Publication Cover
Applicable Analysis
An International Journal
Volume 101, 2022 - Issue 6
944
Views
2
CrossRef citations to date
0
Altmetric
Articles

Adaptive BEM for elliptic PDE systems, part I: abstract framework, for weakly-singular integral equations

ORCID Icon & ORCID Icon
Pages 2085-2118 | Received 16 Apr 2020, Accepted 16 Jul 2020, Published online: 03 Aug 2020

Abstract

In the present work, we consider weakly-singular integral equations arising from linear second-order elliptic PDE systems with constant coefficients, including, e.g. linear elasticity. We introduce a general framework for optimal convergence of adaptive Galerkin BEM. We identify certain abstract conditions for the underlying meshes, the corresponding mesh-refinement strategy, and the ansatz spaces that guarantee that the weighted-residual error estimator is reliable and converges at optimal algebraic rate if used within an adaptive algorithm. These conditions are satisfied, e.g. for discontinuous piecewise polynomials on simplicial meshes as well as certain ansatz spaces used for isogeometric analysis. Technical contributions include the localization of (non-local) fractional Sobolev norms and local inverse estimates for the (non-local) boundary integral operators associated to the PDE system.

2010 Mathematics Subject Classifications:

1. Introduction

1.1. State of the art

For the Laplace model problem, adaptive boundary element methods (BEM) using (dis)continuous piecewise polynomials on triangulations have been intensively studied in the literature. In particular, optimal convergence of adaptive mesh-refining algorithms has been proved for polyhedral boundaries [Citation1–3] as well as smooth boundaries [Citation4]. The work [Citation5] allows to transfer these results to piecewise smooth boundaries; see also the discussion in the review article [Citation6]. In [Citation7], these results have been generalized to the Helmholtz problem. In recent years, we have also shown optimal convergence of adaptive isogeometric BEM (IGABEM) using one-dimensional splines for the 2D Laplace problem [Citation8,Citation9]. However, the important case of 3D IGABEM remained open. Moreover, a generalization to other PDE operators is highly non-trivial (see (Equation6) below), but especially linear elasticity is of great interest in the context of isogeometric analysis.

In [Citation10], we have considered isogeometric finite element methods (IGAFEM). We have derived an abstract framework which guarantees that, first, the classical residual FEM error estimator is reliable, and second, the related adaptive algorithm yields optimal convergence; see [Citation10, Section 2 and 4]. We then showed that, besides standard FEM with piecewise polynomials, this abstract framework covers IGAFEM with hierarchical splines (see [Citation10, Section 3 and 5]) as well as IGAFEM with analysis suitable T-splines (see the recent work [Citation11]).

The aim of the present work is to develop such an abstract framework also for BEM, which is mathematically much more demanding than FEM. In ongoing research [Citation12], we aim to show that this framework covers, besides standard discretizations with piecewise polynomials, also IGABEM with hierarchical splines resp. T-splines.

To this end, the present work focusses on weakly-singular integral equations. For a given Lipschitz domain ΩRd with compact boundary Γ:=Ω and right-hand side f:ΓC, we consider (1) (Vφ)(x):=ΓG(xy)φ(y)dy=f(x)for almost all xΓ.(1) Here, the fundamental solution G stems from an elliptic PDE operator (2) Pu:=i=1di=1di(Aiiiu)+i=1dbiiu+cu,(2) where the coefficients Aii=Aii¯,bi,cCD×D are constant for some fixed dimension D1.

1.2. Outline & contributions

In Section 2, we fix some general notation, recall Sobolev spaces on the boundary, and precisely state the considered problem. Section 3 can be paraphrased as follows. We formulate an adaptive algorithm (Algorithm 3.3) of the form (3) SOLVE ESTIMATE MARK REFINE (3) driven by some weighted-residual a posteriori error estimator (see (Equation4) below) in the frame of conforming Galerkin BEM. The algorithm particularly generates meshes T, BEM solutions Φ in associated nested ansatz spaces XX+1L2(Γ)DH1/2(Γ)D, and error estimators η for all N0. We formulate five Assumptions (M1)–(M5) on the underlying meshes (Section 3.1), five Assumptions (R1)–(R5) on the mesh refinement (Section 3.2), and six Assumptions (S1)–(S6) on the BEM spaces (Section 3.3). First, these assumptions are sufficient to guarantee that the a posteriori error estimator η associated with the BEM solution Φ is reliable, i.e. there exists a constant Crel>0 such that (4) Crel1φΦH1/2(Γ)η:=h1/2Γ(fVΦ)L2(Γ)for all N0,(4) where hL(Γ) denotes the local mesh-size function and Γ is the surface gradient. Second, Theorem 3.3 states that Algorithm 3.3 leads to linear convergence at optimal algebraic rate with respect to the number of mesh elements. In Theorem 3.7, we briefly note that the introduced conditions have already been implicitly proved for standard discretizations with piecewise polynomials on conforming triangulations. Moreover, we mention expected applications to adaptive IGABEM on quadrilateral meshes in Remark 3.8.

Section 4 is devoted to the proof of Theorem 3.3. To prove reliability (Equation4), we use a localization argument (Proposition 4.2), which generalizes earlier works [Citation13,Citation14] for standard discretizations. More precisely, we prove that (5) vH1/2(Γ)2CsplitTTTΠ(T)|v|H1/2(TT)2(5) for all vH1/2(Γ)D that are L2-orthogonal onto the ansatz space X corresponding to some mesh T, where Csplit>0 is independent of v and Π(T) denotes the patch of TT. Finally, the proof of reliability (Equation4) requires the derivation of a local Poincaré-type inequality (Proposition 4.9). In Remark 4.10, we note that one obtains at least plain convergence limφΦH1/2(Γ)=0 if Algorithm 3.3 is steered by the so-called Faermann estimator C~rel1φΦH1/2(Γ)ϝ:=TTTΠ(T)|fVΦ|H1/2(TT)21/2C~eff1φΦH1/2(Γ), which is reliable and efficient. To prove linear convergence at optimal rate for the weighted-residual estimator (Equation4), we show that the assumptions of Section 3 imply the axioms of adaptivity [Citation6]. The latter are properties for abstract mesh-refinements and abstract error estimators, which automatically yield the desired convergence result. In contrast to [Citation1,Citation3] which (implicitly) verify the axioms of adaptivity only for the Laplace problem, our analysis allows for general PDE operators (Equation2). The crucial step is the generalization (Proposition 4.13) of the non-trivial local inverse inequality for the non-local boundary integral operator V: With the help of a Caccioppoli-type inequality (Lemma 4.12), we prove that there exists a constant Cinv>0 such that (6) h1/2ΓVψL2(Γ)CinvψH1/2(Γ)+h1/2ψL2(Γ)for all ψL2(Γ)D;(6) see [Citation5] for standard BEM discretizations of the Laplacian. Similar estimates hold also for the other boundary integral operators related to (Equation2), namely the double-layer integral operator K, its adjoint K, and the hypersingular integral operator W. These are stated and proved in Appendix B; again we refer to [Citation5] for standard BEM discretizations of the Laplacian. In each case, ellipticity of the PDE operator is not required for the inverse inequalities.

While the present work focusses on the numerical analysis aspects only, we refer to the literature (see, e.g. [Citation7,Citation15–18]) for numerical experiments for the Laplace problem, the Helmholtz problem, and linear elasticity.

2. Preliminaries

In this section, we fix some general notation, recall Sobolev spaces on the boundary, and precisely state the considered problem. Throughout the work, let ΩRd for d2 be a bounded Lipschitz domain as in [Citation19, Definition 3.28] and Γ:=Ω its boundary.

2.1. General notation

Throughout and without any ambiguity, || denotes the absolute value of scalars, the Euclidean norm of vectors in Rn, as well as the Hausdorff measure of any d-dimensional set in Rn. Let Bε(x):={yRn:|xy|<ε} denote the open ball around x with radius ε>0. For ω1,ω2Rn, let Bε(ω1):=xω1Bε(x). Moreover, let diam(ω1):=sup{|xy|:x,yω1}, and dist(ω1,ω2):=inf{|xy|:xω1,yω2}. We write AB to abbreviate ACB with some generic constant C>0, which is clear from the context. Moreover, AB abbreviates ABA. Throughout, mesh-related quantities have the same index, e.g. X is the ansatz space corresponding to the mesh T. The analogous notation is used for meshes T, T, T, etc.

2.2. Sobolev spaces

For σ[0,1], we define the Hilbert spaces H±σ(Γ) as in [Citation19, page 99] by use of Bessel potentials on Rd1 and liftings via bi-Lipschitz mappingsFootnote1 that describe Γ. For σ=0, it holds that H0(Γ)=L2(Γ) with equivalent norms. We thus may define H0(Γ):=L2(Γ).

For σ(0,1], any measurable subset ωΓ, and all vHσ(Γ), we define the associated Sobolev–Slobodeckij norm (7) vHσ(ω)2:=vL2(ω)2+|v|Hσ(ω)2with|v|Hσ(ω)2:=ωω|v(x)v(y)|2|xy|d1+2σdxdyif σ(0,1),ΓvL2(ω)2if σ=1.(7) It is well known that Hσ(Γ) provides an equivalent norm on Hσ(Γ); see, e.g. [Citation20, Lemma 2.19] and [Citation19, Theorem 3.30 and page 99] for σ(0,1) and [Citation18, Theorem 2.28] for σ=1. Here, Γ() denotes the usual (weak) surface gradient which can be defined for almost all xΓ as follows. Since Γ is a Lipschitz boundary, there exist an open cover (Oj)j=1J in Rd of Γ such that each ωj:=OjΓ can be parametrized by a bi-Lipschitz mapping γωj:ωˆjωj, where ωˆjRd1 is an open set. By Rademacher's theorem, γωj is almost everywhere differentiable. The corresponding Gram determinant det(DγωjDγωj) is almost everywhere positive. Moreover, by definition of the space H1(Γ), vH1(Γ) implies that vγωjH1(ωˆj). With the weak derivative (vγωj)L2(ωˆj)d, we can hence define (8) (Γv)|ωj:=Dγωj(DγωjDγωj)1(vγωj)γωj1for all vH1(Γ).(8) This definition does not depend on the particular choice of the open sets (Oj)j=1J and the corresponding parametrizations (γωj)j=1J; see, e.g. [Citation18, Theorem 2.28]. With (Equation8), we immediately obtain the chain rule (9) (vγωj)=Dγωj((Γv)γωj())for all vH1(Γ).(9) For σ(0,1], Hσ(Γ) is a realization of the dual space of Hσ(Γ); see [Citation19, Theorem 3.30 and page 99]. With the duality bracket ;, we define an equivalent norm (10) ψHσ(Γ):=sup{v;ψ:vHσ(Γ)vHσ(Γ)=1}for all ψHσ(Γ).(10) Moreover, we abbreviate (11) (v;ψ):=v¯;ψfor all vHσ(Γ),ψHσ(Γ).(11) [Citation19, page 76] states that the inclusion Hσ1(Γ)Hσ2(Γ) for 1σ1σ21 is continuous and dense. In particular, Hσ(Γ)L2(Γ)Hσ(Γ) form a Gelfand triple in the sense of [Citation21, Section 2.1.2.4] for all σ(0,1], where ψL2(Γ) is interpreted as function in Hσ(Γ) via (12) v;ψ:=(v¯;ψ)L2(Γ)=Γvψdxfor all vHσ(Γ),ψL2(Γ).(12) Here, (;)L2(Γ) is the usual complex scalar product on L2(Γ).

So far, we have only dealt with scalar-valued functions. For D1, σ[0,1], v=(v1,,vD)Hσ(Γ)D, we define vH±σ(Γ)2:=j=1DvjH±σ(Γ)2. If σ>0, and ωΓ is an arbitrary measurable set, we define vHσ(ω) and |v|Hσ(ω) analogously. With the definition (13) Γv:=Γv1ΓvDL2(Γ)D2for all vH1(Γ)D,(13) it holds that |v|H1(ω)=ΓvL2(ω). Note that Hσ(Γ)D with σ(0,1] can be identified with the dual space of Hσ(Γ)D, where we set (14) v;ψ:=j=1Dvj;ψjfor all vHσ(Γ)D,ψHσ(Γ)D.(14) As before, we abbreviate (15) (v;ψ):=v¯;ψfor all  vHσ(Γ)D,ψHσ(Γ)D(15) and set (16) v;ψ:=(v¯;ψ)L2(Γ)=j=1DΓvjψjdxfor all vHσ(Γ)D,ψL2(Γ)D.(16) The spaces Hσ(Γ) can also be defined as trace spaces or via interpolation, where the resulting norms are always equivalent with constants depending only on the dimension d and the boundary Γ. More details and proofs are found, e.g. in the monographs [Citation19–21].

