Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Skip to content
Publicly Available Published by De Gruyter June 19, 2018

A Hybrid High-Order Method for Highly Oscillatory Elliptic Problems

  • Matteo Cicuttin , Alexandre Ern and Simon Lemaire EMAIL logo

Abstract

We devise a Hybrid High-Order (HHO) method for highly oscillatory elliptic problems that is capable of handling general meshes. The method hinges on discrete unknowns that are polynomials attached to the faces and cells of a coarse mesh; those attached to the cells can be eliminated locally using static condensation. The main building ingredient is a reconstruction operator, local to each coarse cell, that maps onto a fine-scale space spanned by oscillatory basis functions. The present HHO method generalizes the ideas of some existing multiscale approaches, while providing the first complete analysis on general meshes. It also improves on those methods, taking advantage of the flexibility granted by the HHO framework. The method handles arbitrary orders of approximation k0. For face unknowns that are polynomials of degree k, we devise two versions of the method, depending on the polynomial degree (k-1) or k of the cell unknowns. We prove, in the case of periodic coefficients, an energy-error estimate of the form (ε12+Hk+1+(εH)12), and we illustrate our theoretical findings on some test-cases.

MSC 2010: 65N30; 65N08; 76R50

1 Introduction

Over the last few years, many advances have been accomplished in the design of arbitrary-order polytopal discretization methods. Such methods are capable of handling meshes with polytopal cells, and possibly including hanging nodes. The use of polytopal meshes can be motivated by the increased flexibility, when meshing complex geometries, or when using agglomeration techniques for mesh coarsening (see, e.g., [7]). Classical examples of polytopal methods are the (polytopal) Finite Element Method (FEM) [46, 44], which typically uses non-polynomial basis functions to enforce continuity, and non-conforming methods such as the Discontinuous Galerkin (DG) [5, 16, 10] and the Hybridizable Discontinuous Galerkin (HDG) [15] methods. We also mention the Weak Galerkin (WG) [47] method (see [13] for its links to HDG).

More recently, new paradigms have emerged. One salient example is the Virtual Element Method (VEM) [9], which is formulated in terms of virtual (i.e., non-computed) conforming functions. The key idea is that the virtual space contains those polynomial functions leading to optimal approximation properties, whereas the remaining functions need not be computed (only their degrees of freedom need to be) provided some suitable local stabilization is introduced. The degrees of freedom in the VEM are attached to the mesh vertices, and, as the order of the approximation is increased, also to the mesh edges, faces, and cells. Another recent polytopal method is the Hybrid High-Order (HHO) method, which has been introduced for locking-free linear elasticity in [17], and for diffusion in [19]. The HHO method has been formulated originally as a non-conforming method, using polynomial unknowns attached to the mesh faces and cells. The HHO method has been bridged in [14] both to HDG (by identifying a suitable numerical flux trace), and to the non-conforming VEM considered in [6] (by identifying an isomorphism between the HHO degrees of freedom and a local virtual finite-dimensional space, which again contains those polynomial functions leading to optimal approximation properties). The focus here is on HHO methods. HHO methods offer several assets, including a dimension-independent construction, local conservativity, and attractive computational costs, especially in 3D. Indeed, the HHO stencil is more compact than for methods involving degrees of freedom attached to the mesh vertices, and static condensation allows one to eliminate cell degrees of freedom, leading to a global problem expressed in terms of face degrees of freedom only, whose number grows quadratically with the polynomial order, whereas the growth of globally coupled degrees of freedom is typically cubic for DG methods.

In this work, we are interested in elliptic problems featuring heterogeneous/anisotropic coefficients that are highly oscillatory. The case of slowly varying coefficients has already been treated in [18, 20], where error estimates tracking the dependency of the approximation with respect to the local heterogeneity/anisotropy ratios have been derived. Let Ω be an open, bounded, connected polytopal subset of d, d{2,3}, and ε>0, supposedly much smaller than the diameter of the domain Ω, encode the highly oscillatory nature of the coefficients. We consider the model problem

(1.1){-div(𝔸εuε)=fin Ω,uε=0on Ω,

where fL2(Ω) is non-oscillatory, and 𝔸ε is an oscillatory, uniformly elliptic and bounded matrix-valued field on Ω. It is well known that the Hk+2-norm of the solution uε to problem (1.1) scales as ε-(k+1), meaning that monoscale methods (including the monoscale HHO method of order k0 of [18, 20]) provide an energy-norm decay of the error of order (hε)k+1. To be accurate, such methods must hence rely on a mesh resolving the fine scale, i.e., with size hε. Since ε is supposedly much smaller than the diameter of Ω, an accurate approximation necessarily implies an overwhelming number of degrees of freedom. In a multi-query context, where the solution is needed for a large number of right-hand sides (e.g., an optimization loop, with f as a control and (1.1) as a distributed constraint), a monoscale solve is hence unaffordable. In that context, multiscale methods may be preferred. Multiscale methods aim at resolving the fine scale in an offline step, reducing the online step to the solution of a system of small size, based on oscillatory basis functions computed in the offline step, on a coarse mesh with size Hε. In a single-query context, multiscale methods are also interesting since they allow one to organize computations in a more efficient way.

Multiscale approximation methods on classical element shapes (such as simplices or quadrangles/hexahedra) have been analyzed extensively in the literature. Examples include, e.g., the multiscale Finite Element Method (msFEM) [34, 35, 23] (with energy-error bound of the form (ε12+H+(εH)12) in the periodic case), its variant using oversampling [34, 24] (with improved error bound of the form (ε12+H+εH) in the periodic case), or the Petrov–Galerkin variant of the msFEM using oversampling [36]. Let us also mention [3] (see also [33]), which is an extension to arbitrary orders of approximation of the classical msFEM (with error bound of the form (ε12+Hk+(εH)12) in the periodic case using H1-conforming finite elements of degree k1). These methods all rely on the assumption that a conforming finite element basis is available for the (coarse) mesh under consideration. Recent research directions essentially focus on the approximation of problems that do not assume scale separation, and on reducing and possibly eliminating the cell resonance error. One can cite, e.g., the Generalized msFEM (GmsFEM) [22], or the Local Orthogonal Decomposition (LOD) approach [32, 41]. We also mention that other paradigms exist to approximate oscillatory problems, like the Heterogeneous Multiscale Method (HMM) [21, 1].

On general polytopal meshes, the literature on multiscale methods is more scarce. For constructions in the spirit of the msFEM, one can cite the msFEM à la Crouzeix–Raviart of [39, 40], the so-called Multiscale Hybrid-Mixed (MHM) [4, 43] approach, and the (polynomial-based) method of [26] in the HDG context. Each one of these methods has its proper design, but they all share the same construction principles: they are based, more or less directly, on oscillatory basis functions that solve local Neumann problems with polynomial boundary data, and result in global systems (posed on the coarse mesh) that can be expressed in terms of face unknowns only. In the following, we will thus refer to those methods as skeletal-msFEM. The MHM approach actually presents a small difference with respect to the two other approaches since it is based on a hybridized primal formulation, which leads to consider flux-type unknowns at interfaces instead of potential-type unknowns; as a consequence, and in order to impose the compatibility constraint, one needs to solve a saddle-point global problem, whereas for the two other approaches, one ends up with a coercive problem. For the msFEM à la Crouzeix–Raviart, an error bound of the form (ε12+H+(εH)12) is proved in [39] in the periodic case. However, the analysis is led under the assumption that there exists a finite number of reference elements in the mesh sequence. For the MHM approach, which is designed in the same spirit, the same type of upper bound for the error is expected. Yet, in [43], the authors claim that their method is able to get rid of the resonance error (without oversampling); we clarify this issue in Remark A.3 below. For the HDG-like method, the analysis that is provided in [26] is sharp only in the regime Hε. As a consequence, there is, to date, no complete polytopal analysis available in the literature for skeletal-msFEM. Moreover, we observe that in the three methods, the discretization of the (non-oscillatory) right-hand side is realized in a somewhat suboptimal way, which can become a limiting issue in a multi-query context. In the msFEM à la Crouzeix–Raviart, the discretization is realized through a projection of the loading term onto the space spanned by the oscillatory basis functions. In the MHM and HDG-like approaches, the whole (local) space H1 is considered. In all cases, the approximation of the right-hand side does not take advantage of the fact that the latter is non-oscillatory. Let us mention, as another construction in the spirit of the msFEM, the work [38], which exploits in the DG context the ideas introduced in [3]. The drawback, which is inherent to DG methods, is the large size of the online systems. For constructions in the spirit of the GmsFEM, let us mention in the HDG context the contributions [25, 11] (that are based on [26]), and the work [42] in the WG context.

