space Here, we give a non-existence result for a dual Lagrange multiplier space for the serendipity elements. A similar result for simplicial triangulations and quadratic finite elements is given in [LW02]. We denote by W h (γ k) the finite element space of piecewise bilinear hat functions on γ k. In case of standard triquadratic finite elements, the dual Lagrange multiplier space with tensor product structure contains W h (γ k). Unfortunately, there exists no dual Lagrange multiplier space for the serendipity elements yielding optimal a priori estimates with suppϕ i = suppµ i, where ϕ i are the serendipity nodal finite element basis functions on the interface γ k. Lemma 6. Under the assumption that suppϕ i = suppµ i, there exists no dual Lagrange multiplier space M h (γ k ) such that W h (γ k) M h (γ k ). Proof: We prove this by contradiction. Assume that α i µ i = φ l p, (2) i 5

where φ l p is the bilinear hat function associated with the interior vertex p having the coordinates (0, 0), see Figure 3. Suppose the coordinates of the four corners of the element T be (, 0), (0, 0), (0, ) and (, ), and of the element T 2 be (0, 0), (, 0), (, ) and (0, ). PSfrag replacements ϕ j0 at (0, ) T T 2 φ l p at (0, 0) Figure 3: 2D interface of 3D hexahedral triangulation Because of the duality, the functions µ i are biorthogonal to the finite element basis functions ϕ i on the interface. Hence, after multiplying (2) by some finite element basis function ϕ j and integrating over the interface γ k, we get α j = γ k ϕ j φ l p dσ γ k ϕ j dσ. Let j 0 be the interior vertex with coordinates (0, ) such that j 0 and p share one edge, see Figure 3. Then, we can write ϕ j0 φ l p dσ = ϕ j0 φ l p dσ + ϕ j0 φ l p dσ = γ k T T 2 8 and thus α j0 0. Since the basis functions µ i are locally linearly independent, we obtain suppµ j0 supp i α iµ i. By construction, we find suppµ j0 suppφ l p, which contradicts (2). 3.3 Lagrange multiplier spaces for the serendipity elements The previous subsection shows that there does not exist a dual Lagrange multiplier space for the serendipity elements containing the bilinear hat function at each vertex and satisfying suppϕ i = suppµ i. Here, we consider two different Lagrange multiplier spaces for the serendipity elements. The essential point is that the Lagrange multiplier space should lead to an optimal and stable discretization scheme. For this purpose, the assumptions [P0] [P2] are crucial. The first idea is to choose a standard Lagrange multiplier space, see [BMP93, BMP94]. In this case, the basis functions for each interior element T S k;hk of the interface γ k (i.e., T S k;hk with T γ k = ) are the serendipity elements in 2D. All nodes on γ k do not carry a degree of freedom for the Lagrange multipliers. Therefore, in order to satisfy assumption [P], it is necessary to modify the definition of the basis functions for the elements touching the boundary γ k of the interface γ k. Suppose an element T S k;hk with T γ k has n degrees of freedom for the Lagrange multipliers. Then, the local Lagrange multiplier basis function µ i at a node x i of T is chosen to be a polynomial 6

of minimal degree such that µ i (x j ) = δ ij for all x j, j =,, n. Here, δ ij is the Kronecker delta. These Lagrange multiplier basis functions are continuous. Since, in general, M h (γ k ) ( H /2 (γ k ) ), we can work with discontinuous Lagrange multipliers. Working with a continuous Lagrange multiplier space which locally contains the linear functions has the advantage that assumption [P] is satisfied. Here too, assumption [P0] is trivially satisfied by construction. To verify assumption [P2], we take ϕ := n k k= a kϕ k in W 0,h (γ k ) and define µ := n k k= a kµ k. Then (ϕ, µ) 0,γk = M ˆT = n k i,j= a i a j (ϕ i, µ j ) 0,γk = n k i,j= a i a j γ k ϕ i µ j dσ. The local mass matrix M ˆT for an element having all the vertices in γ k is given by 30 90 60 90 30 2 2 30 90 60 90 30 90 60 90 30 90 30 30 2 2 2 30 30 2 2 2 30 30 30 2 2 30 60 30 30 2 2 90 2 30 30 2 30 2 2 30 30 Similarly, computing the local mass matrices for the different boundary cases, we find that all the eigenvalues of the local mass matrices are greater than 6 00 and smaller than. Then (ϕ, µ) 0,γk, ϕ 2 0,γ k and µ 2 0,γ k are equivalent to n k i= h2 s(k) a2 i, which guarantees assumption [P2]. The coupling of the local mass matrices yield a global mass matrix which is sparse but has a band structure of band-width O(/h). Thus, the inverse of the global mass matrix M s on the slave side is dense. As a consequence, we obtain a stiffness matrix associated with the variational problem (3), which is not sparse. Then we cannot apply static condensation, and the multigrid method discussed in [WK0] cannot be used. To overcome this difficulty, we introduce a new approach for the serendipity elements. The idea is to use a Lagrange multiplier space which yields a sparse inverse of the global mass matrix M s on the slave side. In this case, we use the tensor product Lagrange multiplier space introduced in Subsection 3.. To satisfy the condition dim W 0,h (γ k ) = dim M h (γ k ), each non-empty face f T Γ of the element T of the slave side is enriched with a bubble function. The bubble function b H (T ) corresponding to the face f of T has the property that b T \f = 0 and f b dσ 0. We define K s := {T T s(k);hs(k), k N T Γ contains at least a face of T }, the set of elements on the slave side each of which has at least one face f on Γ. Now, the space of bubble functions B h is formed by N s bubbles, where N s is the number of elements in N k= S k;h k, and each of them is corresponding to a face f of an element T K s, where f T Γ. This leads to one additional degree of freedom for each non-empty face f T Γ of T K s. There are many possibilities to define such a bubble function. Here, the triquadratic nodal finite element function associated with the center of gravity of the face is used as a bubble function corresponding to this face. Although we need only the restriction of bubble functions to the associated face to satisfy assumption [P0], each bubble function is supported on the whole element. Now, the 8 9 4 9 9 8 9 4 4 9 8 9 9 4 9 8. 7

modified unconstrained product space Xh t can be written as X h t = Xs h B h, where Xh s is the unconstrained product space associated with the serendipity elements. In the sequel, the space Xh t will be called augmented serendipity space and the corresponding elements augmented serendipity elements. This leads to a mass matrix M s on the slave side having a special structure. Suppose ˆϕ i and ˆλ i, ( i 9) be the local basis functions of the standard triquadratic finite elements and their dual Lagrange multipliers, respectively. Here, the first four basis functions correspond to the vertices, the second four ones correspond to the midpoints of the edges, and the last one corresponds to the center of gravity of the reference element ˆT. Then, the local basis functions of the serendipity elements can be written as ˆϕ s i = ˆϕ i + α i ˆϕ 9, ( i 8), where α i = 4 for ( i 4) and α i = 2 for (5 i 8). Using the biorthogonality of ˆϕ i and ˆλ i, we have ˆϕ s ˆλ i j dσ = ( ˆϕ i + α i ˆϕ 9 ) ˆλ j dσ = δ ij ˆϕ i dσ + α i δ 9j ˆϕ 9 dσ. ˆT ˆT ˆT In fact, the mass matrix on the reference element ˆT is 36 0 0 0 0 0 0 0 0 0 36 0 0 0 0 0 0 0 0 0 36 0 0 0 0 0 0 0 0 0 36 0 0 0 0 0 M ˆT = 0 0 0 0 9 0 0 0 0. (3) 0 0 0 0 0 9 0 0 0 0 0 0 0 0 0 9 0 0 0 0 0 0 0 0 0 9 0 9 9 9 2 2 2 2 4 9 9 9 9 9 9 To show the consequence of our new Lagrange multiplier space, we consider the global mass matrix M s on the slave side in more detail. In the following, we use the same notation for the vector representation of the solution and the solution as an element in Xh t and M h. The matrix A is the stiffness matrix associated with the bilinear form a(, ) on Xh t Xt h, and the matrices B and B T are associated with the bilinear form b(, ) on Xh t M h. Then, the algebraic formulation of the saddle point problem (4) is given by [ ] [ ] [ ] A B T uh fh =. (4) B 0 λ h 0 We recall the grouping of the degrees of freedom of Xh t introduced in Section. After augmenting the serendipity space with the space of bubble functions B h, we further decompose the degrees of freedom associated with the interior nodes of γ k, k N, on the slave side into two groups (u s, u b ). Here, the block vector u s contains nodal values of u at the interior nodes of γ k, k N, corresponding to the vertices and edges on the slave side, and u b stands for all nodal values corresponding to the bubble functions on the slave side. With this decomposition, we can write u T h = (ut i, ut m, u T s, u T b ). The block vector λ h containing the nodal values of the Lagrange multiplier is similarly decomposed with λ T h = (λt s, λt b ). In terms of this decomposition, we can rewrite the algebraic form of the ˆT 8