2.3. Continuous problem

We consider a general second-order linear system of PDEs (17) Pu:=i=1di=1di(Aiiiu)+i=1dbiiu+cu,(17) where the coefficients Aii,bi,cCD×D are constant for some fixed dimension D1. We suppose that Aii=Aii¯. Moreover, we assume that P is elliptic on H01(Ω)D in the sense of the Lax–Milgram lemma, i.e. the sesquilinear form (18) (u;v)P:=Ωi=1di=1d(Aiiiu)iv+i=1d(biiu)v+(cu)vdx(18) satisfies that (19) (u;u)PCelluH1(Ω)2for all uH01(Ω)D.(19) of the matrices Aii in the sense of [Citation19, page 119]. Here, the standard complex scalar product on CD is denoted by wz=j=1Dw¯jzj.

Let G:Rd{0}CD×D be a corresponding (matrix-valued) fundamental solution in the sense of [Citation19, page 198], i.e. a distributional solution of PG=δ, where δ denotes the Dirac delta distribution. For ψL(Γ)D, we define the single-layer operator as (20) (Vψ)(x):=ΓG(xy)ψ(y)dyfor all xΓ.(20) According to [Citation19, page 209 and 219–220] and [Citation22, Corollary 3.38], this operator can be extended for arbitrary σ(1/2,1/2] to a bounded linear operator (21) V:H1/2+σ(Γ)DH1/2+σ(Γ)D.(21) [Citation19, Theorem 7.6] states that V is always coercive, i.e. elliptic up to some compact perturbation. We assume that it is elliptic even without perturbation, i.e. (22) Re(Vψ;ψ)CellψH1/2(Γ)2for all ψH1/2(Γ)D.(22) This is particularly satisfied for the Laplace problem or for the Lamé problem, where the case d = 2 requires an additional scaling of the geometry Ω; see, e.g. [Citation20, Chapter 6]. Moreover, the sesquilinear form (V;) is continuous due to (Equation21), i.e. it holds with Ccont:=VH1/2(Γ)DH1/2(Γ)D that (23) |(Vψ;ξ)|CcontψH1/2(Γ)ξH1/2(Γ)for all ψ,ξH1/2(Γ)D.(23) Given a right-hand side fH1(Γ)D, we consider the boundary integral equation (24) Vφ=f.(24) Such equations arise from (and are even equivalent to) the solution of Dirichlet problems of the form Pu=0 in Ω with u = g on Γ for some gH1/2(Γ)D; see, e.g. [Citation19, page 226–229] for more details. The Lax–Milgram lemma provides existence and uniqueness of the solution φH1/2(Γ)D of the equivalent variational formulation of (Equation24) (25) (Vφ;ψ)=(f;ψ)for all ψH1/2(Γ)D.(25) In particular, we see that V:H1/2(Γ)DH1/2(Γ)D is an isomorphism. In the Galerkin boundary element method, the test space H1/2(Γ)D is replaced by some discrete subspace XL2(Γ)DH1/2(Γ)D. Again, the Lax–Milgram lemma guarantees existence and uniqueness of the solution ΦX of the discrete variational formulation (26) (VΦ;Ψ)=(f;Ψ)for all ΨX.(26) Moreover, Φ can in fact be computed by solving a linear system of equations. Note that (Equation21) implies that VΨH1(Γ)D for arbitrary ΨX. The additional regularity fH1(Γ)D instead of fH1/2(Γ)D is only needed to define the residual error estimator (Equation37) below. For a more detailed introduction to boundary integral equations, the reader is referred to the monographs [Citation19–21].

3. Axioms of adaptivity (revisited)

The aim of this section is to formulate an adaptive algorithm (Algorithm 3.3) for conforming BEM discretizations of our model problem (Equation24), where adaptivity is driven by the residual a posteriori error estimator (see (Equation37) below). We identify conditions for the underlying meshes, the mesh-refinement, as well as the boundary element spaces which ensure that the residual error estimator is reliable and fits into the general framework of [Citation6] and which hence guarantee optimal convergence behavior of the adaptive algorithm. We mention that we have already identified similar (but not identical) conditions for the finite element method in [Citation10, Section 3]. The main result of this work is Theorem 3.3 which is proved in Section 4.

3.1. Meshes

Throughout, T is a mesh of the boundary Γ=Ω of the bounded Lipschitz domain ΩRd in the following sense:

  • T is a finite set of compact Lipschitz domains on Γ, i.e. each element T has the form T=γT(Tˆ), where Tˆ is a compactFootnote2 Lipschitz domain in Rd1 and γT:TˆT is bi-Lipschitz.

  • T covers Γ, i.e. Γ=TTT.

  • For all T,TT with TT, the intersection TT has (d1)-dimensional Hausdorff measure zero.

We suppose that there is a countably infinite set T of admissible meshes. In order to ease notation, we introduce for TT the corresponding mesh-width function (27) hL(Γ)with h|T=hT:=|T|1/(d1)for all TT.(27) For ωΓ, we define the patches of order qN0 inductively by (28) π0(ω):=ω,πq(ω):={TT:Tπq1(ω)}.(28) The corresponding set of elements is (29) Πq(ω):={TT:Tπq(ω)},i.e., πq(ω)=Πq(ω).(29) If ω={z} for some zΓ, we simply write πq(z):=πq({z}) and Πq(z):=Πq({z}). For ST, we define πq(S):=πq(S) and Πq(S):=Πq(S). To abbreviate notation, the index q = 1 is omitted, e.g. π(ω):=π1(ω) and Π(ω):=Π1(ω).

We assume the existence of constants Cpatch,Clocuni,Cshape,Ccent,Csemi>0 such that the following assumptions are satisfied for all TT:

(M1)

Bounded element patch: The number of elements in a patch is uniformly bounded, i.e. #Π(T)Cpatchfor all TT.

(M2)

Local quasi-uniformity: Neighboring elements have comparable diameter, i.e. diam(T)/diam(T)Clocunifor all TT and all TΠ(T).

(M3)

Shape-regularity: It holds that Cshape1diam(T)/hTCshapefor all TT.

(M4)

Elements lie in the center of their patches: It holdsFootnote3 that diam(T)Ccentdist(T,Γπ(T))for all TT.

(M5)

Local seminorm estimate: For all vH1(Γ), it holds that |v|H1/2(π(z))Csemidiam(π(z))1/2|v|H1(π(z))for all zΓ.

The following proposition shows that (M5) is actually always satisfied. However, in general the multiplicative constant depends on the shape of the point patches. The proof is inspired by [Citation23, Proposition 2.2], where an analogous assertion for norms instead of seminorms is found. For σ=1/2 and d = 2, we have already shown the assertion in the recent own work [Citation16, Lemma 4.5]. For polyhedral domains Ω with triangular meshes, it is proved in [Citation24, Proposition 3.3] via interpolation techniques. A detailed proof for our setting is found in [Citation17, Proposition 5.2.2], where we essentially follow the proof of [Citation16, Lemma 4.5].

Proposition 3.1

Let ωˆRd1 be a bounded and connected Lipschitz domain and γω:ωˆωΓ be bi-Lipschitz, i.e. there exists a constant Clipref>0 such that (30) Clipref1|st||γω(s)γω(t)|diam(ω)Clipref|st|for all s, tωˆ.(30) Then, for arbitrary σ(0,1) there exists a constant Csemi(ωˆ)>0 such that (31) |v|Hσ(ω)Csemi(ωˆ)diam(ω)1σ|v|H1(ω)for all vH1(Γ).(31) The constant Csemi(ωˆ)>0 depends only on the dimension d, σ, the set ωˆ, and Clipref.

3.2. Mesh refinement

For TT and an arbitrary set of marked elements MT, we associate a corresponding refinement T:=refine(T,M)T with MTT, i.e. at least the marked elements are refined. Moreover, we suppose for the cardinalities that #T<#T if M and T=T else. Let refine(T)T be the set of all T such that there exist meshes T(0),,T(J) and marked elements M(0),,M(J1) with T=T(J)=refine(T(J1),M(J1)),,T(1)=refine(T(0),M(0)) and T(0)=T. We assume that there exists a fixed initial mesh T0T with T=refine(T0).

We suppose that there exist Cson2 and 0<ρson<1 such that all meshes TT satisfy for arbitrary marked elements MT with corresponding refinement T:=refine(T,M), the following elementary properties (R1)–(R3):

(R1)

Son estimate: One step of refinement leads to a bounded increase of elements, i.e. #TCson#T,

(R2)

Father is union of sons: Each element is the union of its successors, i.e. T={TT:TT}for all TT.

(R3)

Reduction of sons: Successors are uniformly smaller than their father, i.e. |T|ρson|T|for all TT and all TT with TT.

By induction and the definition of refine(T), one easily sees that (R2)–(R3) remain valid if T is an arbitrary mesh in refine(T). In particular, (R2)–(R3) imply that each refined element TTT is split into at least two sons, wherefore (32) #(TT)#T#Tfor all Trefine(T).(32) Besides (R1)–(R3), we suppose the following less trivial requirements (R2)–(R3) with generic constants Cclos,Cover>0:

(R4)

Closure estimate: Let (T)N0 be a sequence in T such that T+1=refine(T,M) with some MT for all N0. Then, it holds that #T#T0Cclosj=01#Mjfor all N0.

(R5)

Overlay property: For all T,TT, there exists a common refinement Trefine(T)refine(T) which satisfies the overlay estimate #TCover(#T#T0)+#T.

3.3. Boundary element space

With each TT, we associate a finite dimensional space of vector valued functions (33) XL2(Γ)DH1/2(Γ)D.(33) Let ΦX be the corresponding Galerkin approximation of φH1/2(Γ)D from (Equation24), i.e. (34) (VΦ;Ψ)=(f;Ψ)for all ΨX.(34) We note the Galerkin orthogonality (35) (fVΦ;Ψ)=0for all ΨX,(35) as well as the resulting Céa type quasi-optimality (36) φΦH1/2(Γ)CCeaminΨXφΨH1/2(Γ)with CCea:=Ccont/Cell.(36) We assume the existence of constants Cinv>0, qloc,qproj,qsuppN0, and 0<ρunity<1 such that the following properties (S1)–(S4) hold for all TT:

(S1)

Inverse inequality: For all ΨX, it holds that h1/2ΨL2(Γ)CinvΨH1/2(Γ).

(S2)

Nestedness: For all Trefine(T), it holds that XX.

(S3)

Local domain of definition: For all Trefine(T), TTΠqloc(TT)TT, and ΨX, it holds that Ψ|πqproj(T){Ψ|πqproj(T):ΨX}.

(S4)

Componentwise local approximation of unity: For all TT and all j{1,,D}, there exists some Ψ,T,jX with Tsupp(Ψ,T,j)πqsupp(T), such that only the jth component does not vanish, i.e. (Ψ,T,j)j=0for jj and 1(Ψ,T,j)jL2(supp(Ψ,T,j))ρunity|supp(Ψ,T,j)j|1/2.

Remark 3.2

Clearly, (S4) is in particular satisfied if X is a product space, i.e. X=j=1D(X)j, and each component (X)jL2(Γ) satisfies (S4).

Besides (S1)–(S4), we suppose that there exist constants Csz>0 as well as qszN0 such that for all TT and ST, there exists a linear Scott–Zhang-type operator J,S:L2(Γ)D{ΨX:Ψ|(TS)=0} with the following properties (S5)–(S6):

(S5)

Local projection property. Let qloc,qprojN0 from (S3). For all ψL2(Γ)D and TT with Πqloc(T)S, it holds that (J,Sψ)|T=ψ|Tif ψ|πqproj(T){Ψ|πqproj(T):ΨX}.

(S6)

Local L2-stability. For all ψL2(Γ)D and TT, it holds that J,SψL2(T)CszψL2(πqsz(T)).

3.4. Error estimator

Let TT. Due to the regularity assumption fH1(Γ)D, the mapping property (Equation21), and XL2(Γ)D, it holds that fVΨH1(Γ)D for all ΨX. This allows to employ the weighted-residual a posteriori error estimator (37a) η:=η(T)with η(S)2:=TSη(T)2for all ST,(37a) where the local refinement indicators read (37b) η(T)2:=hT|fVΦ|H1(T)2for all TT.(37b) The latter estimator goes back to the works [Citation25,Citation26], where reliability (Equation42) is proved for standard 2D BEM with piecewise polynomials on polygonal geometries, while the corresponding result for 3D BEM is found in [Citation15].

3.5. Adaptive algorithm

We consider the following concrete realization of the abstract algorithm from [Citation6, Algorithm 2.2].