In this work, we devise a multiscale HHO (msHHO) method, which can be seen as a generalization (in particular to arbitrary orders of approximation) of the msFEM à la Crouzeix–Raviart of [39, 40]. Our contribution is twofold. First, we provide an analysis (in the periodic setting) of the method that is valid on general polytopal mesh sequences (in particular, we do not postulate the existence of reference elements); in that respect, this work presents the first complete polytopal analysis of a skeletal-msFEM. Note that considering general element shapes in the periodic setting is clearly not a good strategy (cf., e.g., [31]); however, this setting is not our final target. Second, taking advantage of the flexibility offered by the HHO framework, we improve on the existing methods. We introduce (polynomial) cell unknowns, that we use for the integration of the right-hand side (cf. Remarks 5.8 and 5.16 below). The non-oscillatory loading is hence discretized through a coarse-scale polynomial projection, while the size of the online system remains unchanged since the cell unknowns are locally eliminated in the offline step. Two versions of the msHHO method are proposed herein, both employing polynomials of arbitrary order k0 for the face unknowns. For the mixed-order msHHO method, the cell unknowns are polynomials of order (k-1) (if k1), whereas they are polynomials of order k0 for the equal-order msHHO method. The mixed-order msHHO method does not require stabilization, whereas a simple stabilization (which avoids computing additional oscillatory basis functions) is introduced in the equal-order case. We prove for both methods an energy-error estimate of the form (ε12+Hk+1+(εH)12)=:gk(H) in the periodic case. The analysis of the msHHO method differs from that of the monoscale HHO method since the local fine-scale space does not contain polynomial functions up to order (k+1); in this respect, our key approximation result is Lemma 4.5 below. With respect to [39], we also simplify the analysis and weaken the regularity assumptions (cf. Remark 4.6 below). Our analysis finally sheds new light on the relationship between the non-computed functions of the local virtual space and the associated local stabilization. To motivate the design and use of a high-order method, we note, as it was already pointed out in [3], that the upper bound gk(H) is minimal for Hk=(ε12/2(k+1))2/(2k+3), hence as k0 increases, Hk increases whereas gk(Hk) decreases. The msHHO method we devise is meant to be a first step in the design of an accurate and computationally effective multiscale approach on general meshes. The next step will be to address the resonance phenomenon and the more realistic setting of no scale separation.

The article is organized as follows. In Sections 2 and 3 we introduce, respectively, the continuous and discrete settings. In Section 4, we introduce the fine-scale approximation space, exhibiting its (oscillatory) basis functions and studying, locally, its approximation properties. In Section 5, we introduce the two versions of the msHHO method, analyze their stability, and derive energy-error estimates. In Section 6, we present some numerical illustrations in the periodic and locally periodic settings. Finally, in Appendix A we collect some useful estimates on the first-order two-scale expansion.

2 Continuous Setting

From now on, and in order to lead the analysis, we assume that the diffusion matrix 𝔸ε satisfies 𝔸ε()=𝔸(/ε) in Ω, where 𝔸 is a symmetric and d-periodic matrix field on d. Letting Q:=(0,1)d, we define, for 1p+ and m, the following periodic spaces:

Lperp(Q):={vLlocp(d):v is d-periodic},
Wperm,p(Q):={vWlocm,p(d):v is d-periodic},

with the classical conventions that Wperm,2(Q) is denoted Hperm(Q) and that the subscript “loc” can be omitted for p=+. Letting 𝒮d() denote the set of real-valued d×d symmetric matrices, we also define, for real numbers 0<ab,

𝒮ab:={𝕄𝒮d():a|𝝃|2𝕄𝝃𝝃b|𝝃|2 for all 𝝃d}.

We assume that there exist real numbers 0<αβ such that

(2.1)𝔸()𝒮αβa.e. in d.

Assumption (2.1) ensures that 𝔸εL(Ω;d×d) is such that 𝔸ε()𝒮αβ a.e. in Ω for any ε>0, and hence guarantees the existence and uniqueness of the solution to (1.1) in H01(Ω) for any ε>0. More importantly, assumption (2.1) ensures that the (whole) family (𝔸ε)ε>0 G-converges [2, Section 1.3.2] to some constant symmetric matrix 𝔸0𝒮αβ. Henceforth, we denote ρ:=βα1 the (global) heterogeneity/anisotropy ratio of both (𝔸ε)ε>0 and 𝔸0. Letting (𝒆1,,𝒆d) denote the canonical basis of d, the expression of 𝔸0 is known to read, for integers 1i,jd,

(2.2)[𝔸0]ij=Q𝔸(𝒆j+μj)(𝒆i+μi)=Q𝔸(𝒆j+μj)𝒆i,

where, for any integer 1ld, the so-called corrector μlHper1(Q) is the solution with zero mean-value on Q to the problem

(2.3){-div(𝔸(μl+𝒆l))=0in d,μl is d-periodic.

For further use, we also define the linear operator ε:Lperp(Q)Lp(Ω), 1p+, such that, for any function χLperp(Q), ε(χ)Lp(Ω) satisfies ε(χ)()=χ(/ε) in Ω. In particular, for any integers 1i,jd, we have [𝔸ε]ij=ε(𝔸ij). A useful property of ε is the relation l(ε(χ))=1εε(lχ), valid for any function χWper1,p(Q) and any integer 1ld.

The homogenized problem reads

(2.4){-div(𝔸0u0)=fin Ω,u0=0on Ω.

We introduce the so-called first-order two-scale expansion

(2.5)ε1(u0):=u0+εl=1dε(μl)lu0.

Note that (uε-ε1(u0)) does not a priori vanish on the boundary of Ω.

3 Discrete Setting

We denote by + a countable set of meshsizes having 0 as its unique accumulation point, and we consider mesh sequences of the form (𝒯H)H. For any meshsize H, a mesh𝒯H is a finite collection of non-empty disjoint open polytopes (polygons/polyhedra) T, called elements or cells, such that Ω¯=T𝒯HT¯ and H=maxT𝒯HHT, HT standing for the diameter of the cell T. The mesh cells being polytopal, their boundary is composed of a finite union of portions of affine hyperplanes in d called facets (each facet has positive (d-1)-dimensional measure). A closed subset F of Ω¯ is called a face if either

  1. there exist T1,T2𝒯H such that F=T1T2Z, where Z is an affine hyperplane supporting a facet of both T1 and T2 (and F is termed interface), or

  2. there exists T𝒯H such that F=TΩZ, where Z is an affine hyperplane supporting a facet of both T and Ω (and F is termed boundary face).

Interfaces are collected in the set Hi, boundary faces in Hb, and we let H:=HiHb. The diameter of a face FH is denoted HF. For all T𝒯H, we define

T:={FH:FT}

the set of faces lying on the boundary of T; note that the faces in T compose the boundary of T. For any T𝒯H, we denote by 𝒏T the unit normal vector to T pointing outward T, and for any FT, we let 𝒏T,F:=𝒏TF (by definition, 𝒏T,F is a constant vector on F).

We adopt the following notion of admissible mesh sequence; cf. [16, Section 1.4] and [20, Definition 2.1].

Definition 3.1 (Admissible Mesh Sequence).

The mesh sequence (𝒯H)H is admissible if, for all H, 𝒯H admits a matching simplicial sub-mesh 𝔗H (meaning that the cells in 𝔗H are sub-cells of the cells in 𝒯H and that the faces of these sub-cells belonging to the skeleton of 𝒯H are sub-faces of the faces in H), and there exists a real number γ>0, called mesh regularity parameter, such that, for all H, the following holds:

  1. For all simplex S𝔗H of diameter HS and inradius RS, γHSRS.

  2. For all T𝒯H, and all S𝔗T:={S𝔗H:ST}, γHTHS.

Two classical consequences of Definition 3.1 are that, for any mesh 𝒯H belonging to an admissible mesh sequence,

  1. the quantity card(T) is bounded independently of the diameter HT for all T𝒯H (see [16, Lemma 1.41]),

  2. mesh faces have a comparable diameter to the diameter of the cells to which they belong (see [16, Lemma 1.42]).

For any q, and any integer 1ld, we denote by lq the linear space spanned by l-variate polynomial functions of total degree less than or equal to q. We let

Nlq:=dim(lq)=(q+lq).

Let a mesh 𝒯H be given. For any T𝒯H, dq(T) is composed of the restriction to T of polynomials in dq, and for any FH, d-1q(F) is composed of the restriction to F of polynomials in dq (this space can also be described as the restriction to F of polynomials in d-1qΘ-1, where Θ is any affine bijective mapping from d-1 to the affine hyperplane supporting F). We also introduce, for any T𝒯H, the following broken polynomial space:

d-1q(T):={vL2(T):vFd-1q(F) for all FT}.