saddle point problem (4) as A ii A im A is A ib 0 0 A mi A mm A ms A mb Mm T M T bm A si A sm A ss A sb D s Mbs T A bi A bm A bs A bb 0 D b 0 M m D s 0 0 0 0 M bm M bs D b 0 0 u i u m u s u b λ s λ b = f i f m f s f b 0 0. (5) Recalling the algebraic structure (5) of the bilinear form b(, ) restricted to Xh t M h, we have [ ] Mm D s 0 B =, M bm M bs D b where D b and D s are diagonal matrices, and M bs, M m and M bm are rectangular matrices. The matrix D b is diagonal due to the fact that the bubble functions are supported only in one element, and the diagonal form of D s follows from the structure of the local mass matrix, see (3). Hence, the global mass matrix M s on the slave side can be written as [ ] Ds 0 M s =. M bs The great benefit of this Lagrange multiplier space is that the inverse of the mass matrix M s can be computed very easily, and the inverse is sparse. In fact, the inverse of the mass matrix M s is [ ] M s = D s D b D b M bs Ds 0 D b Thus, the solution at the slave side depends locally on the solution at the master side. Here, we have to invert only two diagonal matrices and scale M bs to compute the inverse of the mass matrix M s. The stiffness matrix associated with the variational problem (3) is sparse, and efficient iterative solver like multigrid can easily be adapted to the nonconforming situation. Additionally, the condition number of the mass matrix is better compared to the standard Lagrange multiplier space. Furthermore, the degree of freedom corresponding to the bubble functions can locally be eliminated by static condensation. Since the matrix D b is diagonal, the sixth and the fourth line of the system (5) give u b = D b (M bm u m + M bs u s ), and λ b = D [ b fb A bi u i (A bm A bb D b M bm )u m (A bs A bb D ] b M bs )u s. Now, we eliminate u b and λ b from the system (5) and obtain a new system Âû h = ˆF h, where û T h = (ut i, ut m, ut s, λt s ). Defining M := D b M bm and M 2 := D b M bs we have A ii A im A ib M A is A ib M 2 0 Â = A mi M T A bi A mm A mb M M T (A bm A bb M ) A ms A mb M 2 M T (A bs A bb M 2) Mm T A si M2 T A bi A sm A sb M M2 T (A bm A bb M ) A ss A sb M 2 M2 T (A bs A bb M 2) D s, 0 M m D s 0. 9

and the right hand side can be written as ˆF h = f i f m M T f b f s M T 2 f b 0 We observe that the matrix  is symmetric, if A is symmetric and it has exactly the same structure as the saddle point matrix arising from mortar finite element method with a dual Lagrange multiplier space, see [Woh0]. Because of this structure of the algebraic system, we can apply the multigrid method proposed in [WK0]. Remark 7. There is also a possibility to use wavelets to get a mass matrix of special structure so that the inversion can be cheaper, and the inverse is sparse. In [Ste00], locally supported and piecewise polynomial wavelets are studied on non-uniform meshes which give a lower triangular mass matrix with higher order finite elements in triangular meshes.. 4 Numerical results Here, we present some numerical examples in 3D for linear and quadratic mortar finite elements. We consider three different cases for quadratic mortar finite elements. The first one is the standard triquadratic finite element space with the dual Lagrange multiplier space introduced in Subsection 3.. The second one is the serendipity space with a standard Lagrange multiplier space given in Subsection 3.2. Finally, the third one is the augmented serendipity space associated with the tensor product Lagrange multiplier space. Our numerical results show the same asymptotic behaviour as predicted by the theory. The implementation is based on the finite element toolbox ug, [BBJ + 97]. We do not discuss and analyze an iterative solver for the arising linear systems. Working with dual Lagrange multiplier spaces has the advantage that the flux can locally be eliminated, and static condensation yields a positive definite system on the unconstrained product space. In [WK0, Woh0], the modification of the system has been carried out and a local modification of the transfer operators of lower complexity