3.6. Optimal convergence

Define (39) T(N):={TT:#T#T0N}for all NN0(39) and for all s>0 (40) Capprox(s):=supNN0minTT(N)(N+1)sη[0,].(40) We say that the solution φH1/2(Γ)D lies in the approximation class s with respect to the estimator if (41) φAsest:=Capprox(s)<.(41) By definition, φAsest< implies that the error estimator η on the optimal meshes T decays at least with rate O((#T)s). The following main theorem states that each possible rate s>0 is in fact realized by Algorithm 3.1. We stress that the range of possible s>0 depends, in particular, on the chosen admissible meshes T. The proof is given in Section 4. It essentially follows by verifying the axioms of adaptivity from [Citation6]. Such an optimality result was first proved in [Citation3] for the Laplace operator P=Δ on a polyhedral domain Ω. As ansatz space, they considered piecewise constants on shape-regular triangulations. [Citation1] in combination with [Citation5] extends the assertion to piecewise polynomials on shape-regular curvilinear triangulations of some piecewise smooth boundary Γ. Independently, [Citation4] proved the same result for globally smooth Γ and general self-adjoint and elliptic boundary integral operators.

Theorem 3.4

Let (T)N0 be the sequence of meshes generated by Algorithm 3.3. Then, there hold the following assertions (i)–(iii):

  1. Suppose (M1)–(M5), and (S4). Then, the residual error estimator satisfies reliability, i.e. there exists a constant Crel>0 such that (42) φΦH1/2(Γ)Crelηfor all  TT.(42)

  2. Suppose (M1)–(M5), (R2)–(R3), (S1)–(S2) and (S4). Then, for arbitrary 0<θ1 and Cmin[1,], the estimator converges linearly, i.e. there exist constants 0<ρlin<1 and Clin1 such that (43) η+j2Clinρlinjη2for all j,N0.(43)

  3. Suppose (M1)–(M5), (R1)–(R5), and (S1)–(S6). Then, there exists a constant 0<θopt1 such that for all 0<θ<θopt and Cmin[1,), the estimator converges at optimal rate, i.e. for all s>0 there exist constants copt,Copt>0 such that (44) coptφAsestsupN0(#T#T0+1)sηCoptφAsest,(44) where the lower bound requires only (R1) to hold.

All involved constants Crel,Clin,qlin,θopt, and Copt depend only on the assumptions made as well as the dimensions d,D, the coefficients of the differential operator P, and Γ, while Clin,ρlin depend additionally on θ and the sequence (Φ)N0, and Copt depends furthermore on Cmin, and s>0. The constant copt depends only on Cson,#T0,s, and if there exists 0 with η0=0, then also on 0 and η0.

Remark 3.5

If the sesquilinear form (V;) is Hermitian, then Clin, ρlin, and Copt are independent of (Φ)N0; see Remark 4.14 below.

Remark 3.6

Let Γ0Γ be an open subset of Γ=Ω and let E0:L2(Γ0)DL2(Γ)D denote the operator that extends a function defined on Γ0 to a function on Γ by zero. We define the space of restrictions H1/2(Γ0):={v|Γ0:vH1/2(Γ)} endowed with the quotient norm v0inf{vH1/2(Γ):v|Γ0=v0} and its dual space H~1/2(Γ0):=H1/2(Γ0). According to [Citation5, Section 2.1], E0 can be extended to an isometric operator E0:H~1/2(Γ0)DH1/2(Γ)D. Then, one can consider the integral equation (45) (VE0φ)|Γ0=f|Γ0,(45) where (VE0())|Γ0:H~1/2(Γ0)DH1/2(Γ0)D. In the literature, such problems are known as screen problems; see, e.g. [Citation21, Section 3.5.3]. Theorem 3.3 holds analogously for the screen problem (Equation45). Indeed, the works [Citation1,Citation3–5] cover this case as well. However, to ease the presentation, we focus on closed boundaries Γ0=Γ=Ω.

Remark 3.7

(a) Let us additionally assume that X contains all componentwise constant functions, i.e. (46) xXfor all xCD.(46) Then, under the assumption that hL(Ω)0 as , one can show that X:=N0X¯=H1/2(Γ)D. To see this, recall that H1/2(Γ)D is continuously and densely embedded in L2(Γ)D which is itself continuously and densely embedded in H1/2(Γ)D. For ψH1/2(Γ)D and arbitrary ε>0, let ψεH1/2(Γ)D with ψψεH1/2(Γ)ε. We abbreviate the projection operator J:=J,T for all N0. For all TT, the projection property (S5) in combination with our additional assumption (Equation46), the triangle inequality, and the local L2-stability (S6) show that (1J)ψεL2(T)=(S5)(1J)ψε1|πqsz(T)|πqsz(T)ψεdxL2(T)(S6)(1+Csz)ψε1|πqsz(T)|πqsz(T)ψεdxL2(πqsz(T)). With this, the Poincaré-type inequality from Lemma 4.6 below, and (M1)–(M3), we see that (1J)ψεL2(T)hT1/2|ψε|H1/2(πqsz(T))hL(Γ)1/2|ψε|H1/2(πqsz(T)). Summing over all elements, we obtain that (1J)ψεH1/2(Γ)2(1J)ψεL2(Γ)2hL(Γ)TT|ψε|H1/2(πqsz(T))2. With (M1)–(M4), Proposition 4.1 and Lemma 4.8 from below prove that TT|ψε|H1/2(πqsz(T))2|ψε|H1/2(Γ)2. Overall, this shows that minψXψψH1/2(Γ)ψψεH1/2(Γ)+(1J)ψεH1/2(Γ)ε+hL(Γ)1/2|ψε|H1/2(Γ). Since limhL(Γ)=0 and ϵ was arbitrary, this concludes the proof.

(b) The latter observation allows to follow the ideas of [Citation27] and to show that the adaptive algorithm yields convergence provided that the sesquilinear forms (;)P as well as (V;) are only coercive, i.e. elliptic up to some compact perturbation and that the continuous problem is well-posed; see also the introductory text of Section 4.4. This includes, e.g. adaptive BEM for the Helmholtz equation; see [Citation20, Section 6.9]. For details, the reader is referred to [Citation7,Citation27].

3.7. Application to BEM with piecewise polynomials on triangulations

For d = 2, 3, we fix the reference simplex Tref as the closed convex hull of the d vertices {0,e1,,ed1}. The convex hull of any d−1 vertices is called facet. A set T of subsets of Γ is called κ-shape regular triangulation if the following properties (i)–(v) are satisfied:

  1. T is a finite set of elements T of the form T=γT(Tˆ), where γT:TrefT is a bi-Lipschitz mapping whose Lipschitz constants are bounded from above by κ.

  2. T covers Γ, i.e. Γ=TTT.

  3. There are no hanging nodes in the sense that the intersection TT of any T,TT with TT is either empty or a common facet, i.e. TT=γT(f)=γT(f) for some facets f and f of Tref.

  4. The parametrizations of neighboring elements are compatible in the sense that for all nodes z (i.e. images of the {0,e1,,ed1} under an element map γT), there exists an interval π~(z) for d = 2 and a convex polygonal π~(z) for d = 3 respectively as well as a bijective and bi-Lipschitz continuous mapping γz:π~(z)π(z) such that γz1γT is affine for all TΠ(z).

  5. If d = 2, T is locally-quasi uniform in the sense that diam(T)κdiam(T) for all T,TT with TT.

Up to the fact that we allow γT to be bi-Lipschitz instead of C1, this definition is slightly stronger than [Citation5, Definition 2.4]. The property (iii) stems from [Citation21, Assumption 4.3.25] and is stronger than the corresponding assumption [Citation5, Definition 2.4 (iii)]. Further, (i) implies [Citation5, Definition 2.4 (v)], i.e. for all TT, there holds with the extremal eigenvalues λmin() and λmax() that (47) suptTrefdiam(T)2λmin(DγT(t)DγT(t))+λmax(DγT(t)DγT(t))diam(T)21;(47) see, e.g. [Citation24, (3.26)–(3.27)] or [Citation17, Lemma 5.2.1].

Let T0 be a κ0-shape regular triangulation. For d = 2, we define refine() as in [Citation28] via 1D-bisection in the parameter domain. For d = 3, we define refine() as in [Citation29] via newest vertex bisection in the parameter domain. In particular, all corresponding refinements TT=refine(T0) are again κ-shape regular triangulations with some fixed κ depending on κ0. We also note that the number of different π~(z) in (iv) is uniformly bounded, i.e. there exist only finitely many reference node patches. Finally, let pN0 be a fixed polynomial order. For each T, we associate the space of (transformed) piecewise polynomials (48) X:=Pp(T):={ΨL2(Γ):ΨγT is a polynomial of degree p for all TT}.(48) For this concrete setting, we already pointed out that [Citation1] in combination with [Citation5] proved linear convergence (Equation43) at optimal rate (Equation44) if P=Δ is the Laplace operator. The following theorem generalizes this result to arbitrary P as in Section 2.3.

Theorem 3.8

Piecewise polynomials on κ-shape regular triangulations satisfy the abstract properties (M1)–(M5), (R1)–(R5), and (S1)–(S6), where the constants depend only on the dimension D, the regularity constant κ, the initial mesh T, and the polynomial order p. By Theorem 3.4, this implies reliability (Equation42) of the error estimator and linear convergence (Equation43) at optimal rate (Equation44) for the adaptive strategy from Algorithm 3.3.

Proof.

The elementary mesh properties (M1)–(M3) are verified in [Citation5, Section 2.3]. (M4) is stated in [Citation5, Section 4.1]. (M5) follows from Proposition 3.1 together with the fact that there are only finitely many reference node patches.

For d = 2, the son estimate (R1) is clearly satisfied with Cson=2. For d = 3, it is well known that NVB satisfies (R1) with Cson=4. Further, (R1) holds true by definition. Reduction of sons (R3) is obviously satisfied in the parameter domain, i.e. |γT1(T)||γT1(T)|/2 for all TTrefine(T) with TT. Since γT is bi-Lipschitz, this property transfers to the physical domain, i.e. |γT1(T)|ρson|γT1(T)|, where 0<ρson<1 depends only on κ; see, e.g. [Citation17, Section 4.5.3] for details. For d = 2, (R4)–(R5) are found in [Citation28, Theorem 2.3]. For d = 3, the closure estimate (R4) is proved in [Citation30, Theorem 2.4], [Citation31, Theorem 6.1], or [Citation32, Theorem 2], where the latter result avoids any additional assumption on T0. The overlay property is proved in [Citation31, Proof of Lemma 5.2] or [Citation33, Section 2.2].

The inverse inequality (S1) for piecewise polynomials on the boundary is proved, e.g. in [Citation5, Lemma A.1]. Nestedness (S2) is trivially satisfied. Also (S3) is trivially satisfied with qloc,qproj=0. Clearly, (S4) holds with (Ψ,T,j)j:=0 for jj and (Ψ,T,j)j:=χT, where χT denotes the indicator function on T. Finally, for STT, we define with the elementwise L2(T)-orthogonal projection P,T:L2(T)D{Ψ|T:ΨX} (49) J,S:L2(Γ)X,ψJ,S:=P,Tψon all TS,0on all TTS.(49) This definition immediately yields (S5)–(S6) with qsz=0.

Remark 3.8

We mention that Theorem 3.8 is also valid if d = 2 and X is chosen as set of (transformed) splines which are piecewise polynomials with certain differentiability conditions at the break points. The required properties are (implicitly) verified in [Citation16]. As in [Citation10] (resp. [Citation11]), where we have verified the abstract FEM framework of [Citation10] for IGAFEM with hierarchical splines [Citation34] and the mesh-refinement from [Citation10] (resp. T-splines with the mesh-refinement from [Citation35]), the verification of the present abstract BEM framework for 3D IGABEM will be addressed in the future work [Citation12].

4. Proof of Theorem 3.4

In the following sections, we prove Theorem 3.4. Reliability (Equation42) is treated explicitly in Section 4.2. It follows immediately from an auxiliary result on the localization of the Sobolev–Slobodeckij norm which is investigated in Section 4.1. To prove Theorem 3.4 (ii)–(iii), we verify the following abstract properties (E1)–(E4) for the error estimator. Together with (R1), the closure estimate (R4), and the overlay property (R5), these already imply linear convergence of the estimator at optimal algebraic rate; see [Citation6].

There exist Cρ,Cqo, Cref, Cdrel, Cson, Cclos, Cover1, and 0ρred,εqo,εdrel<1 such that there hold:

(E1)

Stability on non-refined elements: For all TT and Trefine(T), it holds that |η(TT)η(TT)|ϱ,:=CρΦΦH1/2(Γ).

(E2)

Reduction on refined elements: For all TT and Trefine(T), it holds that η(TT)2ρredη(TT)2+ϱ,2.

(E3)

General quasi-orthogonality: It holds that 0εqo<supδ>01(1+δ)(1(1ρred)θ)2+δ1, and for all ,NN0, the sequence (T)N0 from Algorithm 3.3 satisfies that j=+N(ϱj,j+12εqoηj2)Cqoη2.

(E4)

Discrete reliability: For all TT and all Trefine(T), there exists TTR,T with #R,Cref(#T#T) such that ϱ,2εdrelη2+Cdrel2η(R,)2.