The term “broken” refers to the fact that no continuity is required between adjacent faces for functions in d-1q(T). For any T𝒯H, we denote by (ΦTq,i)1iNdq a set of basis functions of the space dq(T), and for any FH, we denote by (ΦFq,j)1jNd-1q a set of basis functions of the space d-1q(F). We define, for any T𝒯H and FH, ΠTq and ΠFq as the L2-orthogonal projectors onto the spaces dq(T) and d-1q(F), respectively. Whenever no confusion can arise, we write, for all T𝒯H, all FT, and all vH1(T), ΠFq(v) instead of ΠFq(vF).

We conclude this section by recalling some classical results, that are valid for any mesh 𝒯H belonging to an admissible mesh sequence in the sense of Definition 3.1. For any T𝒯H and FT, the trace inequalities

(3.1)vL2(F)ctr,dHF-12vL2(T)for all vdq(T),
(3.2)vL2(F)ctr,c(HT-1vL2(T)2+HTvL2(T)d2)12for all vH1(T),

hold [16, Lemmas 1.46 and 1.49], as well as the local Poincaré inequality

(3.3)vL2(T)cPHTvL2(T)dfor all vH1(T) such that Tv=0,

where cP=π-1 for convex elements [8]; estimates in the non-convex case can be found, e.g., in [45]. Finally, proceeding as in [27, Lemma 5.6], one can prove using the above trace and Poincaré inequalities that

(3.4)|v-ΠTq(v)|Hm(T)+HT12|v-ΠTq(v)|Hm(F)cappHTs-m|v|Hs(T)vHs(T),

for integers 1sq+1 and 0m(s-1). All of the above constants are independent of the meshsize and can only depend on the underlying polynomial degree q, the space dimension d, and the mesh regularity parameter γ.

Henceforth, we use the symbol c to denote a generic positive constant, whose value can change at each occurrence, provided it is independent of the micro-scale ε, any meshsize HT or H, and the homogenized solution u0. We also track the direct dependency of the error bounds on the parameters α,β characterizing the spectrum of the diffusion matrix. The value of the generic constant c can depend on the space dimension d, the underlying polynomial degree, the mesh regularity parameter γ, and on some higher-order norms of the rescaling 𝔸β of the diffusion matrix or the correctors μl that will be made clear from the context.

4 Fine-Scale Approximation Space

Let k and let 𝒯H be a member of an admissible mesh sequence in the sense of Definition 3.1. In this section, we introduce the fine-scale approximation space on which we will base our multiscale HHO method. We first construct in Section 4.1 a set of cell-based and face-based basis functions, then we provide in Section 4.2 a local characterization of the underlying space, finally we study its approximation properties in Section 4.3.

4.1 Oscillatory Basis Functions

The oscillatory basis functions consist of cell- and face-based basis functions.

4.1.1 Cell-Based Basis Functions

Let T𝒯H. If k=0, we do not define cell-based basis functions. Assume now that k1. For all 1iNdk-1, we consider the problem

(4.1)inf{T[12𝔸εφφ-ΦTk-1,iφ]:φH1(T),ΠFk(φ)=0 for all FT}.

Problem (4.1) admits a unique minimizer. This minimizer, that we will denote φε,Tk+1,iH1(T), can be proved to solve, for real numbers (λF,jT)FT, 1jNd-1k satisfying the compatibility condition

FTFj=1Nd-1kλF,jTΦFk,j=-TΦTk-1,i,

the constrained Neumann problem

(4.2){-div(𝔸εφε,Tk+1,i)=ΦTk-1,iin T,𝔸εφε,Tk+1,i𝒏T,F=j=1Nd-1kλF,jTΦFk,jon all FT,ΠFk(φε,Tk+1,i)=0on all FT.

The superscript k+1 is meant to remind us that the functions φε,Tk+1,i are used to generate a linear space which has the same approximation capacity as the polynomial space of order at most k+1, as will be shown in Section 4.3.

Remark 4.1 (Practical Computation).

To compute the functions φε,Tk+1,i for all 1iNdk-1, one considers in practice a (shape-regular) matching simplicial mesh 𝒯hT of the cell T, with size h smaller than ε. Then one can solve problem (4.2) approximately by using a classical (equal-order) monoscale HHO method (or any other monoscale approximation method). For the implementation of the monoscale HHO method, we refer to [12]. One can either consider a weak formulation in {φH1(T):ΠFk(φ)=0 for all FT}, which leads to a coercive problem, or a weak formulation in H1(T), which leads to a saddle-point system with Lagrange multipliers. Equivalent considerations apply below to the computation of the face-based basis functions. Note that the error estimates we provide in this work for our approach do not take into account the local approximations of size h and assume that (4.2) and (4.4) below are solved exactly.

4.1.2 Face-Based Basis Functions

Let T𝒯H. For all FT and all 1jNd-1k, we consider the problem

(4.3)inf{T[12𝔸εφφ]:φH1(T),ΠFk(φ)=ΦFk,j,Πσk(φ)=0 for all σT{F}}.

Problem (4.3) admits a unique minimizer. This minimizer, that we will denote φε,T,Fk+1,jH1(T), can be proved to solve, for real numbers (λσ,qT,F)σT, 1qNd-1k satisfying the compatibility condition

σTσq=1Nd-1kλσ,qT,FΦσk,q=0,

the constrained Neumann problem

(4.4){-div(𝔸εφε,T,Fk+1,j)=0in T,𝔸εφε,T,Fk+1,j𝒏T,σ=q=1Nd-1kλσ,qT,FΦσk,qon all σT,ΠFk(φε,T,Fk+1,j)=ΦFk,jon F,Πσk(φε,T,Fk+1,j)=0on all σT{F}.

4.2 Discrete Space

We introduce, for any T𝒯H, the space

(4.5)Vε,Tk+1:={vεH1(T):div(𝔸εvε)dk-1(T),𝔸εvε𝒏Td-1k(T)},

with the convention that d-1(T):={0}. We recall that the condition 𝔸εvε𝒏Td-1k(T) is equivalent to 𝔸εvε𝒏T,Fd-1k(F) for all FT. Proceeding as in [14, Section 2.4], it can easily be shown that the dimension of Vε,Tk+1 is (Ndk-1+card(T)×Nd-1k) (or card(T) if k=0).

Proposition 4.2 (Characterization of Vε,Tk+1).

For any TTH, the family {(φε,Tk+1,i)1iNdk-1,(φε,T,Fk+1,j)FFT, 1jNd-1k} forms a basis for the space Vε,Tk+1.

Proof.

To establish the result, we only need to prove that

Vε,Tk+1Span{(φε,Tk+1,i)1iNdk-1,(φε,T,Fk+1,j)FT, 1jNd-1k},

since the converse inclusion follows from the definition of the oscillatory basis functions, and the cardinal of the family fits the dimension of Vε,Tk+1. Let vεVε,Tk+1. Then there exist real numbers (θTi)1iNdk-1 (only if k1) and (θT,Fj)FT, 1jNd-1k, satisfying the compatibility condition

FTFj=1Nd-1kθT,FjΦFk,j=-Ti=1Ndk-1θTiΦTk-1,i(=0 if k=0),

such that

{-div(𝔸εvε)=i=1Ndk-1θTiΦTk-1,i(=0 if k=0)in T,𝔸εvε𝒏T,F=j=1Nd-1kθT,FjΦFk,jon all FT.

Let us now introduce

ζ:=vε-i=1Ndk-1θTiφε,Tk+1,i-σTj=1Nd-1kxσk,j(vε)φε,T,σk+1,j,

where, for all σT, the real numbers (xσk,j(vε))1jNd-1k solve the linear system

j=1Nd-1k(σΦσk,jΦσk,q)xσk,j(vε)=σvεΦσk,qfor all 1qNd-1k.

It can easily be checked that -div(𝔸εζ)=0 in T and that 𝔸εζ𝒏T,Fd-1k(F) and ΠFk(ζ)=0 on all FT. Using the compatibility conditions, we also infer that T𝔸εζ𝒏T=0, which means that the previous system for ζ is compatible. Hence, ζ0, which concludes the proof. ∎

Remark 4.3 (Space Vε,Tk+1).

The definition of the space Vε,Tk+1 is reminiscent of that considered in the non-conforming VEM in the case where 𝔸ε=𝕀d; see [6] and also [14].

We define HTd-10(T) such that, for any FT, HTF:=HF. We will need the following inverse inequality on the normal component of 𝔸εvε for a function vεVε,Tk+1; for completeness, we also establish a bound on the divergence.

Lemma 4.4 (Inverse Inequalities).

The following holds for all vεVε,Tk+1:

HTdiv(𝔸εvε)L2(T)+HT12𝔸εvε𝒏TL2(T)cβ12𝔸ε12vεL2(T)d,

with c independent of ε, HT, α and β.

Proof.