has been proposed. The introduced multigrid has a level-independent convergence rate and is of optimal complexity. Unfortunately, in the case of a standard Lagrange multiplier space no local elimination of the flux can be carried out. Following the approach in [WK0], the sparsity of the modified system and the efficiency of the multigrid solver is lost. In that case, we apply a multigrid method for saddle point problems. This technique has been considered for mortar elements in [WW98] and further analyzed in [BDW99, WW02]. It turns out that we do not have to work in a positive definite subspace, and the smoother can be realized in an inner and outer iteration scheme. As in the other approach, level-independent multigrid convergence rates can be established. However, the numerical solution process is slower if we have to work with the saddle point approach. We point out that the more efficient multigrid method for the modified positive definite system can only be applied when the inverse of M s is sparse, whereas the saddle point multigrid method is more general. Although we do not have a dual Lagrange multiplier space for the serendipity elements, the tensor product Lagrange multiplier space in combination with the introduction of bubble functions on the slave side of the interface yield an optimal discretization scheme. We present some numerical results in 3D illustrating the performance of the different Lagrange multiplier spaces. In particular, 20

we compare the discretization errors in the L 2 - and H - norm for the solution for linear and quadratic mortar finite elements. The discretization errors in the flux across the interface are compared in a mesh-dependent Lagrange multiplier norm, which is defined by µ µ h 2 h := N m= T S m;hm h T µ µ h 2 0,T, where h T is the diameter of the element T. For all our examples, we have used uniform refinement. In each refinement step, one element is refined into eight subelements. We denote by Xh l and Xf h the unconstrained finite element spaces associated with the standard finite element spaces for the trilinear and the triquadratic case, respectively. Similarly, Xh s and Xh t are the unconstrained finite element spaces associated with the serendipity elements and the augmented serendipity elements as defined in the previous section, respectively. The corresponding solutions are denoted by u l h, uf h, us h and ut h, respectively. Remark 8. We note that the concept of dual Lagrange multiplier spaces can be generalized to distorted hexahedral meshes. In that case, the mapping between the actual element and the reference element has a non-constant Jacobian. As a consequence, we have to compute for each face on the interface a biorthogonal basis with respect to local nodal one. This can be easily done by solving a local mass matrix system. By construction, the sum of the local dual Lagrange multiplier basis functions is one. Defining the global Lagrange multiplier basis functions by gluing the local ones together, we find that the constants are included in the Lagrange multiplier space. As a consequence, it is easy to verify that the discretization error is of order h for the lowest order conforming finite elements. Our first example is given by u = f on Ω := (0, ) 2 (0, 2), where Ω is decomposed into two subdomains Ω := (0, ) 3 and Ω 2 := (0, ) 2 (, 2). The right hand side f and the boundary conditions are chosen such that the exact solution is given by u(x, y, z) = 2 exp( x 2 y 2 z 2 ) sin(2 yx) (x + y + z) (x y z). In Figure 4, the decomposition of the domain, the initial triangulation and the isolines of the solution at the interface z = are shown. We have used a nonconforming initial triangulation, where the Figure 4: Decomposition of the domain and initial triangulation (left), isolines of the solution at the interface (right), Example lower cube is refined once, and the upper cube is left as it is. The errors in the L 2 -, H - and the weighted Lagrange multiplier norms are given in Tables 3. The tables show that we get the predicted asymptotic convergence rates. For the quadratic case, the errors in the L 2-2