4.1. Localization of the Sobolev–Slobodeckij norm

Let TT. In contrast to the integer case, for σ(0,1), the norm Hσ(Γ) is not additive in the sense that vHσ(Γ)2TTvHσ(T)2for all vHσ(Γ)D. Although the upper bound ‘≲’ is in general false (see [Citation36, Section 3]), the lower bound ‘⪆’ can be proved elementarily for arbitrary vHσ(Γ)D.

Proposition 4.1

Let 0<σ<1 and TT. Then, (M1) implies the existence of a constant Csplit>0 such that for any vHσ(Γ)D, there holds that (50) TTTΠ(T)|v|Hσ(TT)2Csplit|v|Hσ(Γ)2.(50) The constant Csplit depends only on the constant from (M1)

Proof.

With the abbreviation (51) V(x,y):=|v(x)v(y)|2|xy|d1+2σfor all x,yΓ with xy,(51) (M1) shows that TTTΠ(T)|v|Hσ(TT)2=TTTΠ(T)|v|Hσ(T)2+2TTV(x,y)dxdy+|v|Hσ(T)2=2TTTΠ(T)TTV(x,y)dxdy+TTV(x,y)dxdy2(Cpatch+1)|v|Hσ(Γ)2. This concludes the proof.

However, if one replaces the elements T by some overlapping patches, then also the converse inequality is satisfied for functions vHσ(Γ)D which are L2-orthogonal to the ansatz space X.

Proposition 4.2

Let 0<σ<1 and TT. Then, (M1)–(M4) and (S4) imply the existence of a constant Csplit>0 such that for any vHσ(Γ)D which satisfies that (v;(Ψ,T,j)j)L2(Γ)=0 for all TT and all j{1,,D}, where Ψ,T,j are the functions from (S4), it holds that (52) vHσ(Γ)2CsplitTTTΠ(T)|v|Hσ(TT)2.(52) The constant Csplit depends only on the dimension d,σ, Γ, and the constants from (M1)–(M4) and (S4).

With this result, one can immediately construct a reliable and efficient error estimator, namely the so-called Faermann estimator; see Remark 4.10. For d = 2, the result of the proposition goes back to [Citation13], where X is chosen as space of splines transformed via the arclength parametrization γ:[a,b]Γ onto the one-dimensional boundary. In the recent own works [Citation37], we generalized the assertion to rational splines, where we could also drop the restriction that γ is the arclength parametrization. For d = 3, [Citation14] proved the result for discrete spaces which contain certain (transformed) polynomials of degree p{0,1,5,6} on a curvilinear triangulation of Γ. Our proof of Proposition 4.2 is inspired by [Citation14]. The key ingredient is the assumption (S4) which is exploited in Lemma 4.7. Before proving Proposition 4.2, we provide an easy corollary which is the key ingredient for the proof of reliability (Equation42).

Corollary 4.3

Let TT. Then, (M1)–(M5) and (S4) imply the existence of a constant Crel>0 such that for any vH1(Γ)D which satisfies that (v;Ψ,T,j)L2(Γ)=0 for all TT and all j{1,,D}, where Ψ,T,j are the functions from (S4), it holds that (53) vH1/2(Γ)Crelh1/2ΓvL2(Γ).(53) The constant Crel depends only on the dimension d, Γ, as well as the constants from (M1)–(M5) and (S4).

To prove Proposition 4.2, we start with the following basic estimate, which is proved in [Citation38, Lemma 8.2.4] or in [Citation17, Lemma 5.3.1].

Lemma 4.4

For all λ>0, there is a constant C(λ)>0 such that for all xRd and all ε>0, there holds that (54) ΓBε(x)|xy|d+1λdyC(λ)ελ.(54) The constant C(λ) depends only on the parameter λ, the dimension d, and Γ.

The following lemma is the first step towards the localization of the norm vHσ(Γ) for certain functions vHσ(Γ)D. In [Citation14, Lemma 3.1], this result is stated for triangular meshes. The elementary proof extends to our situation; see also [Citation17, Lemma 5.3.2] for details.

Lemma 4.5

Let 0<σ<1 and TT. Then, (M4) implies the existence of a constant C>0 such that for all vHσ(Γ)D, it holds that (55) vHσ(Γ)2TTTΠ(T)|v|Hσ(TT)2+CTTdiam(T)2σvL2(T)2.(55) The constant C depends only on the dimension d, σ,Γ, and the constant from (M4).

It remains to control the second summand in (Equation55). To this end, we need the following elementary Poincaré-type inequality of [Citation13, Lemma 2.5].

Lemma 4.6

For any σ(0,1) and any measurable ωΓ, there holds for all vHσ(ω) that (56) vL2(ω)2diam(ω)d1+2σ2|ω||v|Hσ(ω)2+1|ω|ωv(x)dx2.(56)

We start to estimate the second summand in (Equation55).

Lemma 4.7

Let σ(0,1), TT and TT. Then, (M1)–(M3) and (S4) imply the existence of a constant C>0 such that for all vHσ(Γ)D with (vj;Ψ,T,j)L2(Γ)=0 for all j{1,,D}, where Ψ,T,j are the functions from (S4), it holds that (57) hσvL2(T)C|v|Hσ(πqsupp(T)),(57) where qsupp is the constant from (S4). The constant C depends only on the dimension d, σ,Γ, and the constants from (M1)–(M3) and (S4).

Proof.

We prove (Equation57) for each component vj of v, where j{1,,D}. Then, squaring and summing up all components, we conclude the proof. (S4) and Lemma 4.6 show that (58) vjL2(T)2vjL2(supp(Ψ,T,j))2diam(supp(Ψ,T,j))d1+2σ2|supp(Ψ,T,j)||vj|Hσ(supp(Ψ,T,j))2+1|supp(Ψ,T,j)|supp(Ψ,T,j)vj(x)dx2.(58) Now, we apply the orthogonality and (S4) to get for the second summand that 1|supp(Ψ,T,j)|supp(Ψ,T,j)vj(x)dx2=1|supp(Ψ,T,j)|supp(Ψ,T,j)v¯j(x)(1Ψ,T,j(x))dx21|supp(Ψ,T,j)|vjL2(supp(Ψ,T,j))21(Ψ,T,j)jL2(supp(Ψ,T,j))2ρunity2vjL2(supp(Ψ,T,j)2. Inserting this in (Equation58) gives that (59) (1ρunity2)vjL2(supp(Ψ,T,j))2diam(supp(Ψ,T,j))d1+2σ2|supp(Ψ,T,j)||vj|Hσ(supp(Ψ,T,j))2.(59) With (S4) and (M1)–(M3), we see that diam(supp(Ψ,T,j))diam(πqsupp(T))diam(T)hT. Further, (S4) implies that |supp(Ψ,T,j)||T|=hTd1. Inserting this in (Equation59) and using again (S4), we derive that vjL2(T)2vjL2(supp(Ψ,T,j))2hT2σ|vj|Hσ(supp(Ψ,T,j))2hT2σ|vj|Hσ(πqsupp))2. Altogether, this concludes the proof.

The following lemma allows us to further estimate the term |v|Hσ(πqsupp(T)) of (Equation57).

Lemma 4.8

Let qN0 and TT. Then, (M1)–(M4) imply the existence of a constant C(q)>0 such that for all vHσ(Γ)D and all TT there holds that (60) |v|Hσ(πq(T))2C(q)T,TΠq(T)TT|v|Hσ(TT)2.(60) The constant depends only on the dimension d,σ,q, and the constants from (M1)–(M4).

Proof.

Without loss of generality, we may assume that D = 1. We prove the assertion in two steps.

Step 1: Let T0,T1,,Tm be a chain of elements in Πq(T) with TiTj= for |ij|>1 and TiTj if |ij|=1, where 1mq. We set Tij:=k=ijT for ij and prove by induction on m that there exists a constant C1(m)>0 which depends only on d,σ,q,m, and (M2)–(M4), such that (61) |v|Hσ(T0m)2C1(m)i=0m1|v|Hσ(TiTi+1)2.(61) For m = 1, (Equation61) with C1(1)=1 even holds with equality. Thus the induction hypothesis reads: for all 1m1<q and for any chain T0,,Tm1 of elements in Πq(T), it holds that (62) |v|Hσ(T0m1)2C1(m1)i=0m2|v|Hσ(TiTi+1)2.(62) Let TmΠq(T) with TmTi= for im2 and TmTi for i = m−1. For all x,yΓ,xy, we abbreviate V(x,y):=|v(x)v(y)|2|xy|d1+2σ. The definition (Equation7) of the Sobolev–Slobodeckij seminorm shows that |v|Hσ(T0m)2=T0mT0mV(x,y)dxdy=T0m1T0m1V(x,y)dxdy+TmTmV(x,y)dxdy+2TmT0m1V(x,y)ddxdy=|v|Hσ(T0m1)2+|v|Hσ(Tm)2+2TmT0m2V(x,y)dxdy+2TmTm1V(x,y)dxdy|v|Hσ(T0m1)2+|v|Hσ(Tm1Tm)2+2TmT0m2V(x,y)dxdy. With the induction hypothesis (Equation62), it remains to estimate TmT0m2V(x,y)dxdy. First, we note that for xT0m2,yTm,zTm1, it holds that (63) V(x,y)=|v(x)v(y)|2|xy|d1+2σ2|v(x)v(z)|2|xy|d1+2σ+2|v(z)v(y)|2|xy|d1+2σ.(63) Moreover, (M4) shows that |xy|dist(Tm,Γπ(Tm))diam(Tm). Since x,y,zT0m, (M2) shows max{|xz|,|yz|}diam(Tm). Hence, we can proceed the estimate of (Equation63) V(x,y)V(x,z)+V(z,y). This implies that TmT0m2V(x,y)dxdy=1|Tm1|Tm1TmT0m2V(x,y)dxdydz1|Tm1|Tm1TmT0m2V(x,z)+V(y,z)dxdydz=1|Tm1|Tm1T0m2|Tm|V(x,z)dxdz+Tm1Tm1|T0m2|V(y,z)dydzmax{|Tm|,|T0m2|}|Tm1||v|Hσ(T0m1)2+|v|Hσ(Tm1Tm)2. Note that max{|Tm|,|T0m2|}/|Tm1|1 by (M2)–(M3). Together with the induction hypothesis (Equation62), this concludes the induction step.

Step 2: We come to the assertion itself. By definition, we have that |v|H1/2(πq(T))2=T~,T~Πq(T)T~T~V(x,y)dxdy. Let T~,T~Πq(T). First, we suppose that T~T~=. Then, there exists a chain as in Step 1 with T~=T0 and T~=Tm. Step 1 proves that T~T~V(x,y)dxdy|v|Hσ(T0m)2T,TΠq(T)TT|v|Hσ(TT)2. If T~=T~, the same estimate holds true. Since the number of T~,T~Πq(T) is uniformly bounded by a constant, which depends only on the constant of (M1) and q, this estimate concludes the proof.

With the property (M5), one immediately derives the following Poincaré inequality.

Proposition 4.9

Let TT and TT. Then, (M1)–(M5) and (S4) imply the existence of a constant Cpoinc>0 such that for all vH1(Γ)D which satisfy that (v;Ψ,T,j)L2(Γ)=0 for all j{1,,D}, where Ψ,T,j are the functions from (S4), it holds that (64) h1vL2(T)Cpoinc|v|H1πqsupp+1(T),(64) where qsupp is the constant from (S4). The constant Cpoinc depends only on the dimension d,Γ, and the constants from (M1)–(M5) and (S4).

Proof.

We apply Lemma 4.7 and Lemma 4.8 to see that h1/2vL2(T)2|v|H1/2(πqsupp(T))2T,TΠqsupp(T)TT|v|H1/2(TT)2. For T,TT with TT, we fix some point z(T,T)TT. With (M5), we continue our estimate h1/2vL2(T)2|v|H1/2(πqsupp(T))2T,TΠqsupp(T)TT|v|H1/2(π(z(T,T))2T,TΠqsupp(T)TTdiamπ(z(T,T)ΓvL2(π(T))2.(M1)–(M3) imply that hTh on πqsupp+1(T), and that the last term of the latter estimate can be bounded from above (up to a multiplicative constant) by h1/2ΓvL2(πqsupp+1(T))2. This concludes the proof.

With all the preparations, we can finally prove the main result of this section.

Proof

Proof of Proposition 4.2

Together with (M3), Lemma 4.5 proves that vHσ(Γ)2TTTΠ(T)|v|Hσ(TT)2+TThT2σvL2(T)2. It remains to estimate the second sum. With Lemma 4.7 and Lemma 4.8, we see that (65) TThT2σvL2(T)2TT|v|Hσ(πqsupp(T))2TTT,TΠqsupp(T)TT|v|Hσ(TT)2.(65) If TT and T,TΠqsupp(T) with TT, then TΠqsupp(T) and TΠ(T). Plugging this into (Equation65) shows that TThT2σvL2(T)2TTTΠqsupp(T)TΠ(T)|v|Hσ(TT)2, and #Πqsupp(T)1 (see (M1)) concludes the proof.