Note that the functions on the left-hand side are (piecewise) polynomials, but the function on the right-hand side is not a polynomial in general. Let us first bound the divergence. Let dε:=div(𝔸εvε)dk-1(T). Let S be a simplicial sub-cell of T. Considering the standard bubble function bSH01(S) (equal to the scaled product of the barycentric coordinates in S taking the value one at the barycenter of S), we infer using integration by parts that, for some c>0 depending on mesh regularity,

cdεL2(S)2SdεbSdε=Sdiv(𝔸εvε)bSdε
=-S𝔸εvε(bSdε)β12𝔸ε12vεL2(S)dHS-1dεL2(S),

where the last bound follows by applying an inverse inequality to the polynomial function bSdε. Summing over all the simplicial sub-cells and invoking mesh regularity, we conclude that

div(𝔸εvε)L2(T)cβ12HT-1𝔸ε12vεL2(T)d.

Let us now bound the normal component at the boundary. Let σ be a sub-face of a face FT, and let ST be the simplex of the sub-mesh such that σ is a face of S. Then rS:=[div(𝔸εvε)]Sdk-1(S)dk(S) and rσ:=[𝔸εvε𝒏T]σd-1k(σ). Note that 𝒏Tσ=𝒏Sσ. Invoking [28, Appendix A], we infer that there is a vector-valued polynomial function 𝒒 in the Raviart–Thomas–Nédélec (RTN) finite element space of order k in S so that div(𝒒)=rS in S, 𝒒𝒏Tσ=rσ on σ, and

𝒒L2(S)dcmin𝒛𝑯(div;S)div(𝒛)=rS in S𝒛𝒏Tσ=rσ on σ𝒛L2(S)d,

with c depending on γ (but not on k) and 𝑯(div;S):={𝒛L2(S)d:div(𝒛)L2(S)}. Since the function [𝔸εvε]S is in 𝑯(div;S) and satisfies the requested conditions on the divergence in S and the normal component on σ, we conclude that 𝒒L2(S)dc𝔸εvεL2(S)d. A discrete trace inequality in the RTN finite element space shows that

𝔸εvε𝒏TL2(σ)=𝒒𝒏TL2(σ)cHσ-12𝒒L2(S)dcHσ-12𝔸εvεL2(S)d,

where c depends on γ and k. We conclude by invoking mesh regularity. ∎

4.3 Approximation Properties

We now investigate the approximation properties of the space Vε,Tk+1 for all T𝒯H. Our aim is to study how well the first-order two-scale expansion ε1(u0) can be approximated in the discrete space Vε,Tk+1. Let us define πε,Tk+1(u0)Vε,Tk+1 such that Tπε,Tk+1(u0)=Tε1(u0) and

(4.6){-div(𝔸επε,Tk+1(u0))=-div(𝔸0ΠTk+1(u0))dk-1(T)in T,𝔸επε,Tk+1(u0)𝒏T=𝔸0ΠTk+1(u0)𝒏Td-1k(T)on T.

Note that the data in (4.6) are compatible. From (4.6) we infer that, for any wH1(T),

(4.7)T𝔸επε,Tk+1(u0)w=T𝔸0ΠTk+1(u0)w.

Lemma 4.5 (Approximation in Vε,Tk+1).

Assume that the correctors μl are in W1,(Rd) for any 1ld, and that u0Hk+2(T)W1,(T). Then the following holds:

(4.8)𝔸ε12(ε1(u0)-πε,Tk+1(u0))L2(T)dcβ12ρ12(HTk+1|u0|Hk+2(T)+ε|u0|H2(T)+ε12|T|12|u0|W1,(T)),

with c independent of ε, HT, u0, α, β, and possibly depending on d, k, γ, max1ldμlW1,(Rd).

Proof.

Subtracting/adding 𝔸0u0 and using (4.7) with w=ε1(u0)|T-πε,Tk+1(u0) which is in H1(T), we infer that

𝔸ε12(ε1(u0)-πε,Tk+1(u0))L2(T)d2=T(𝔸εε1(u0)-𝔸0u0)(ε1(u0)-πε,Tk+1(u0))
+T𝔸0(u0-ΠTk+1(u0))(ε1(u0)-πε,Tk+1(u0)).

Using the Cauchy–Schwarz inequality and the fact that ε1(u0)T-πε,Tk+1(u0) has zero mean-value on T by construction, we infer that

𝔸ε12(ε1(u0)-πε,Tk+1(u0))L2(T)dβ12ρ12(u0-ΠTk+1(u0))L2(T)d+α-12supwH1(T)|ε(w)|wL2(T)d,

with

ε(w)=T(𝔸εε1(u0)-𝔸0u0)wandH1(T)={wH1(T):Tw=0}.

The first term on the right-hand side is bounded using the approximation properties (3.4) of ΠTk+1 with m=1 and s=k+2, and the second term is bounded in Lemma A.2 (take D=T). ∎

Remark 4.6 (Alternative Estimate).

An alternative estimate to (4.8) can be derived under the slightly stronger regularity assumptions that there is κ>0 so that 𝔸C0,κ(d;d×d), and that u0Hmax(k+2,3)(T). The proof of this estimate follows the strategy advocated in [39], where one invokes Lemma A.4 instead of Lemma A.2 at the end of the proof of Lemma 4.5 to infer that

𝔸ε12(ε1(u0)-πε,Tk+1(u0))L2(T)dcβ12ρ12(HTk+1|u0|Hk+2(T)+(ε+(εHT)12)|u0|H2(T)
   +εHT|u0|H3(T)+ε12HT-12|u0|H1(T)),

with c independent of ε, HT, u0, α, β, and possibly depending on d, k, γ, 𝔸βC0,κ(d;d×d). This local estimate leads to the same global error estimate for (both versions of) the msHHO method described hereafter as (4.8); see in particular the end of the proof of Theorem 5.6.

5 The msHHO Method

In this section, we introduce and analyze the multiscale HHO (msHHO) method. We consider first in Section 5.1 a mixed-order version and then in Section 5.2 an equal-order version concerning the polynomial degree used for the cell and face unknowns. Let 𝒯H be a member of an admissible mesh sequence in the sense of Definition 3.1.

5.1 The Mixed-Order Case

Let k1. For all T𝒯H, we consider the following local set of discrete unknowns:

U¯Tk:=dk-1(T)×d-1k(T).

Any element v¯TU¯Tk is decomposed as v¯T:=(vT,vT). For any FT, we denote vF:=vTFd-1k(F). We introduce the local reduction operator I¯Tk:H1(T)U¯Tk such that, for any vH1(T), I¯Tkv:=(ΠTk-1(v),ΠTk(v)), where ΠTk(v)d-1k(T) is defined, for any FT, by ΠTk(v)F:=ΠFk(v). Reasoning as in [14, Section 2.4], it can be proved that, for all T𝒯H, the restriction of I¯Tk to Vε,Tk+1 is an isomorphism from Vε,Tk+1 to U¯Tk. Thus, the triple (T,Vε,Tk+1,I¯Tk) defines a finite element in the sense of Ciarlet.

We define the local multiscale reconstruction operator

pε,Tk+1:U¯TkVε,Tk+1

such that, for any v¯T=(vT,vT)U¯Tk, pε,Tk+1(v¯T)Vε,Tk+1 satisfies

Tpε,Tk+1(v¯T)=TvT

and solves, for all wεVε,Tk+1, the well-posed local Neumann problem

(5.1)T𝔸εpε,Tk+1(v¯T)wε=-TvTdiv(𝔸εwε)+TvT𝔸εwε𝒏T.

Note that (5.1) can be equivalently rewritten

(5.2)T𝔸εpε,Tk+1(v¯T)wε=TvT𝔸εwε-T(vT-vT)𝔸εwε𝒏T.

Integrating by parts the left-hand side of (5.1) and exploiting the definition (4.5) of the space Vε,Tk+1, one can see that, for any v¯TU¯Tk,

(5.3)ΠTk-1(pε,Tk+1(v¯T))=ΠTk-1(vT)=vT,ΠTk(pε,Tk+1(v¯T))=ΠTk(vT)=vT.

Owing to (4.5) and (5.1), we infer that, for all vH1(T),

(5.4)T𝔸ε(v-pε,Tk+1(I¯Tkv))wε=0for all wεVε,Tk+1,

so that pε,Tk+1I¯Tk:H1(T)Vε,Tk+1 is the 𝔸ε-weighted elliptic projection. As a consequence, we have, for all vH1(T),

(5.5)𝔸ε12(v-pε,Tk+1(I¯Tkv))L2(T)d=infwεVε,Tk+1𝔸ε12(v-wε)L2(T)d.

Since the operator pε,Tk+1I¯Tk preserves the mean value, its restriction to Vε,Tk+1 is the identity operator.

Remark 5.1 (Comparison with the Monoscale HHO Method).