and H -norm are of order h 3 and h 2, whereas they are of order h 2 and h for the linear case. The errors in the weighted Lagrange multiplier norm for the quadratic and linear cases are of order h 5/2 and h 3/2, respectively. Theoretically, the errors in the weighted Lagrange multiplier norm for quadratic and linear case are expected to be of order h 2 and h, respectively. The better convergence rates are due to the fact that the error in the H -norm is equally distributed and the Lagrange multiplier space has an O(h 5/2 ) and O(h 3/2 ) approximation property in the considered norm. For the three different second order approaches, the quantitative results are asymptotically almost the same in the L 2 - and H -norm. Only in the weighted Lagrange multiplier norm, the full triquadratic yields better results. Table : Discretization errors in the L 2 -norm, (Example ) level # elem. u u l h 0 u u f h 0 u u s h 0 u u t h 0 0 9.004794e+00 8.867203e-0 9.2e-0 9.25653e-0 72 9.09292e-0 3.87528e-0 4.650386e-0 4.574443e-0 2 576 4.363422e-0 3.70530e-02 4.563530e-02 4.558460e-02 3 4608.033778e-0 7.0696e-03 7.649595e-03 7.650e-03 4 36864 2.640337e-02 9.606873e-04 9.65404e-04 9.63983e-04 Table 2: Discretization errors in the H -norm, (Example ) level # elem. u u l h u u f h u u s h u u t h 0 9.000886e+00 9.356647e-0 9.705044e-0 9.6780e-0 72 9.446222e-0 6.608442e-0 7.237820e-0 7.07652e-0 2 576 6.449633e-0.35306e-0.490407e-0.485072e-0 3 4608 3.005969e-0 4.988503e-02 5.40438e-02 5.08725e-02 4 36864.50883e-0.26993e-02.269328e-02.2676e-02 Table 3: Discretization errors in the weighted Lagrange multiplier norm, (Example ) level # elem. λ λ l h h λ λ f h h λ λ s h h λ λ t h h 0 9 4.685044e-0 5.949429e+00 6.980608e+00 8.650037e+00 72 3.072290e+00 4.254093e+00 4.30030e+00 5.528655e+00 2 576.8248e+00 4.052985e-0 6.575e-0 7.399e-0 3 4608 6.30025e-0.07732e-0.493473e-0.466488e-0 4 36864 2.246235e-0.669782e-02 2.04704e-02 2.40364e-02 In Example 2, we choose a L-shaped domain. The domain Ω := ((0, ) 2 (0, 2)) ([, 2) (0, ) 2 ) is decomposed into three cubes, Ω := (0, ) 3, Ω 2 := (0, ) 2 (, 2) and Ω 3 := (, 2) (0, ) 2. We have shown the decomposition of the domain and the initial triangulation in the left picture of Figure 5, and the isolines of the solution at the interface z = are shown in the right. As before, we solve a Poisson problem u = f 22

with the right hand side function f and the Dirichlet boundary conditions determined by ( the exact solution u(x, y, z) = (x ) 2 + (z ) 2) 5/6 ( cos 6 y 2 + x 2 + 6 ). Figure 5: Decomposition of the domain and initial triangulation (left), isolines of the solution at the interface z = (right), Example 2 We have tabulated the errors in different norms in Tables 4 6. Here, the solution is not H 3 -regular. Therefore, we cannot expect the same order of convergence as in the previous example. Since the solution u H 8/3 ɛ (Ω) for ɛ > 0, we expect the same order of convergence as in the previous example in the linear case. In the quadratic case, theoretically the orders of h 8/3 and of h 5/3 are expected in the L 2 - and H -norm, respectively. In all three cases of quadratic finite elements, we observe the asymptotic rates in the L 2 - and H -norm, which are better than predicted by the theory. The quantitative results are also almost the same in these norms. Here too, we observe the better convergence rates for the errors in the weighted Lagrange multiplier norm. Table 4: Discretization errors in the L 2 -norm, (Example 2) level # elem. u u l h 0 u u f h 0 u u s h 0 u u t h 0 0 0.327466e+00 7.3859e-0 8.066003e-0 7.95793e-0 80 8.047675e-0.748627e-0 2.039559e-0 2.008262e-0 2 640 2.057468e-0 4.86375e-02 4.936495e-02 4.90636e-02 3 520 6.7225e-02 6.422065e-03 6.969e-03 6.443622e-03 4 40960.76695e-02 8.056078e-04 8.0656e-04 8.066527e-04 In Examples and 2, there is not any significant difference in the accuracy between the different quadratic mortar solutions either in the L 2 -norm or in the H -norm. However, there is some quantitative difference in the errors in the weighted Lagrange multiplier norm between different quadratic mortar solutions. In this norm, the standard triquadratic finite elements with the tensor product Lagrange multiplier space gives the best result, whereas the difference between the augmented serendipity elements with the tensor product Lagrange multiplier space and the serendipity elements with the standard Lagrange multiplier space is quite negligible in Example 2. In Example, the serendipity elements with the standard Lagrange multiplier space yields better result than the augmented serendipity elements with the tensor product Lagrange multiplier space. However, the difference is not 23