4.2. Reliability (42)

Let TT. Recall that V:H1/2(Γ)DH1/2(Γ)D is an isomorphism. Due to Galerkin orthogonality (Equation35), Corollary 4.3 leads to (66) φΦH1/2(Γ)fVΦH1/2(Γ)h1/2Γ(fVΦ)L2(Γ)=η.(66)

Remark 4.10

Proposition 4.1 and Proposition 4.2 show that (67) φΦH1/2(Γ)2fVΦH1/2(Γ)2TTTΠ(T)|fVΦ|H1/2(TT)2.(67) This is even true for arbitrary fH1/2(Γ)D without the additional restriction fH1(Γ)D. In particular, (68) ϝ(T)2:=TΠ(T)|fVΦ|H1/2(TT)2for all TT(68) provides a local error indicator. The corresponding error estimator ϝ is often referred to as Faermann estimator. In BEM, it is the only known estimator which is reliable and efficient (without further assumptions as, e.g. the saturation assumption [Citation39, Section 1]). Obviously, one could replace the residual estimator η in Algorithm 3.3 by ϝ. However, due to the lack of an h-weighting factor, it is unclear whether the reduction property (E2) of Section 4.2 is satisfied. [Citation24, Theorem 7] proves at least plain convergence of ϝ even for fH1/2(Γ)D if one uses piecewise constants on affine triangulations of Γ as ansatz space. The proof immediately extends to our current situation, where the assumptions (M1)–(M5), (R2)–(R3), and (S1)–(S2) are employed. The key ingredient is the construction of an equivalent mesh-size function h~L(Γ) which is contractive on each element which touches a refined element, i.e. there exists a uniform constant 0<ρctr<1 such that (69) h~|Tρctrh~|Tfor all Trefine(T) and all  TΠ(TT).(69) The existence of such a mesh-size function is proved in [Citation6, Section 8.7] for shape-regular triangular meshes. The proof works verbatim for the present setting.

4.3. Convergence of Φ+1ΦH1/2(Γ)

Nestedness (S2) ensures that X:=N0X¯ is a closed subspace of H1/2(Γ)D and hence admits a unique Galerkin solution ΦX. Note that Φ is also a Galerkin approximation of Φ. Hence, the Céa lemma (Equation36) with φ replaced by Φ proves that ΦΦH1/2(Γ)0 as . In particular, we obtain that limΦ+1ΦH1/2(Γ)=0.

4.4. An inverse inequality for V

In Proposition 4.13, we establish an inverse inequality for the single-layer operator V. Throughout this section, neither the ellipticity of P nor the ellipticity of V are exploited (and we can drop these assumption here). Indeed, it is sufficient to assume that P is only coercive. Then, the definitions and properties presented in Section 2.3 remain valid; see [Citation19, page 119]. For the Laplace operator P=Δ, an inverse estimate as in Proposition 4.13 was already proved in [Citation3, Theorem 3.1] for shape-regular triangulations of a polyhedral boundary Γ. Independently, [Citation4] derived a similar result for globally smooth Γ and arbitrary self-adjoint and elliptic boundary integral operators. In [Citation5, Theorem 3.1], [Citation3, Theorem 3.1] is generalized to piecewise polynomial ansatz functions on shape-regular curvilinear triangulations. In particular, our Proposition 4.13 does not only extend these results to arbitrary general meshes as in Section 3.1, but is also completely novel for, e.g. linear elasticity. The proof follows the lines of [Citation5, Section 4]. We start with the following lemma, which was proved in [Citation40, Theorem 4.1] on shape-regular triangulations. With Lemma 4.6, the proof immediately extends to our situation; see also [Citation17, Lemma 5.3.11].

Lemma 4.11

For TT, let P0(T)DL2(Γ)D be the set of all functions whose D components are T-piecewise constant functions on Γ. Let P:L2(Γ)DP0(T)D be the corresponding L2-projection. Then, (M1) and (M3) imply for arbitrary 0<σ<1 the existence of a constant C>0 such that (70) (1P)ψHσ(Γ)ChσψL2(Γ)for all ψL2(Γ).(70) The constant C depends only on the dimension D, the boundary Γ,σ, and the constants from (M3).

In contrast to [Citation5], we cannot use the Caccioppoli-type inequality from [Citation41, Lemma 5.7.1] which is only shown for the Poisson problem there. Therefore, we prove the following generalization. For an open set ORd and an arbitrary uH2(O), we abbreviate |u|H1(O):=uL2(O) and |u|H2(O):=i=1d|iu|H1(O)21/2.

Lemma 4.12

Let r>0, xRd, and uH1(B2r(x))D be a weak solution of Pu=0. Then, u|Br(x)C(Br(x))D and there exists a constant C>0 such that (71) |u|H2(Br(x))CuL2(B2r(x))+1+r+r2r|u|H1(B2r(x)).(71) The constant C depends only on the dimensions d,D, and the coefficients of the partial differential operator P.

Proof.

By [Citation19, Theorem 4.16], there holds that u|B3r/2(x)Hk(B3r/2(x))D for all kN0, and the Sobolev embedding theorem proves that u|B3r/2(x)C(B3r/2(x))D. In particular, u is a strong solution of Pu=0 on B3r/2(x). To prove (Equation71), let λRD be an arbitrary constant vector, and define u~:=uϕ with the affine bijection ϕ:B3/2(0)B3r/2(x), ϕ(y~)=ry~+x for y~B3/2(0). Since the coefficients of P are constant and u is a strong solution, there holds for all y~B3/2(0) with y:=ϕ(y~) that (72) i=1di=1di(Aiii(u~λ))(y~)=i=1di=1di(Aiii(uλ))(y)r2=r2i=1dbii(uλ)(y)+c(uλ)(y)+cλ.(72) We define the right-hand side as f~C(B3/2(0)), i.e. f~(y~):=r2i=1dbii(uλ)(ϕ(y~))+c(uλ)(ϕ(y~))+cλ. This shows that u~λ is a strong (and thus weak) solution of a coercive system of second-order PDEs with smooth coefficients and smooth right-hand side. The application of [Citation19, Theorem 4.16] yields the existence of a constant C1>0, which depends only on d,D, and the coefficients of the matrices Aii, such that (73) |u~λ|H2(B1(0))C1u~λH1(B3/2(0))+f~L2(B3/2(0)).(73) Standard scaling arguments prove that |u~λ|H2(B1(0))r2rd/2|u|H2(Br(x)),u~λL2(B3/2(0))1rd/2uλL2(B3r/2(x)),|u~λ|H1(B3/2(0))rrd/2|u|H1(B3r/2(x)),f~L2(B3/2(0))r2rd/2|u|H1(B3r/2(x))+r2rd/2uλL2(B3r/2(x))+r2|λ|. Plugging this into (Equation73), we obtain that (74) |u|H2(Br(x))1+r2r2uλL2(B3r/2(x))+1+rr|u|H1(B3r/2(x))+rd/2|λ|.(74) We choose λ as the integral mean λ:=B3r/2(x)u(y)dy/|B3r/2(x)|. The Cauchy–Schwarz inequality implies that |λ|uL1(B3r/2(x))/|B3r/2(x)|uL2(B3r/2(x))/|B3r/2(x)|1/2rd/2uL2(B3r/2(x)). Using this and the Poincaré inequality in (Equation74), we see that |u|H2(Br(x))uL2(B3r/2(x))+1+r+r2r|u|H1(B3r/2(x)). Together with the fact that B3r/2(x)B2r(x), this concludes the proof.

For the proof of the next proposition, we need the linear and continuous single-layer potential from [Citation19, Theorem 6.11] (75) V~:H1/2(Γ)DH1(U)D,(75) where U is an arbitrary bounded domain with ΓU. The single-layer operator V is just the trace of V~, i.e. (76) V=V~()|Γ:H1/2(Γ)DH1/2(Γ)D;(76) see [Citation19, page 219–220]. Indeed, for ψL(Γ), [Citation19, page 201–202] states the following integral representation: (77) (V~ψ)(x)=ΓG(xy)ψ(y)dyfor all xU.(77)

Proposition 4.13

Suppose (M1)–(M5). For TT, let wL(Γ) be a weight function which satisfies for some α>0 and all TT that (78) wL(T)αw(x)for almost all xπ(T).(78) Then, there exists a constant Cinv,V>0 such that for all ψL2(Γ)D, it holds that (79) wΓVψL2(Γ)Cinv,Vw/h1/2L(Γ)ψH1/2(Γ)+wψL2(Γ).(79) The constant Cinv,V depends only on (M1)–(M5), Γ, the coefficients of P, and the admissibility constant α. The particular choice w=h1/2 shows that (80) h1/2ΓVψL2(Γ)Cinv,VψH1/2(Γ)+h1/2ψL2(Γ).(80)

Proof.

The proof works essentially as in [Citation5, Section 4]. Therefore, we mainly emphasize the differences and refer to [Citation17, Proposition 5.3.15] for further details.

By (M4) and with the abbreviation δ1(T):=diam(T)/(2Ccent) and UT:=Bδ1(T)(T), there holds for all TT that UTΓB2δ1(T)(T)Γπ(T). This provides us with an open covering of ΓTTUT. We show that this is even locally finite in the sense that there exists a constant C>0 with #{TT:xUT}C for all xRd: Let xRd. Clearly, we may assume that xTTUT. Choose T0T with xUT0 such that δ1(T0) is minimal, and let x0T0 with |xx0|<δ1(T0). If TT with xUT, the triangle inequality yields that dist({x0},T)<2δ1(T). By choice of δ1(T), (M4) hence yields that x0π(T). Thus, {TT:xUT}{TT:x0π(T)}, and (M1) implies that (81) #{TT:xUT}#{TT:x0π(T)}Cpatch2.(81) We fix (independently of T) a bounded domain URd with UTU for all TT. We define for TT the near-field and the far-field of uV:=V~ψ by (82) uV,Tnear:=V~(ψχΓUT)anduV,Tfar:=V~(ψχΓUT).(82) In the following five steps, the near-field and the far-field are estimated separately. The first two steps deal with the near-field, whereas the last three steps deal with the far-field.

Step 1: As in [Citation5, Lemma 4.1], one shows that for all TT, all T-piecewise (componentwise) constant functions ΨTP0(T)D with supp(ΨT)π(T) satisfy that (83) V~ΨTH1(UT)h1/2ΨTL2(π(T)).(83) The proof of [Citation5] uses only (Equation81) and the fact that |xG(xy)||xy|d+1 (x,y)U×Γ with xy for the Laplacian fundamental solution G. However, according to [Citation19, Theorem 6.3 and Corollary 6.5] this fact is also valid for general coercive second-order PDEs with C coefficients. Moreover, [Citation5] bounds only the H1-seminorm in (Equation83), but the L2-norm can be bounded similarly due to |G(xy)|max{|log|xy||,|xy|d+2}|xy|d+1 (see again [Citation19, Theorem 6.3 and Corollary 6.5]).

Step 2: With Step 1, one shows as in [Citation5, Proposition 4.2] that uV,TnearH1(U) and uV,Tnear|ΓH1(Γ) with (84) TTwΓuV,TnearL2(T)2+TTw/h1/2L(T)2uV,TnearH1(UT)2wψL2(Γ)2.(84) In the proof, one applies the stability of V:L2(Γ)DH1(Γ)D (see (Equation21)) and V~:H1/2(Γ)DH1(U)D (see (Equation75)). Moreover, the approximation property (Equation70) is exploited by splitting ψχΓUT=P(ψχΓUT)+(1P)(ψχΓUT) and choosing ΨT:=P(ψχγUT). Note that [Citation5] only proves (Equation84) with |uV,Tnear|H1(UT)2 instead of uV,TnearH1(UT)2.

Step 3: We consider the far-field. We set Ωext:=RdΩ¯. According to [Citation19, Theorem 6.11], for all TT, the potential uV,Tfar is a solution of the transmission problem PuV,Tfar=0on ΩΩext,[uV,Tfar]Γ=0in H1/2(Γ)D,[DνuV,Tfar]Γ=ψχΓUTin H1/2(Γ)D, where []Γ and  [Dν()]Γ denote the jump of the traces and the conormal derivatives respectively (see [Citation19, page 117] for a precise definition) across the boundary Γ. Twofold integration by parts that uses these jump conditions shows that PuV,Tfar=0 weakly on UT. Since Bδ1(T)(x)UT for all xT, Lemma 4.12 shows that uV,TfarC(Bδ1(T)/2(x)) with (85) |uV,Tfar|H2(Bδ1(T)/2(x))uV,TfarL2(Bδ1(T)(x))+diam(T)1|uV,Tfar|H1(Bδ1(T)(x)).(85) Note that [Citation5] proves (Equation85) even without the term uV,TfarL2(Bδ1(T)(x)). Indeed, since the kernel of the Laplace operator contains all constants, [Citation5] employs a Poincaré inequality to bound uV,TfarL2(Bδ1(T)(x)) by diam(T)1|uV,Tfar|H1(Bδ1(T)(x)).