In the monoscale HHO method, the reconstruction operator is simpler to construct since it maps onto dk+1(T) (which is a proper subspace of Vε,Tk+1 whenever 𝔸ε is a constant matrix on T), whereas in the multiscale context, we explore the whole space Vε,Tk+1 to build the reconstruction. One advantage of doing this is that we no longer need stabilization in the present case. Another advantage is that we recover the characterization of pε,Tk+1I¯Tk as the 𝔸ε-weighted elliptic projector onto Vε,Tk+1, that is lost in the monoscale case as soon as 𝔸ε is not a constant matrix on T.

The local bilinear form aε,T:U¯Tk×U¯Tk is defined as

aε,T(u¯T,v¯T):=T𝔸εpε,Tk+1(u¯T)pε,Tk+1(v¯T).

We introduce the following semi-norm on U¯Tk:

(5.6)v¯TT2:=vTL2(T)d2+HT-12(vT-vT)L2(T)2.

Lemma 5.2 (Local Stability).

The following holds:

aε,T(v¯T,v¯T)cαv¯TT2for all v¯TU¯Tk,

with constant c independent of ε, HT, α and β.

Proof.

Let v¯TU¯Tk. To derive an estimate on vTL2(T)d, we define vεVε,Tk+1 such that

(5.7){-div(𝔸εvε)=-vTdk-1(T)in T,𝔸εvε𝒏T=vT𝒏Td-1k(T)on T,

and satisfying, e.g., Tvε=0 (the way the constant is fixed is unimportant here). Note that data in (5.7) are compatible. Then the following holds:

T𝔸εvεz=TvTzfor all zH1(T).

Using this last relation where we take z=pε,Tk+1(v¯T), and using (5.2) where we take wε=vεVε,Tk+1 defined in (5.7), we infer that

-TvTvT+TvTvT𝒏T=-TvTdiv(𝔸εvε)+TvT𝔸εvε𝒏T
=T𝔸εvεvT-T(vT-vT)𝔸εvε𝒏T
=T𝔸εvεpε,Tk+1(v¯T)=TvTpε,Tk+1(v¯T).

After an integration by parts, this yields

vTL2(T)d2=Tpε,Tk+1(v¯T)vT+T(vT-vT)vT𝒏T.

By the Cauchy–Schwarz inequality and the discrete trace inequality (3.1), we then obtain

(5.8)vTL2(T)dc(α-12𝔸ε12pε,Tk+1(v¯T)L2(T)d+HT-12(vT-vT)L2(T)).

To bound the second term on the right-hand side, we use (5.3) to infer that

[vT-vT]T=[ΠTk-1(pε,Tk+1(v¯T))]T-ΠTk(pε,Tk+1(v¯T))
=ΠTk(ΠTk-1(pε,Tk+1(v¯T))-pε,Tk+1(v¯T)).

Using the L2-stability of ΠTk, the continuous trace inequality (3.2), the local Poincaré inequality (3.3) (since pε,Tk+1(v¯T)-ΠTk-1(pε,Tk+1(v¯T)) has zero mean-value on T), and the H1-stability of ΠTk-1, we infer that

(5.9)HT-12(vT-vT)L2(T)cα-12𝔸ε12pε,Tk+1(v¯T)L2(T)d.

This concludes the proof. ∎

We define the skeleton 𝒯H of the mesh 𝒯H as 𝒯H:=FHF. We introduce the broken polynomial spaces

(5.10)dk-1(𝒯H):={vL2(Ω):vTdk-1(T) for all T𝒯H},
(5.11)d-1k(H):={vL2(𝒯H):vFd-1k(F) for all FH}.

The global set of discrete unknowns is defined to be

U¯Hk:=dk-1(𝒯H)×d-1k(H),

so that any v¯HU¯Hk can be decomposed as v¯H:=(v𝒯H,vH). For any given discrete unknown v¯HU¯Hk, we denote v¯T:=(vT,vT)U¯Tk its restriction to the mesh cell T𝒯H. Note that unknowns attached to mesh interfaces are single-valued, in the sense that, for any FHi such that F=T1T2Z for T1,T2𝒯H, vF:=vHFd-1k(F) is such that vF=vT1F=vT2F. To take into account homogeneous Dirichlet boundary conditions, we further introduce the subspace U¯H,0k:={v¯HU¯Hk:vF0 for all FHb}. We define the global bilinear form aε,H:U¯Hk×U¯Hk such that

aε,H(u¯H,v¯H):=T𝒯Haε,T(u¯T,v¯T)=T𝒯HT𝔸εpε,Tk+1(u¯T)pε,Tk+1(v¯T).

Then the discrete problem reads: Find u¯ε,HU¯H,0k such that

(5.12)aε,H(u¯ε,H,v¯H)=Ωfv𝒯Hfor all v¯HU¯H,0k.

Setting v¯HH2:=T𝒯Hv¯TT2 on U¯Hk, with T introduced in (5.6), this defines a norm on U¯H,0k since elements in U¯H,0k are such that vF0 for all FHb.

Lemma 5.3 (Well-Posedness).

The following holds, for all v¯HU¯Hk:

aε,H(v¯H,v¯H)=T𝒯H𝔸ε12pε,Tk+1(v¯T)L2(T)d2=:v¯Hε,H2cαv¯HH2,

with constant c independent of ε, H, α and β. As a consequence, the discrete problem (5.12) is well-posed.

Proof.

This is a direct consequence of Lemma 5.2. ∎

Remark 5.4 (Non-conforming Finite Element (ncFE) Formulation).

Consider the discrete space

Vε,H,0k+1:={vε,HL2(Ω):vε,HTVε,Tk+1 for all T𝒯H,ΠFk([[vε,H]]F)=0 for all FH},

where [[]]F denotes the jump operator for all interfaces FHi (the sign is irrelevant) and the actual trace for all boundary faces FHb. Consider the following ncFE method: Find uε,HVε,H,0k+1 such that

(5.13)a~ε,H(uε,H,vε,H)=T𝒯HTfΠTk-1(vε,H)for all vε,HVε,H,0k+1,

where

a~ε,H(uε,H,vε,H):=T𝒯HT𝔸εuε,Hvε,H.

Then, using that the restriction of I¯Tk to Vε,Tk+1 is an isomorphism from Vε,Tk+1 to U¯Tk and that the restriction of pε,Tk+1I¯Tk to Vε,Tk+1 is the identity operator, it can be shown that u¯ε,H solves (5.12) if and only if u¯ε,T=I¯Tk(uε,HT) for all T𝒯H where uε,H solves (5.13). This proves that (5.12) is indeed a high-order extension of the method in [39], up to a different treatment of the right-hand side: ΠTk-1(vε,H) is used instead of vε,H.

Let uε be the oscillatory solution to (1.1) and let u¯ε,H be the discrete msHHO solution to (5.12). Let us define the discrete error such that

(5.14)e¯ε,HU¯H,0k,e¯ε,T:=I¯Tkuε-u¯ε,Tfor all T𝒯H.

Note that e¯ε,H is well-defined as a member of U¯H,0k since the oscillatory solution uε is in H01(Ω) and functions in H01(Ω) are single-valued at interfaces and vanish at the boundary.

Lemma 5.5 (Discrete Energy-Error Estimate).

Let the discrete error e¯ε,H be defined by (5.14). Assume that u0Hk+2(Ω). Then the following holds:

(5.15)e¯ε,Hε,Hcρ12(βT𝒯HHT2(k+1)|u0|Hk+2(T)2+T𝒯H𝔸ε12(uε-πε,Tk+1(u0))L2(T)d2)12,

with constant c independent of ε, H, u0, α and β.

Proof.

Lemma 5.3 implies that

e¯ε,Hε,H=supv¯HU¯H,0kaε,H(e¯ε,H,v¯H)v¯Hε,H.

Let v¯HU¯H,0k. Performing an integration by parts, and using the facts that the flux 𝔸0u0𝒏F is continuous across any interface FHi since u0H2(Ω), and that v¯HU¯H,0k, we infer that

aε,H(u¯ε,H,v¯H)=Ωfv𝒯H=T𝒯HT𝔸0u0vT-T𝒯HT(vT-vT)𝔸0u0𝒏T.

Using (5.2) with wε=pε,Tk+1(I¯Tkuε), we then infer that

aε,H(e¯ε,H,v¯H)=T𝒯HT(𝔸εpε,Tk+1(I¯Tkuε)-𝔸0u0)vT-T𝒯HT(𝔸εpε,Tk+1(I¯Tkuε)-𝔸0u0)𝒏T(vT-vT).