Table 5: Discretization errors in the H -norm, (Example 2) level # elem. u u l h u u f h u u s h u u t h 0 0.02784e+00 8.0853e-0 8.9430e-0 8.049068e-0 80 8.094756e-0 3.93427e-0 4.259780e-0 4.59907e-0 2 640 4.22967e-0.803404e-0.83986e-0.84987e-0 3 520 2.479979e-0 4.695275e-02 4.732e-02 4.70404e-02 4 40960.27749e-0.73059e-02.73325e-02.73569e-02 Table 6: Discretization errors in the weighted Lagrange multiplier norm, (Example 2) level # elem. λ λ l h h λ λ f h h λ λ s h h λ λ t h h 0 0 9.99273e-0 5.98895e+00 4.678395e+00 7.046350e+00 80 2.467e+00.83235e+00 3.00680e+00 2.08278e+00 2 640 8.795363e-0 5.338894e-0 7.909840e-0 6.36054e-0 3 520 4.48377e-0 7.548844e-02 9.720802e-02.002547e-0 4 40960.720963e-0.5730e-02.577079e-02.684356e-02 essential, and the asymptotic convergence orders are the same. We remark that for both of these examples the asymptotic phase starts from the third step. This is due to the fact that the initial triangulation is very coarse. For the next two examples, we consider only the serendipity elements. In our third example, the domain Ω := (0, ) 2 (0, 2.5) is decomposed into three subdomains Ω := (0, ) 3, Ω 2 := (0, ) 2 (, 2), and Ω 3 := (0, ) 2 (2, 2.5). The right hand side f and the boundary conditions of u = f are chosen such that the exact solution is given by u(x, y, z) = 5(z.4)((x 0.5) 2 + 4(y 0.3) 3 ) + z(z ) sin(4πxy)(2(x y) 2 + (y + x ) 2 ). In Figure 6, we have shown the decomposition of the domain, the initial nonmatching triangulation and the isolines of the solution at the interface z = 2. Figure 6: Decomposition of the domain and initial triangulation (left), isolines of the solution at the interface z = 2 (right), Example 3 Here, we have three subdomains and two interfaces. The middle cube is taken as the slave 24

side. We start with a nonconforming coarse initial triangulation having 23 elements. The errors along with their order of convergence at every step of refinement in different norms are given in Tables 7 9. As before, we get the correct asymptotic rates for both cases of the serendipity elements, and the errors in the L 2 - and H -norm are almost the same for both approaches. In the weighted Lagrange multiplier norm the serendipity elements yield less errors than the augmented serendipity elements. However, the difference is quite negligible, and the asymptotic rate of convergence is optimal in both cases. Table 7: Discretization errors in the L 2 -norm, (Example 3) level # elem. u u l h 0 u u s h 0 u u t h 0 0 23 8.9337e-0.7670e-0.760480e-0 0 84 2.582954e-0.79 2.997997e-02 2.54 3.00899e-02 2.55 2 472 6.366337e-02 2.02 3.6695e-03 3.03 3.6773e-03 3.04 3 776.607229e-02.99 4.46663e-04 3.04 4.462098e-04 3.04 4 94208 4.03862e-03 2.00 5.393667e-05 3.05 5.39429e-05 3.05 Table 8: Discretization errors in the H -norm, (Example 3) level # elem. u u l h u u s h u u t h 0 23 8.70532e-0 5.57577e-0 5.290887e-0 0 84 5.643329e-0 0.53.47860e-0.90.482833e-0.84 2 472 2.626420e-0.0 3.95936e-02.92 3.920488e-02.92 3 776.293053e-0.02 9.352708e-03 2.07 9.332480e-03 2.07 4 94208 6.446694e-02.00 2.295583e-03 2.03 2.293897e-03 2.02 Table 9: Discretization errors in the weighted Lagrange multiplier norm, (Example 3) level # elem. λ λ l h h λ λ s h h λ λ t h h 0 23 7.43364e+00 2.76237e+0 2.758347e+0 0 84 5.657720e+00 0.39 2.006842e+00 3.78 3.320707e+00 3.05 2 472.855735e+00.6 7.048806e-0.5 8.042462e-0 2.05 3 776 4.868778e-0.93.00359e-0 2.82.599e-0 2.80 4 94208.832775e-0.4.56494e-02 2.68.879805e-02 2.62 In our last example, we consider a domain Ω := (0, 2) (0, ) (0, 2), which is decomposed into four subdomains Ω := (0, ) 3, Ω 2 := (0, ) 2 (, 2), Ω 3 := (, 2) (0, ) 2 and Ω 4 := (, 2) (0, ) (, 2). We have shown the decomposition of the domain and the initial triangulation in the left picture of Figure 7, the isolines of the solution on the plane y = 2 in the middle, and the flux of the exact solution at the interface x = is shown in the right one. Here, Ω 2 and Ω 3 are taken to be the slave sides and the rest are the master sides. The problem for this example is given by a reaction-diffusion equation div(a u) + u = f in Ω, 25