Step 4: With inequality (Equation85) at hand, one can prove the following local far-field bound for the single-layer potential V~ (86) h1/2ΓuV,TfarL2(T)h1/2uV,TfarL2(T)uV,TfarH1(UT).(86) The proof works as in [Citation5, Lemma 4.4] and relies on a standard trace inequality on Ω, the Caccioppoli inequality (Equation85) as well as the Besicovitch covering theorem. Note that in [Citation5], the estimates (Equation85) and thus (Equation86) even hold without the L2-norm of uV,Tfar, since the Laplace problem is considered.

Step 5: Finally, (Equation81), (Equation82), (Equation84), (Equation86), and the stability of V~:H1/2(Γ)DH1(U)D (see (Equation75)) easily lead to the far-field bound for V~ (87) TTwΓuV,TfarL2(T)2TTwuV,TfarL2(T)2w/h1/2L(Γ)2ψH1/2(Γ)2+wψL2(Γ)2.(87) For the simple proof, we refer to [Citation5, Proposition 4.5]. By definition (Equation82), (Equation87) together with (Equation84) from Step 2 concludes the proof.

[Citation5] does not only treat the single-layer operator V:H1/2(Γ)DH1/2(Γ)D but also derives similar inverse estimates as in (Equation79) for the double-layer operator, the adjoint double-layer operator, and the hyper-singular operator. With similar techniques as in Proposition 4.13, we will also generalize this result in Appendix B. However, we will indeed only need the inverse estimate (Equation80) for the single-layer operator in the remainder of the paper.

4.5. Stability on non-refined elements (E1)

We show that the assumptions (M1)–(M5) and (S1)–(S2) imply stability (E1), i.e. the existence of Cstab1 such that for all TT, and all Trefine(T), it holds that (88) |η(TT)η(TT)|CstabΦΦH1/2(Γ).(88) In Section 4.6, we will fix the constant Cϱ for ϱ, defined in (E1) such that CstabCϱ. The reverse triangle inequality and the fact that h=h on ω:=(TT) prove that |η(TT)η(TT)|=h1/2ΓV(φΦ)L2(ω)h1/2ΓV(φΦ)L2(ω)h1/2ΓV(ΦΦ)L2(ω)h1/2ΓV(ΦΦ)L2(Γ). (S2) shows that ΦΦX. Therefore, the inverse inequalities from (S1) and (Equation80) are applicable, which implies (Equation88). The constant Cstab depends only on d, D, Γ, the coefficients of P, and the constants from (M1)–(M5) and (S1).

4.6. Reduction on refined elements (E2)

We show that the assumptions (M1)–(M5), (R2)–(R3), and (S1)–(S2) imply reduction on refined elements (E2), i.e. the existence of Cred1 and 0<ρred<1 such that for all TT and all Trefine(T), it holds that (89) η(TT)2ρredη(TT)2+CredΦΦH1/2(Γ)2.(89) With this, we can fix the constant for ϱ, defined in (E1) as (90) Cϱ:=max{Cstab,Cred1/2}.(90) Let ω:=(TT). First, we apply the triangle inequality η(TT)=h1/2ΓV(φΦ)L2(ω)h1/2ΓV(φΦ)L2(ω)+h1/2ΓV(ΦΦ)L2(ω). Clearly, (R2)–(R3) show that ω=(TT)=(TT) and hρson1/(d1)h on ω. Thus we can proceed the estimate as follows: η(TT)ρson1/(2d2)h1/2ΓV(φΦ)L2(ω)+h1/2ΓV(ΦΦ)L2(ω)=ρson1/(2d2)η(TT)+h1/2ΓV(ΦΦ)L2(ω). Since ΦXX according to (S2), we can apply the inverse estimates (S1) and (Equation80). Together with the Young inequality, we derive for arbitrary δ>0 that η(TT)2(1+δ)ρson1/(d1)η(TT)2+(1+δ1)Cinv,V2(1+Cinv)2ΦΦH1/2(Γ)2. Choosing δ>0 sufficiently small, we obtain (Equation89). The constant Cred depends only on d, D, Γ, the coefficients of P, and the constants from (M1)–(M5), (R2)–(R3), and (S1).

4.7. General quasi-orthogonality (E3)

According to [Citation6, Section 4.3], Section 4.3, Section 4.5, and Section 4.6 already imply estimator convergence limη=0. Therefore, reliability (Equation42) implies error convergence limφΦH1/2(Γ)=0. In particular, we obtain that φX=N0X¯. Recall that we have already fixed the constant Cϱ in (Equation90). We introduce the principal part of P as the corresponding partial differential operator without lower-order terms (91) P0v:=i=1di=1di(Aiiiv).(91) According to [Citation19, Lemma 4.5], the principal part is coercive on H01(Ω)D. We denote its corresponding single-layer operator which can be defined as in Section 2.3 by V0:H1/2(Γ)DH1/2(Γ)D. Our assumption Aii=Aii¯ easily implies that V0 is self-adjoint; see, e.g. [Citation19, page 218]. With (Equation76) and (EquationA14) below, we particularly see the stability of VV0:H1/2(Γ)DH1ε(Γ)D for all ε>0. Thus the Rellich compactness theorem [Citation19, Theorem 3.27] implies that VV0:H1/2(Γ)DH1/2(Γ)D is compact. This yields that V is an elliptic operator which can be written as the sum of a self-adjoint operator V0 plus a compact operator VV0. From [Citation27,Citation42], we thus derive the general quasi-orthogonality (E3) (see also [Citation17, Section 4.4.3] for all details), i.e. the existence of (92) 0εqo<supδ>01(1+δ)(1(1ρred)θ)2+δ1(92) and Cqo1 such that (93) j=+N(CϱΦj+1ΦjH1/2(Γ)2εqoηj2)Cqoη2for all ,NN0.(93)

Remark 4.14

If the sesquilinear form (V;) is Hermitian, (Equation93) follows from the Pythagoras theorem φΦjV2+Φj+1ΦjV2=φΦjV2 and norm equivalence j=+NΦj+1ΦjH1/2(Γ)2j=+NΦj+1ΦjV2=φΦV2φΦ+NV2φΦH1/2(Γ)2. Together with reliability (Equation42), this proves (Equation93) even for εqo=0, and Cqo is independent of the sequence (Φ)N0.

4.8. Discrete reliability (E4)

The proof of (E4) is inspired by [Citation3, Proposition 5.3] which considers piecewise constants on shape-regular triangulations as ansatz space. Under the assumptions (M1)–(M5), (Equation32), and (S1)–(S6), we show that there exist Cdrel,Cref1 such that for all TT and all Trefine(T), the subset (94) R,:=Πqsupp+qloc+2(TT)(94) satisfies that CϱΦΦH1/2(Γ)Cdrelη(R,),TTR,, and #R,Cref(#T#T). The last two properties are obvious with Cref=Cpatchqsupp+qloc+2 by validity of (M1) and (Equation32). The first estimate is proved in three steps:

Step 1: For S1:=TT, let J,S1 be the corresponding projection operator from (S5)–(S6). Ellipticity (Equation22), nestedness (S2) of the ansatz spaces, and the definition (Equation34) of the Galerkin approximations yield that ΦΦH1/2(Γ)2Re(V(ΦΦ);ΦΦ)L2(Γ)=Re(V(φΦ);(1J,S1)(ΦΦ))L2(Γ). (S3) shows that (ΦΦ)|πproj(T){Ψ|πproj(T):ΨX} for any TTΠqloc(TT). Moreover, one easily sees that (95) Πqloc(T)TT=S1for all TTΠqloc(TT).(95) Hence, the local projection property (S5) of J,S1 is applicable and proves that J,S1(ΦΦ)=ΦΦ on Γπqloc(TT). With S2:=Πqloc(TT), Lemma A.1 provides a smooth cut-off function χ:=χ~S2H1(Γ) with 0χ1 such that χ=1 on S2, χ=0 on Γπ(S2), and |Γχ|h1, where the hidden constant depends only on d and (M1)–(M4). This leads to (96) ΦΦH1/2(Γ)2Re(χV(φΦ);(1J,S1)(ΦΦ))L2(Γ).(96) We bound the two terms I:=Re(χV(φΦ);ΦΦ)L2(Γ) and II:=Re(χV(φΦ);J,S1(ΦΦ))L2(Γ) separately. Since H1/2(Γ)D is the dual space of H1/2(Γ)D, it holds that (97) IχV(φΦ)H1/2(Γ)ΦΦH1/2(Γ).(97) The Cauchy–Schwarz inequality shows that IIh1/2χV(φΦ)L2(Γ)h1/2J,S1(ΦΦ)L2(Γ). Since J,S1:L2(Γ)D{ΨX:Ψ|(TS1)=0}, it holds that supp(J,S1(ΦΦ))(TT). This together with the fact that h=h on (TT), the local L2-stability (S6) and (M1)–(M3) implies that II=h1/2χV(φΦ)L2(Γ)h1/2J,S1(ΦΦ)L2((TT))h1/2χV(φΦ)L2(Γ)h1/2(ΦΦ)L2(Γ). With the inverse inequality (S1) applied to ΦΦX (see (S2)), the latter estimate implies that (98) IIh1/2χV(φΦ)L2(Γ)ΦΦH1/2(Γ).(98) Plugging (Equation97)–(Equation98) into (Equation96) shows that (99) ΦΦH1/2(Γ)h1/2χV(φΦ)L2(Γ)+χV(φΦ)H1/2(Γ).(99)

Step 2: Next, we deal with the first summand of (Equation99). With supp(χ)πqloc+1(TT) and 0χ1, this implies that (100) h1/2χV(φΦ)L2(Γ)h1/2V(φΦ)L2πqloc+1(TT).(100) By Galerkin orthogonality (Equation35), we see that V(φΦ) is L2-orthogonal to all functions of X which includes in particular the functions Ψ,T,j from (S4). Hence, we can apply Proposition 4.9. Together with (M1)–(M3) and recalling (Equation94), (Equation100) proves that h1/2χV(φΦ)L2(Γ)h1/2ΓV(φΦ)L2πqsupp+qloc+2(TT)=η(R,). Step 3: It remains to consider the second summand of (Equation99). Lemma 4.5 in conjunction with shape-regularity (M3) implies that χV(φΦ)H1/2(Γ)2TTTΠ(T)|χV(φΦ)|H1/2(TT)2+h1/2χV(φΦ)L2(Γ). We have already dealt with the second summand in Step 2 (see (Equation100)). For the first one, we fix again some z(T,T)TT for any TT,TΠ(T). (M1)–(M3) and (M5) show that TTTΠ(T)|χV(φΦ)|H1/2(TT)2TTTΠ(T)|χV(φΦ)|H1/2(π(z(T,T)))2TTTΠ(T)h1/2Γ(χV(φΦ))L2(π(z(T,T))2h1/2Γ(χV(φΦ))L2(Γ)2. With the product rule and (EquationA2), we continue our estimate χV(φΦ)H1/2(Γ)2h1/2V(φΦ))L2(supp(χ))2+h1/2ΓV(φΦ))L2(supp(χ))2. Note that we have already dealt with the first summand in Step 2 (see (Equation100)). Finally, supp(χ)πqloc+1(TT) and Πqloc+1(TT)R, (see (Equation94)) prove for the second one that h1/2ΓV(φΦ))L2(supp(χ))2η(R,)2. With this, we conclude the proof of discrete reliability (E4). The constant Cdrel depends only on Cϱ,d,D, Γ, and the constants from (M1)–(M5) and (S1)–(S6).

Acknowledgments

The authors acknowledge support through the Austrian Science Fund (FWF) under grant J4379-N,P29096,W1245.

Disclosure statement

No potential conflict of interest was reported by the authors.

Additional information

Funding

The authors acknowledge support through the Austrian Science Fund (FWF) under grant J4379-N,P29096,W1245.

Notes

1 For ωˆRd1 and ωRd, a mapping γ:ωˆω is bi-Lipschitz if it is bijective and γ as well as its inverse γ1 are Lipschitz continuous.

2 A compact Lipschitz domain is the closure of a bounded Lipschitz domain. For d = 2, it is the finite union of compact intervals with non-empty interior.