Adding/subtracting ΠTk+1(u0) on the right-hand side yields aε,H(e¯ε,H,v¯H)=𝔗1+𝔗2 with

𝔗1=T𝒯HT𝔸0(ΠTk+1(u0)-u0)vT-T𝒯HT𝔸0(ΠTk+1(u0)-u0)𝒏T(vT-vT),
𝔗2=T𝒯HT(𝔸εpε,Tk+1(I¯Tkuε)-𝔸0ΠTk+1(u0))vT-T𝒯HT(𝔸εpε,Tk+1(I¯Tkuε)-𝔸0ΠTk+1(u0))𝒏T(vT-vT).

The term 𝔗1 is estimated using Cauchy–Schwarz inequality and the approximation properties (3.4) of the projector ΠTk+1 for m=1 and s=k+2, yielding

|𝔗1|cβ(T𝒯HHT2(k+1)|u0|Hk+2(T)2)12v¯HH.

Considering now 𝔗2, we use the definition (4.6) of πε,Tk+1(u0) and relation (4.7) to infer that

𝔗2=T𝒯HT𝔸ε(pε,Tk+1(I¯Tkuε)-πε,Tk+1(u0))vT-T𝒯HT𝔸ε(pε,Tk+1(I¯Tkuε)-πε,Tk+1(u0))𝒏T(vT-vT).

The first term on the right-hand side can be bounded using the Cauchy–Schwarz inequality, whereas the second term is estimated by means of the inverse inequality from Lemma 4.4 since (pε,Tk+1(I¯Tkuε)-πε,Tk+1(u0))Vε,Tk+1. This yields

|𝔗2|cβ12(T𝒯H𝔸ε12(pε,Tk+1(I¯Tkuε)-πε,Tk+1(u0))L2(T)d2)12v¯HH
cβ12(T𝒯H𝔸ε12(uε-πε,Tk+1(u0))L2(T)d2)12v¯HH,

where the last bound follows from (5.5) since πε,Tk+1(u0)Vε,Tk+1. Since v¯Hε,H2cαv¯HH2 owing to Lemma 5.3, we obtain the expected bound. ∎

Theorem 5.6 (Energy-Error Estimate).

Assume that the correctors μl are in W1,(Rd) for any integer 1ld, and that u0Hk+2(Ω) (recall that k1). Then the following holds:

(T𝒯H𝔸ε12(uε-pε,Tk+1(u¯ε,T))L2(T)d2)12cβ12ρ(T𝒯HHT2(k+1)|u0|Hk+2(T)2+ε|Ω||u0|W1,(Ω)2
(5.16)   +T𝒯H[ε2|u0|H2(T)2+ε|T||u0|W1,(T)2])12,

with c independent of ε, H, u0, α and β. In particular, if the mesh TH is quasi-uniform, and tracking for simplicity only the dependency on ε and H with εHΩ (Ω denotes the diameter of Ω), we obtain an energy-error upper bound of the form (ε12+Hk+1+(εH)12).

Proof.

Using the shorthand notation eε,T:=uε|T-pε,Tk+1(u¯ε,T) for all T𝒯H, the triangle inequality implies that

(T𝒯H𝔸ε12eε,TL2(T)d2)12(T𝒯H𝔸ε12(uε-pε,Tk+1(I¯Tkuε))L2(T)d2)12+e¯ε,Hε,H,

and owing to (5.5), we infer that

(T𝒯H𝔸ε12eε,TL2(T)d2)12(T𝒯H𝔸ε12(uε-πε,Tk+1(u0))L2(T)d2)12+e¯ε,Hε,H.

Lemma 5.5 then implies that

(T𝒯H𝔸ε12eε,TL2(T)d2)12cρ12(βT𝒯HHT2(k+1)|u0|Hk+2(T)2+T𝒯H𝔸ε12(uε-πε,Tk+1(u0))L2(T)d2)12.

To conclude the proof of (5.16), we add/subtract ε1(u0) in the last term on the right-hand side, and invoke the triangle inequality together with Lemma A.5 to bound (uε-ε1(u0)) globally on Ω and Lemma 4.5 to bound (ε1(u0)-πε,Tk+1(u0)) locally on all T𝒯H. Finally, to derive the upper bound for quasi-uniform meshes, we observe that the last term in (5.16) can be estimated as

T𝒯Hε|T||u0|W1,(T)2cεH-1|u0|W1,(Ω)2T𝒯H|T|HTcεH-1|u0|W1,(Ω)2

with c proportional to |Ω|. ∎

Remark 5.7 (Dependency on ρ).

Estimate (5.16) has a linear dependency with respect to the (global) heterogeneity/anisotropy ratio ρ (a close inspection of the proof shows that the term ε12|Ω|12|u0|W1,(Ω) only scales with ρ12). This linear scaling is also obtained with the monoscale HHO method when the diffusivity is non-constant in each mesh cell; cf. [20, Theorem 3.1].

Remark 5.8 (Discretization of the Right-Hand Side).

Note that we could also integrate the right-hand side in (5.12) using pε,Tk+1(v¯T) instead of vT on each T𝒯H, up to the addition on the right-hand sides of the bounds (5.15) and (5.16) of the optimally convergent term cα-12(T𝒯HHT2(k+1)|f|Hk(T)2)12. Indeed, owing to (5.3), we have

T𝒯HTf(vT-pε,Tk+1(v¯T))=T𝒯HT(f-ΠTk-1(f))(vT-pε,Tk+1(v¯T)),

which can be estimated by applying the Cauchy–Schwarz inequality on each T, and

  1. the approximation properties (3.4) of ΠTk-1 with m=0 and s=k for the first factor,

  2. the Poincaré inequality (3.3) (recall that (vT-pε,Tk+1(v¯T)) has zero-mean on T) and the triangle inequality combined with Lemma 5.2 for the second factor.

This alternative approach, that is pursued in [39, 40], necessitates an integration against oscillatory test functions. It is hence computationally more expensive (recall that f is assumed to be non-oscillatory), and may become limiting in a multi-query context.

5.2 The Equal-Order Case

Let k0. For all T𝒯H, we consider now the following local set of discrete unknowns:

U¯Tk:=dk(T)×d-1k(T).

Any v¯TU¯Tk is again decomposed as v¯T:=(vT,vT), and for any FT, we denote vF:=vTFd-1k(F). We redefine the local reduction operator I¯Tk:H1(T)U¯Tk so that, for any vH1(T),

I¯Tkv:=(ΠTk(v),ΠTk(v)).

Reasoning as in [14, Section 2.4], it can be proved that, for all T𝒯H, the restriction of I¯Tk to V~ε,Tk+1 is an isomorphism from V~ε,Tk+1 to U¯Tk, where

V~ε,Tk+1:={vεH1(T):div(𝔸εvε)dk(T),𝔸εvε𝒏Td-1k(T)}.

Thus, the triple (T,V~ε,Tk+1,I¯Tk) defines a finite element in the sense of Ciarlet.

The local multiscale reconstruction operator pε,Tk+1:U¯TkVε,Tk+1 is still defined as in (5.1), so that the key relations (5.4) and (5.5) still hold. In particular, pε,Tk+1I¯Tk:H1(T)Vε,Tk+1 is the 𝔸ε-weighted elliptic projection. However, the restriction of pε,Tk+1I¯Tk to the larger space V~ε,Tk+1 is not the identity operator since pε,Tk+1 maps onto the smaller space Vε,Tk+1. Concerning (5.3), we still have

ΠTk(pε,Tk+1(v¯T))=vT,

but now ΠTk-1(pε,Tk+1(v¯T))=ΠTk-1(vT) is in general different from vT.

This leads us to introduce the symmetric, positive semi-definite stabilization

jε,T(u¯T,v¯T):=αTHT-1(uT-ΠTk(pε,Tk+1(u¯T)))(vT-ΠTk(pε,Tk+1(v¯T))).

The local bilinear form aε,T:U¯Tk×U¯Tk is then defined as

aε,T(u¯T,v¯T):=T𝔸εpε,Tk+1(u¯T)pε,Tk+1(v¯T)+jε,T(u¯T,v¯T).

Remark 5.9 (Variant).

Alternatively, one can discard the stabilization at the price of computing additional cell-based oscillatory basis functions, using the basis functions (ΦTk,i)1iNdk instead of (ΦTk-1,i)1iNdk-1 as proposed in Section 4.1.1. This is the approach pursued in [40] for k=0 where one cell-based oscillatory basis function is added (in the slightly different context of perforated domains).

Recall the local stability semi-norm T defined by (5.6).

Lemma 5.10 (Local Stability and Approximation).

The following holds:

aε,T(v¯T,v¯T)cαv¯TT2for all v¯TU¯Tk.

Moreover, for all vH1(T),

(5.17)jε,T(I¯Tkv,I¯Tkv)12c𝔸ε12(v-pε,Tk+1(I¯Tkv))L2(T)d,

with (distinct) constants c independent of ε, HT, α and β.

Proof.

To prove stability, we adapt the proof of Lemma 5.2. Let v¯TU¯Tk. The bound (5.8) on vTL2(T)d still holds, so that we only need to bound HT-1/2(vT-vT)L2(T). Since ΠTk(pε,Tk+1(v¯T))=vT, we infer that

(vT-vT)=ΠTk(vT-pε,Tk+1(v¯T)),

so that invoking the L2-stability of ΠTk and the triangle inequality while adding/subtracting ΠTk(pε,Tk+1(v¯T)), we obtain

HT-12(vT-vT)L2(T)HT-12(vT-ΠTk(pε,Tk+1(v¯T)))L2(T)+HT-12(pε,Tk+1(v¯T)-ΠTk(pε,Tk+1(v¯T)))L2(T).

The first term on the right-hand side is bounded by α-12jε,T(v¯T,v¯T)12, and the second one has been bounded (with the use of ΠTk-1 instead of ΠTk) in the proof of Lemma 5.2 (see (5.9)) by cα-1/2𝔸ε1/2pε,Tk+1(v¯T)L2(T)d. To prove (5.17), we start from

jε,T(I¯Tkv,I¯Tkv)=αHT-12ΠTk(v-pε,Tk+1(I¯Tkv))L2(T)2.

The result then follows from the application of the discrete trace inequality (3.1), the L2-stability property of ΠTk, and the local Poincaré inequality (3.3) (since Tpε,Tk+1(I¯Tkv)=Tv). ∎

We define the broken polynomial space

dk(𝒯H):={vL2(Ω):vTdk(T) for all T𝒯H},

and the global set of discrete unknowns is defined to be

U¯Hk:=dk(𝒯H)×d-1k(H),

where d-1k(H) is still defined by (5.11). To take into account homogeneous Dirichlet boundary conditions, we consider again the subspace U¯H,0k:={v¯HU¯Hk:vF0 for all FHb}. We define the global bilinear form aε,H:U¯Hk×U¯Hk such that

aε,H(u¯H,v¯H):=T𝒯Haε,T(u¯T,v¯T)=T𝒯H(T𝔸εpε,Tk+1(u¯T)pε,Tk+1(v¯T)+jε,T(u¯T,v¯T)).

Then the discrete problem reads: Find u¯ε,HU¯H,0k such that

(5.18)aε,H(u¯ε,H,v¯H)=Ωfv𝒯Hfor all v¯HU¯H,0k.

Recalling the norm v¯HH2:=T𝒯Hv¯TT2 on U¯H,0k, we readily infer from Lemma 5.10 the following well-posedness result.

Lemma 5.11 (Well-Posedness).

The following holds, for all v¯HU¯Hk:

aε,H(v¯H,v¯H)=T𝒯H(𝔸ε12pε,Tk+1(v¯T)L2(T)d2+jε,T(v¯T,v¯T))=:v¯Hε,H2cαv¯HH2,

with constant c independent of ε, H, α and β. As a consequence, the discrete problem (5.18) is well-posed.

Remark 5.12 (ncFE Interpretation).

As in Remark 5.4, it is possible to give a ncFE interpretation of the scheme (5.18). Let

V~ε,H,0k+1:={vε,HL2(Ω):vε,HTV~ε,Tk+1 for all T𝒯H,ΠFk([[vε,H]]F)=0 for all FH},

and consider the following ncFE method: Find uε,HV~ε,H,0k+1 such that

(5.19)a~ε,H(uε,H,vε,H)=T𝒯HTfΠTk(vε,H)for all vε,HV~ε,H,0k+1,

where

a~ε,H(uε,H,vε,H):=T𝒯Haε,T(I¯Tk(uε,HT),I¯Tk(vε,HT)).

Then it can be shown that u¯ε,H solves (5.18) if and only if u¯ε,T=I¯Tk(uε,HT) for all T𝒯H, where uε,H solves (5.19). The main difference with respect to the mixed-order case is that it is no longer possible to simplify the expression of the bilinear form a~ε,H since the restriction of pε,Tk+1I¯Tk to V~ε,Tk+1 is not the identity operator. As in the monoscale HHO method, the operator pε,Tk+1, which maps onto the smaller space Vε,Tk+1, allows one to restrict the number of computed basis functions while maintaining optimal (and here also ε-robust) approximation properties. The functions (from the discrete space V~ε,Tk+1) that are eliminated (not computed) are handled by the stabilization term.

Lemma 5.13 (Discrete Energy-Error Estimate).

Let the discrete error e¯ε,H be defined by (5.14). Assume that u0Hk+2(Ω). Then the following holds:

e¯ε,Hε,Hcρ12(βT𝒯HHT2(k+1)|u0|Hk+2(T)2+T𝒯H𝔸ε12(uε-πε,Tk+1(u0))L2(T)d2)12,

with constant c independent of ε, H, u0, α and β.

Proof.

The only difference with the proof of Lemma 5.5 is that we now have aε,H(e¯ε,H,v¯H)=𝔗1+𝔗2+𝔗3, where 𝔗1,𝔗2 are defined and bounded in that proof and where

𝔗3:=T𝒯Hjε,T(I¯Tkuε,v¯T).

Since jε,T is symmetric, positive semi-definite, we infer that

|𝔗3|(T𝒯Hjε,T(I¯Tkuε,I¯Tkuε))12(T𝒯Hjε,T(v¯T,v¯T))12
c(T𝒯H𝔸ε12(uε-pε,Tk+1(I¯Tkuε))L2(T)d2)12v¯Hε,H,

where we have used (5.17). We can now conclude as before. ∎

Theorem 5.14 (Energy-Error Estimate).

Assume that the correctors μl are in W1,(Rd) for any 1ld, and that u0Hk+2(Ω)W1,(Ω). Then the following holds:

(T𝒯H𝔸ε12(uε-pε,Tk+1(u¯ε,T))L2(T)d2)12cβ12ρ(T𝒯HHT2(k+1)|u0|Hk+2(T)2+ε|Ω||u0|W1,(Ω)2
(5.20)   +T𝒯H[ε2|u0|H2(T)2+ε|T||u0|W1,(T)2])12,

with c independent of ε, H, u0, α and β. In particular, if the mesh TH is quasi-uniform, and tracking for simplicity only the dependency on ε and H with εHΩ, we obtain an energy-error upper bound of the form (ε12+Hk+1+(εH)12).

Proof.

Identical to that of Theorem 5.6. ∎

Remark 5.15 (Dependency on ρ).

As in the mixed-order case (cf. Remark 5.7), estimate (5.20) has a linear dependency with respect to the (global) heterogeneity/anisotropy ratio ρ.

Remark 5.16 (Discretization of the Right-Hand Side).

The same observation as in Remark 5.8 concerning the discretization of the right-hand side in (5.18) is still valid for the equal-order case.

6 Numerical Results

In this section, we discuss the organization of the computations and we present some numerical results illustrating the above analysis for both the mixed-order and equal-order msHHO methods. Our numerical results have been obtained using the Disk++ library, which is available as open-source under MPL license at the address https://github.com/datafl4sh/diskpp. The numerical core of the library is described in [12]. For the numerical tests presented below, we have used the direct solver PARDISO of the Intel MKL library. The simulations were run on an Intel i7-3615QM (2.3 GHz) with 16 Gb of RAM.

6.1 Offline/Online Solution Strategy

Let us consider the equal-order version (k0) of the msHHO method introduced in Section 5.2. Similar considerations carry over to the mixed-order case (k1) of Section 5.1. To solve problem (5.18), we adopt an offline/online strategy.

  1. In the offline step, all the computations are local, and independent of the right-hand side f. We first compute the cell-based and face-based basis functions, i.e., for all T𝒯H, we compute the Ndk-1 functions φε,Tk+1,i solution to (4.2), and the card(T)×Nd-1k functions φε,T,Fk+1,j solution to (4.4) (cf. Remark 4.1). This first substep is fully parallelizable. In a second time, we compute the multiscale reconstruction operators pε,Tk+1, by solving (5.1) for all T𝒯H. Each computation requires to invert a symmetric positive-definite matrix of size (Ndk-1+card(T)×Nd-1k), which can be performed effectively via Cholesky factorization. This second substep is as well fully parallelizable. Finally, we perform static condensation locally in each cell of 𝒯H, to eliminate the cell unknowns. Details can be found in [20, Section 3.3.1]. Basically, in each cell, this substep consists in inverting a symmetric positive-definite matrix of size Ndk. This last substep is also fully parallelizable.

  2. In the online step, we compute the L2-orthogonal projection of the right-hand side f onto dk(𝒯H), and we then solve a symmetric positive-definite global problem, posed in terms of the face unknowns only. The size of this problem is card(Hi)×Nd-1k. If one wants to compute an approximation of the solution to (1.1) for another f (or for other boundary conditions), only the online step must be rerun.