3 We use the convention dist(T,):=diam(Γ).

References

  • Feischl M, Führer T, Karkulik M, et al. Quasi-optimal convergence rates for adaptive boundary element methods with data approximation, part I: weakly-singular integral equation. Calcolo. 2014;51(4):531–562.
  • Feischl M, Führer T, Karkulik M, et al. Quasi-optimal convergence rates for adaptive boundary element methods with data approximation, part II: hyper-singular integral equation. Electron Trans Numer Anal. 2015;44:153–176.
  • Feischl M, Karkulik M, Markus Melenk J, et al. Quasi-optimal convergence rate for an adaptive boundary element method. SIAM J Numer Anal. 2013;51(2):1327–1348.
  • Gantumur T. Adaptive boundary element methods with convergence rates. Numer Math. 2013;124(3):471–516.
  • Aurada M, Feischl M, Führer T, et al. Local inverse estimates for non-local boundary integral operators. Math Comp. 2017;86(308):2651–2686.
  • Carstensen C, Feischl M, Page M, et al. Axioms of adaptivity. Comput Math Appl. 2014;67(6):1195–1253.
  • Bespalov A, Betcke T, Haberl A, et al. Adaptive BEM with optimal convergence rates for the Helmholtz equation. Comput Methods Appl Mech Engrg. 2019;346:260–287.
  • Feischl M, Gantner G, Haberl A, et al. Optimal convergence for adaptive IGA boundary element methods for weakly-singular integral equations. Numer Math. 2017;136(1):147–182.
  • Gantner G, Praetorius D, Schimanko S. Adaptive isogeometric boundary element methods with local smoothness control. Math Models Methods Appl Sci. 2020;30:261–307.
  • Gantner G, Haberlik D, Praetorius D. Adaptive IGAFEM with optimal convergence rates: hierarchical B-splines. Math Models Methods Appl Sci. 2017;27(14):2631–2674.
  • Gantner G, Praetorius D. Adaptive IGAFEM with optimal convergence rates: T-splines. Comput Aided Geom Design. 2020;81:101906.
  • Gantner G, Praetorius D. Adaptive BEM for elliptic PDE systems, part II: isogeometric analysis for weakly-singular integral equations. In Preparation. 2020.
  • Faermann B. Localization of the Aronszajn–Slobodeckij norm and application to adaptive boundary elements methods. Part I. The two-dimensional case. IMA J Numer Anal. 2000;20(2):203–234.
  • Faermann B. Localization of the Aronszajn–Slobodeckij norm and application to adaptive boundary element methods. Part II. The three-dimensional case. Numer Math. 2002;92(3):467–499.
  • Carstensen C, Maischak M, Stephan EP. A posteriori error estimate and h-adaptive algorithm on surfaces for Symm's integral equation. Numer Math. 2001;90(2):197–213.
  • Feischl M, Gantner G, Haberl A, et al. Adaptive 2D IGA boundary element methods. Eng Anal Bound Elem. 2016;62:141–153.
  • Gantner G. Optimal adaptivity for splines in finite and boundary element methods. PhD thesis, Institute for Analysis and Scientific Computing, TU Wien, 2017.
  • Mitscha-Eibl G. Adaptive BEM und FEM-BEM-Kopplung für die Lamé-Gleichung. Master's thesis, Institute for Analysis and Scientific Computing, TU Wien, 2014.
  • McLean W. Strongly elliptic systems and boundary integral equations. Cambridge: Cambridge University Press; 2000.
  • Steinbach O. Numerical approximation methods for elliptic boundary value problems. New York: Springer; 2008.
  • Sauter SA, Schwab C. Boundary element methods. Berlin: Springer; 2011.
  • Hofmann S, Mitrea M, Taylor M. Singular integrals and elliptic boundary problems on regular Semmes–Kenig– Toro domains. Int Math Res Not. 2009;2010(14):2567–2865.
  • Di Nezza E, Palatucci G, Valdinoci E. Hitchhiker' s guide to the fractional Sobolev spaces. Bull Sci Math. 2012;136(5):521–573.
  • Feischl M, Führer T, Mitscha-Eibl G, Praetorius D, et al. Convergence of adaptive BEM and adaptive FEM-BEM coupling for estimators without h-weighting factor. Comput Meth Appl Math. 2014;14(4):485–508.
  • Carstensen C. An a posteriori error estimate for a first-kind integral equation. Math Comp. 1997;66(217):139–156.
  • Carstensen C, Stephan EP. Adaptive boundary element methods for some first kind integral equations. SIAM J Numer Anal. 1996;33(6):2166–2183.
  • Bespalov A, Haberl A, Praetorius D. Adaptive FEM with coarse initial mesh guarantees optimal convergence rates for compactly perturbed elliptic problems. Comput Methods Appl Mech Engrg. 2017;317:318–340.
  • Aurada M, Feischl M, Führer T, et al. Efficiency and optimality of some weighted-residual error estimator for adaptive 2D boundary element methods. Comput Methods Appl Math. 2013;13(3):305–332.
  • Stevenson R. The completion of locally refined simplicial partitions created by bisection. Math Comp. 2008;77(261):227–241.
  • Binev P, Dahmen W, DeVore R. Adaptive finite element methods with convergence rates. Numer Math. 2004;97(2):219–268.
  • Stevenson R. Optimality of a standard adaptive finite element method. Found Comput Math. 2007;7(2):245–269.
  • Karkulik M, Pavlicek D, Praetorius D. On 2D newest vertex bisection: optimality of mesh-closure and H1-stability of L2-projection. Constr Approx. 2013;38(2):213–234.
  • Manuel Cascon J, Kreuzer C, Nochetto RH, et al. Quasi-optimal convergence rate for an adaptive finite element method. SIAM J Numer Anal. 2008;46(5):2524–2550.
  • Vuong A-V, Giannelli C, Jüttler B, et al. A hierarchical approach to adaptive local refinement in isogeometric analysis. Comput Methods Appl Mech Engrg. 2011;200(49):3554–3567.
  • Morgenstern P, Peterseim D. Analysis-suitable adaptive T-mesh refinement with linear complexity. Comput Aided Geom Design. 2015;34:50–66.
  • Carstensen C, Faermann B. Mathematical foundation of a posteriori error estimates and adaptive mesh-refining algorithms for boundary integral equations of the first kind. Eng Anal Bound Elem. 2001;25(7):497–509.
  • Feischl M, Gantner G, Praetorius D. Reliable and efficient a posteriori error estimation for adaptive IGA boundary element methods for weakly-singular integral equations. Comput Methods Appl Mech Engrg. 2015;290:362–386.
  • Hackbusch W. Integral equations: theory and numerical treatment. Basel: Birkhäuser; 1995.
  • Ferraz-Leite S, Praetorius Dirk. Simple a posteriori error estimators for the h-version of the boundary element method. Computing. 2008;83(4):135–162.
  • Carstensen C, Praetorius D. Averaging techniques for the effective numerical solution of symm's integral equation of the first kind. SIAM J Sci Comput. 2006;27(4):1226–1260.
  • Morrey Jr CB. Multiple integrals in the calculus of variations. Springer, Berlin, reprint of the 1966 original edition, 2008.
  • Feischl M, Führer T, Praetorius D. Adaptive FEM with optimal convergence rates for a certain class of nonsymmetric and possibly nonlinear problems. SIAM J Numer Anal. 2014;52(2):601–625.

Appendices

Appendix 1. Smooth characteristic functions

The following lemma provides a smooth cut-off function as used in the proof of discrete reliability (E4); see Section 4.8. For regular simplicial meshes as in Section 3.7, such functions can easily be constructed by means of hat functions; see [Citation5, Section 5.3]. For the present abstract setting, the construction is more technical.

Lemma A.1

Let TT and ST. Suppose (M1)–(M4). Then there exists a function χ~SH1(Γ) such that for almost all xΓ (A.1a) χ~S(x)=1if xS,(A.1a) (A.1b) 0χ~S(x)1if xπ(S),(A.1b) (A.1c) χ~S(x)=0if xπ(S).(A.1c) Further, there exists a constant C>0 such that (A2) |Γχ~S(x)|Ch(x)1for almost all xΓ.(A2) The constant C depends only on the dimension d and the constants from (M1)–(M4).

Proof.

In the following three steps, we will even prove the existence of a function χ~SC(O) with an open superset OΓ such that the restriction to Γ has the desired properties. With the constants from (M1)–(M2) and (M4), we define for all TT, (A3) δ1(T):=diam(T)2CpatchClocuniCcent,δ2(T):=diam(T)Ccent,δ3(T):=diam(T)2Ccent.(A3) Step 1: First, we construct an equivalent smooth mesh-size function δC(Rd) with uniformly bounded gradient on Γ. Let K1C(Rd) be a standard mollifier with 0K11 on B1(0), K1=0 on RdB1(0), and RdK1dx=1. For s>0, we set Ks():=K1(/s)sd. By convolution, we define (A4) δ:=TTδ1(T)χBδ2(T)(T)Kδ2(T).(A4) Note that supp(χBδ2(T)(T)Kδ2(T))B2δ2(T)(T) for TT. Thus (M4) and the choice (EquationA3) of δ2(T) yield that supp(χBδ2(T)(T)Kδ2(T))Γπ(T). Together with (M1)–(M2) and 0χBδ2(T)(T)Kδ2(T)1, this implies for the interior int(T) of any TT that δ|int(T)TTδ1(T)χπ(T)|int(T)=TΠ(T)δ1(T)CpatchClocuniδ1(T). Note that the restriction to the interior is indeed necessary since the second term is discontinuous across faces of T. However, by continuity of δ, this estimate is also satisfied if int(T) is replaced by T, i.e. δ|TCpatchClocuniδ1(T). The fact that χBδ2(T)(T)Kδ2(T)=1 on T shows that also the converse estimate is valid. This leads to (A5) diam(T)2CpatchClocuniCcent=δ1(T)δ|TCpatchClocuniδ1(T)=δ3(T)for all TT.(A5) In particular, this proves the existence of an open set RdOΓ such that δ>0 on O. Finally, we consider the gradient of δ for xΓ. Recall that supp(χBδ2(T)(T)Kδ2(T))π(T). Together with the Hölder inequality, KsL1(Rd)s1, and (M1)–(M2), this proves that (A6) |δ(x)|=TTδ1(T)χπ(T)(x)|χBδ2(T)(T)Kδ2(T)(x)|TTδ1(T)χπ(T)(x)δ2(T)11.(A6)

Step 2: In this step, we construct χ~S and prove (EquationA.1a)–(EquationA.1c). For xO, we define the quasi-convolution χ~S(x):=RdχS~(y)Kδ(x)(xy)dy,where S~:={Bδ3(T)(T):TS}. Since δ>0 on O, the chain rule shows that any derivative with respect to x of the term Kδ(x)(xy) exists and is continuous at all (x,y)O×S~. This yields that χ~SC(O) and χ~S|ΓH1(Γ); see, e.g. [Citation19, page 98–99]. Since supp(Ks)=Bs(0), it holds that (A7) χ~S(x)=Bδ(x)(x)χS~(y)Kδ(x)(xy)dy.(A7) If Bδ(x)(x)S~ and thus χS~(y)=1, the properties of K1 show that χ~S(x)=1. Due to δ|Tδ3(T) for all TT (which follows from (EquationA5)), this is particularly satisfied if xS. This proves (EquationA.1a). Moreover, (EquationA7) shows that 0χ~S(x)1 for all xRd (and hence verifies (EquationA.1b)), and χ~S(x)=0 if Bδ(x)(x)S~=. For (EquationA.1c), it thus remains to prove that xΓπ(S) implies that Bδ(x)(x)S~=. We prove the contraposition. Let xΓ and suppose that Bδ(x)(x)S~. Then, there exists TS and yRd such that |xy|<δ(x) and dist({y},T)<δ3(T). The triangle inequality yields that (A8) dist({x},T)|xy|+dist({y},T)<δ(x)+δ3(T)2max{δ(x),δ3(T)}.(A8) Now, we differ two different cases. If δ(x)δ3(T), then we have that dist({x},T)<2δ3(T). The choice (EquationA3) of δ3(T) together with (M4) shows that xπ(T)π(S). If δ(x)>δ3(T), then we have that dist({x},T)<2δ(x). Let TT with xT and zT with |xz|=dist({x},T). Together with (EquationA5) and (EquationA8), this yields that dist({z},T)|xz|=dist({x},T)<2δ(x)2CpatchClocuniδ1(T)=diam(T)Ccent. Hence, (M4) implies that zΓπ(T) and thus zπ(T). Any zΓ that is sufficiently close to z also satisfies that dist({z},T)<diam(T)/Ccent and thus zπ(T). Since zT=int(T)¯, z can be particularly chosen in int(T). Hence, we see that int(T)π(T). This is equivalent to TΠ(T), which concludes that xTπ(T)π(S).

Step 3: Finally, we prove (EquationA2). We recall that δ>0 on O; see Step 1. With the identity matrix IRd×d and the matrix (xy)(δ(x))Rd×d, elementary calculations prove for all xOΓ and all yRd that xKδ(x)(xy)=K1xyδ(x)δ(x)I(xy)(δ(x))δ(x)2δ(x)d+K1xyδ(x)δ(x)d1(d)(δ(x)). Considering the norm, we see that |xKδ(x)(xy)|δ(x)d1+|xy||δ(x)|δ(x)d2+δ(x)d1|δ(x)|. Together with supp(Ks)=Bs(0), this shows for all xΓ that |χ~S(x)|=|RdχS~(y)xKδ(x)(xy)dy|Bδ(x)(x)δ(x)d1+|xy||δ(x)|δ(x)d2+δ(x)d1|δ(x)|dyδ(x)1(1+δL(Γ)). Thus  (EquationA5)–(EquationA6) and (M3) prove that |χ~S(x)|h(x)1 for almost all xΓ. Moreover, for smooth functions, the surface gradient Γ is the orthogonal projection of the gradient ∇ onto the tangent plane; see, e.g. [Citation18, Lemma 2.22]). With the outer normal vector ν, this implies that Γχ~S=χ~S(χ~Sν)ν almost everywhere on Γ and concludes the proof with the previous estimate.

Appendix 2

Inverse inequalities for other integral operators

In Proposition 4.13, we have generalized an inverse estimate from [Citation5] for the single-layer operator V:H1/2(Γ)DH1/2(Γ)D. [Citation5] additionally derived similar estimates for the double-layer operator K:H1/2(Γ)DH1/2(Γ)D, the adjoint double-layer operator K:H1/2(Γ)DH1/2(Γ)D, and the hyper-singular operator W:H1/2(Γ)DH1/2(Γ)D; see, e.g. [Citation19, page 218] for a precise definition (where these operators are denoted by 12T,12T~, and R, respectively). Although [Citation5] considered only the Laplace problem, the techniques of the proof of Proposition 4.13 extend the result at least to partial differential operators without lowest-order term cu. With some further effort, one can even prove it for arbitrary PDE operators P with constant coefficients as in Section 2.3, where, as in Section 4.4, ellipticity of P can be replaced by ellipticity up to some compact perturbation. To this end, one requires additional regularity of the trace operator ()|Γ:H3/2(Ω)DH1(Γ)D, which is satisfied for piecewise smooth boundaries Γ; see, e.g. [Citation21, Remark 3.1.18].

For the proof, we will frequently use [Citation19, Theorem 4.24], which reads as follows: Let uH1(Ω) be arbitrary with PuL2(Ω)D in the weak sense and u|ΓH1(Γ). Then, DνuL2(Γ)D and (B.1) DνuL2(Γ)u|ΓH1(Γ)+uH1(Ω)+PuL2(Ω).(B.1)

Proposition B.1

Suppose (M1)–(M5). For TT, let wL(Γ) be a weight function which satisfies for some α>0 and all TT that (B.2) wL(T)αw(x)for almost all xπ(T).(B.2) Then, there exists a constant C1>0 such that for all ψL2(Γ)D and vH1(Γ)D, (B.3) wΓVψL2(Γ)+wKψL2(Γ)C1w/h1/2L(Γ)ψH1/2(Γ)+wψL2(Γ),(B.3) If we additionally suppose that the trace operator satisfies the stability ()|Γ:H3/2(Ω)DH1(Γ)D, there exists a constant C2>0 such that for all ψL2(Γ)D and vH1(Γ)D, (B.4) wΓKvL2(Γ)+wWvL2(Γ)C2w/h1/2L(Γ)vH1/2(Γ)+wΓvL2(Γ).(B.4) The constants C1 and C2 depend only on (M1)–(M5), Γ, the coefficients of P, and the admissibility constant α.

Proof of (EquationB3).

Proof of (EquationB.3)

The bound for wΓVψL2(Γ) is just the assertion of Proposition 4.13. With the results from the proof of Proposition 4.13, one can easily estimate the second summand wKψL2(Γ) as in [Citation5, Section 6]. In particular, the stability of K:L2(Γ)L2(Γ) is exploited, which is stated in [Citation19, page 209]. Further, one uses the fact from [Citation19, page 218] that K=D~νintV~1/2, where D~νint() denotes the interior modified conormal derivative from [Citation19, page 117–118].

Proof of (EquationB4).

Proof of (EquationB.4)

As in the proof of Proposition 4.13, we abbreviate δ1(T):=diam(T)/(2Ccent) and UT:=Bδ1(T)(T) for all TT. Let vT:=|T|1Tvdx. Further, let χ~T:=χ~{T} be the smooth quasi-indicator function of T from Lemma A.1. With the Poincaré inequality (Equation56), the localization properties (Equation55) as well as (Equation60), and (M5), one can easily verify that (B.5) hT1vvTL2(π(T))+hT1/2(vvT)χ~TH1/2(Γ)+(vvT)χ~TH1(Γ)ΓvL2(π(T)).(B.5) We fix (independently of T) a bounded domain URd with UTU for all TT. Let K~:H1/2(Γ)DH1(UΓ)D denote the double-layer potential from [Citation19, Theorem 6.11]. With these preparations and if P has no lower-order terms, i.e. bi=0 for all i{1,,d} as well as c = 0, the proof of (EquationA4) follows as in [Citation5, Section 5 and Section 6]. In particular, one exploits the fact that (B.6) K~vT=vTonΩandK~vT=0onΩext:=RdΩ¯,(B.6) which follows from the representation formula [Citation19, Theorem 6.10] together with the assumption that P has no lower-order terms. The proof of (EquationB.4) then employs the Poincaré inequality (EquationB.5), the property (EquationB.6), the stability of K:H1(Γ)DH1(Γ)D and of W:H1(Γ)DL2(Γ)D from [Citation22, Corollary 3.38] or [Citation19, page 209], and the Caccioppoli inequality (Equation71) in combination with the transmission property [Citation19, Theorem 6.11].

To prove (EquationA12) for general P with lower-order terms, where (EquationB.6) is in general false, we recall the principal part P0 from (Equation91). In the following five steps, we show for the corresponding double-layer operator K0:H1/2(Γ)H1/2(Γ) and the hyper-singular operator W0:H1/2(Γ)H1/2(Γ) that (B.7) KK0:H1/2(Γ)H1(Γ)andWW0:H1/2(Γ)L2(Γ)(B.7) are continuous. Since (EquationB.4) is satisfied for the operators corresponding to P0, this and the trivial estimate ww/h1/2L(Γ) will conclude the proof.

Step 1: Let N~,N~0 be the Newton potentials from [Citation19, Theorem 6.1] corresponding to P,P0. According to [Citation19, Theorem 6.1], they satisfy the mapping property N~,N~0:Hσ(Rd)DHσ+2(Rd)D for all σR. In the proof of the latter stability, the fundamental solution is defined in terms of the Fourier transformation. The definition involves a multivariate polynomial P:RdCD×D resp. P0:RdCD×D associated to P resp. P0 (which is obtained from the differential operator by replacing the derivatives with variables) such that |P(ξ)||P0(ξ)1|=O(|ξ|2) for ξRd and |ξ|; see [Citation19, Equation (6.7)]. Indeed, the latter inequality is the key of the proof. As elementary analysis even shows that |P(ξ)1P0(ξ)1|=|P(ξ)1[IP(ξ)P0(ξ)1]|=O(|ξ|3) with the identity matrix ICD×D, one sees along the lines of [Citation19, Theorem 6.1] the additional regularity (B.8) N~N~0:Hσ(Rd)DHσ+3(Rd)D.(B.8) Since multiplication with a fixed compactly supported smooth function is stable (see, e.g. [Citation19, Theorem 3.20]), we also see for arbitrary χ1,χ2Cc(Rd) the continuity of (B.9) A:Hσ(Rd)DHσ+3(Rd)D,gχ1(N~N~0)(gχ2).(B.9) Step 2: For sufficiently large λ>0, the sesquilinear forms (;)P+λ and (;)P0+λ from (Equation18) are both even elliptic on H01(Ω)D. Let uλH1(Ω) be the unique weak solution of (P+λ)uλ=0 and uλ|Γ=v. Similarly, let u0,λH1(Ω) be the solution of (P0+λ)u0,λ=0 and u0,λ|Γ=v. We extend both functions by zero outside of Ω. Since Puλ=uλ and P0u0,λ=u0,λ, the representation formula [Citation19, Theorem 6.10] yields that (B.10) (K~K~0)v=(uλu0,λ)λ(N~uλN~0u0,λ)+(V~DνuλV~0Dνu0,λ).(B.10) We note that KK0=(K~K~0)()|Γ and WW0=Dν(K~K~0) with the conormal derivative Dν(). To see (EquationB.7) and hence to conclude the proof, we thus only have to bound the trace as well as the conormal derivative of each summand in (EquationB.10) separately, which will be done in the following three steps.

Step 3: By definition, the trace of the first summand in (EquationB.10) vanishes, i.e. (uλu0,λ)|Γ=0. According to (EquationA9) (where the differential operator can also be chosen as P+λ instead of P), the normal derivative satisfies that Dν(uλu0,λ)L2(Γ)(A9)(uλu0,λ)|ΓH1(Γ)+uλu0,λH1(Ω)+(P+λ)(uλu0,λ)L2(Ω). Since the first summand vanishes and (P+λ)(uλu0,λ)L2(Ω)u0,λH1(Ω) due to (P+λ)(uλu0,λ)=(i=1dbiiu0,λ)cu0,λ, the stability of the solution mapping vuλ and vu0,λ gives that (B.11) Dν(uλu0,λ)L2(Γ)vH1/2(Γ).(B.11) Step 4: Due to our assumption that the trace operator ()|Γ:H3/2(Ω)DH1(Γ)D is well-defined and continuous, the stability of the Newton potentials, the stability of the solution mapping vuλ and vu0,λ, the trace of the second summand in (EquationB.10) satisfies that (N~uλN~0u0,λ)|ΓH1(Γ)(N~uλ)|ΓH1(Γ)+(N~0u0,λ)|ΓH1(Γ)N~uλH3/2(Ω)+N~0u0,λH3/2(Ω)uλH1/2(Rd)+u0,λH1/2(Rd)uλL2(Rd)+u0,λL2(Rd)=uλL2(Ω)+u0,λL2(Ω)vH1/2(Γ). Note that N~ and N~0 are indeed potentials, i.e. PN~=0 weakly and P0N~0=0 weakly. With (EquationB.1) (applied for P and P0), the stability of the Newton potentials, and the estimates for the trace of N~uλ and N~0u0,λ, we thus see that Dν(N~uλN~0u0,λ)L2(Γ)(A9)(N~uλ)|ΓH1(Γ)+N~uλH1(Ω)+(N~0u0,λ)|ΓH1(Γ)+N~0u0,λH1(Ω)vH1/2(Γ)+uλH1(Rd)+u0,λH1(Rd)vH1/2(Γ)+uλL2(Rd)+u0,λL2(Rd). Again, the stability of the solution mappings allows to bound the last terms by vH1/2(Γ).

Step 5: To deal with the third summand in (EquationB.10), we first rewrite it as follows: (B.12) V~DνuλV~0Dνu0,λ=V~Dν(uλu0,λ)+(V~V~0)Dνu0,λ.(B.12) Due to the stability of V~()|Γ=V:L2(Γ)H1(Γ) (see (Equation21) and (Equation76)) and (EquationB.11), we have for the first summand in (EquationB.12) that (B.13) (V~Dν(uλu0,λ))|ΓH1(Γ)vH1/2(Γ).(B.13) To deal with the conormal derivative, we apply again (EquationB.1) together with the fact that V~ is a potential, i.e. PV~=0 weakly. This leads to DνV~Dν(uλu0,λ)L2(Γ)(A9)(V~Dν(uλu0,λ))|ΓH1(Γ)+V~Dν(uλu0,λ)H1(Ω). The first summand can be bounded as in (EquationB.13). For the second summand, we use the stability (Equation75) in combination with H1/2(Γ)L2(Γ) and (EquationB.11).

Finally, it only remains to bound the trace as well as the conormal derivative of the second summand in (EquationB.12). Choosing σ=1 in the additional regularity (EquationB.9), one sees as in the proof of [Citation19, page 203] (which proves stability of V~:H1/2(Γ)DH1(Ω)D) that (B.14) V~V~0:H1/2(Γ)DH2(Ω)D(B.14) With the assumption ()|Γ:H3/2(Ω)DH1(Γ)D, this implies that (B.15) ((V~V~0)Dνu0,λ)|ΓH1(Γ)(V~V~0)Dνu0,λH3/2(Ω)Dνu0,λH1/2(Γ).(B.15) Recall that (P0+λ)u0,λ=0 with u0,λ|Γ=v. Thus [Citation19, Theorem 4.25] (which states boundedness of the Dirichlet to Neumann mapping) gives that (B.16) ((V~V~0)Dνu0,λ)|ΓH1(Γ)(A23)Dνu0,λH1/2(Γ)vH1/2(Γ).(B.16) Moreover, (EquationA9), the fact that P(V~V~0)=(i=1dbiiV~0)cV~0, the stability (Equation75), and (EquationB.16) show for the conormal derivative that Dν(V~V~0)Dνu0,λL2(Γ)((V~V~0)Dνu0,λ)|ΓH1(Γ)+(V~V~0)Dνu0,λH1(Ω)+P(V~V~0)Dνu0,λL2(Ω)vH1/2(Γ). Overall, we have estimated the trace as well as the conormal derivative of all terms in (EquationB.10). Since KK0=(K~K~0)()|Γ and WW0=Dν(K~K~0), this verifies the stability (EquationB.7) and thus concludes the proof.