6.2 Periodic Test-Case

We consider the periodic test-case studied in [39] (and also in [43]). We let d=2, and let Ω be the unit square. We consider problem (1.1), with right-hand side f(x,y)=sin(x)sin(y), and oscillatory coefficient

(6.1)𝔸ε(x,y)=a(xε,yε)𝕀2,a(x1,x2)=1+100cos2(πx1)sin2(πx2).

For the coefficient (6.1), the homogenized tensor is given by 𝔸06.72071𝕀2. We fix ε=π1500.021.

Figure 1 Periodic test-case: convergence results in energy-norm for mesh levels l∈{0,…,6}{l\in\{0,\ldots,6\}}; mixed-order msHHO method with polynomial degrees k∈{1,2}{k\in\{1,2\}} and equal-order msHHO method with polynomial degrees k∈{0,1,2}{k\in\{0,1,2\}}. The red vertical line indicates the value of ε.
Figure 1

Periodic test-case: convergence results in energy-norm for mesh levels l{0,,6}; mixed-order msHHO method with polynomial degrees k{1,2} and equal-order msHHO method with polynomial degrees k{0,1,2}. The red vertical line indicates the value of ε.

We consider a sequence of hierarchical triangular meshes of size Hl=0.43×2-l with l{0,,9}, so that H5<ε<H4. A reference solution is computed by solving (1.1) with the (equal-order) monoscale HHO method on the mesh of level lref=9 with polynomial degree kref=2. In Figure 1, we present the (absolute) energy-norm errors obtained with the msHHO method on the meshes 𝒯Hl with l{0,,6}. We consider both the mixed-order msHHO method with polynomial degrees k{1,2} and the equal-order msHHO method with polynomial degrees k{0,1,2}. In all cases, the cell- and face-based oscillatory basis functions are precomputed using the (equal-order) monoscale HHO method on the mesh of level losc=8 with polynomial degree kosc=1. We have verified that the oscillatory basis functions are sufficiently well resolved by comparing our results to those obtained with kosc=2 and obtaining only very marginal differences. The first observation we draw from Figure 1 is that the mixed-order and equal-order msHHO methods employing the same polynomial degree for the face unknowns deliver very similar results; indeed, the error curves are barely distinguishable both for k=1 and k=2. Moreover, we can observe all the main features expected from the error analysis: a pre-asymptotic regime where the term Hk+1 essentially dominates (meshes of levels l{0,1}), the resonance regime (meshes of levels l{2,3,4} essentially), and the asymptotic regime where the mesh actually resolves the fine scale of the model coefficients (meshes of levels l{5,6}). We can also see the advantages of using a higher polynomial order for the face unknowns: the error is overall smaller, the minimal error in the resonance regime is reached at a larger value of H and takes a smaller value (incidentally, the maximal error in the resonance regime takes a smaller value as well), and the asymptotic regime starts for larger values of H.

6.3 Locally Periodic Test-Case

Figure 2 Locally periodic test-case: convergence results in energy-norm for mesh levels l∈{0,…,6}{l\in\{0,\ldots,6\}}; mixed-order msHHO method with polynomial degrees k∈{1,2}{k\in\{1,2\}} and equal-order msHHO method with polynomial degrees k∈{0,1,2}{k\in\{0,1,2\}}. The red vertical line indicates the value of ε.
Figure 2

Locally periodic test-case: convergence results in energy-norm for mesh levels l{0,,6}; mixed-order msHHO method with polynomial degrees k{1,2} and equal-order msHHO method with polynomial degrees k{0,1,2}. The red vertical line indicates the value of ε.

Keeping the same two-dimensional domain Ω as in the periodic test-case of Section 6.2, we consider now a locally periodic test-case where we solve problem (1.1) with unchanged right-hand side f(x,y)=sin(x)sin(y), but with oscillatory coefficient

𝔸ε(x,y)=(a(xε,yε)+ex2+y22)𝕀2,with a given in (6.1),

and with unchanged value of ε. We perform the same numerical experiments as in Section 6.2 using the same mesh level and polynomial order parameters for computing the reference solution and the oscillatory basis functions (we verified similarly the adequate resolution of the oscillatory basis functions). Results are reported in Figure 2. We can draw the same conclusions as in the periodic test-case: similarity of the results delivered by the mixed-order and the equal-order msHHO methods for both k=1 and k=2, presence of the pre-asymptotic, resonance, and asymptotic regimes, and advantages of using a higher polynomial order for the face unknowns.

To briefly assess computational costs, we compute, for those mesh levels in the pre-asymptotic or resonance regimes for which the error is minimal, the computational times to perform the offline and online steps. We report the results in Table 1. We also report the number of degrees of freedom in the global system solved in the online step. We make the experiment for the equal-order msHHO method of orders k=0 and k=2, for respective mesh levels l=2 and l=1. We do not make use of parallelism in our implementation to compute the results. Table 1 shows the interest of higher-order approximations, since a better accuracy is reached at a smaller online computational cost.

Table 1

Offline and online computational times.

Energy-errorOffline time (s)Online time (s)#DoFs
k=0(l=2)0.003386122540.026408
k=2(l=1)0.002646485200.018288

A Estimates on the First-Order Two-Scale Expansion

In this appendix, we derive various useful estimates on the first-order two-scale expansion ε1(u0) defined by (2.5). Except for Lemma A.4, these estimates are classical; we provide (short) proofs since we additionally track the direct dependency of the constants on the parameters α and β characterizing the spectrum of 𝔸 and on the various length scales present in the problem.

A.1 Dual-Norm Estimates

Let D be an open, connected, polytopal subset of Ω; in this work, we will need the cases where D=Ω or where D=T𝒯H. Let D be a length scale associated with D, e.g., its diameter. Our goal is to bound the dual norm of the linear map such that

(A.1)wε(w):=D(𝔸εε1(u0)-𝔸0u0)w

for all wH01(D) (Dirichlet case), or for all wH1(D):={wH1(D):Dw=0} (Neumann case); note that ε(w) does not change if the values of w are shifted by a constant.

Lemma A.1 (Dual Norm, Dirichlet Case).

Assume that the homogenized solution u0 belongs to H2(D) and that, for any 1ld, the corrector μl belongs to W1,(Rd). Then

supwH01(D)|ε(w)|wL2(D)dcβε|u0|H2(D),

with c independent of ε, D, u0, α, β, and possibly depending on d, and on max1ldμlW1,(Rd).

Proof.

For any integer 1id, we have

[𝔸εε1(u0)]i=j=1d[𝔸ε]ijjε1(u0)
=j=1d[𝔸ε]ij(ju0+εl=1d(1εε(jμl)lu0+ε(μl)j,l2u0))
(A.2)=[𝔸0u0]i+l=1dε(θil)lu0+εl,j=1d[𝔸ε]ijε(μl)j,l2u0,

with θil:=𝔸il+j=1d𝔸ijjμl-[𝔸0]il satisfying the following properties:

  1. θilLper(Q) by assumption on 𝔸 and on the correctors μl,

  2. Qθil=0 as a consequence of (2.2),

  3. i=1diθil=0 in d as a consequence of (2.3).

Adapting [37, equation (1.11)] (see also [30, Sections I.3.1 and I.3.3]), we infer that, for any integer 1ld, there exists a skew-symmetric matrix 𝕋lWper1,(Q)d×d, satisfying Q𝕋l=𝟘 and such that, for any integer 1id,

(A.3)θil=q=1dq𝕋qil.

Plugging (A.3) into (A.2), we infer that, for any integer 1id,

[𝔸εε1(u0)]i-[𝔸0u0]i=ε(l,q=1dq(ε(𝕋qil))lu0+l,j=1d[𝔸ε]ijε(μl)j,l2u0).

Since

q(ε(𝕋qil))lu0=q(ε(𝕋qil)lu0)-ε(𝕋qil)q,l2u0,

and recalling the definition (A.1) of ε, this yields

ε(w)=ε(i,l,j=1dD[𝔸ε]ijε(μl)j,l2u0iw-i,l,q=1dDε(𝕋qil)q,l2u0iw)
(A.4)+εi,l,q=1dDq(ε(𝕋qil)lu0)iw.

Since 𝕋qil=-𝕋iql for any integers 1i,qd, we infer by integration by parts of the last term that

ε(w)=ε(i,l,j=1dD[𝔸ε]ijε(μl)j,l2u0iw-i,l,q=1dDε(𝕋qil)q,l2u0iw)
(A.5)+εi,l,q=1dDq(ε(𝕋qil)lu0)nD,iw,

where 𝒏D is the unit outward normal to D. Since wH01(D), we obtain

ε(w)=ε(i,l,j=1dD[𝔸ε]ijε(μl)j,l2u0iw-i,l,q=1dDε(𝕋qil)q,l2u0iw).

Using the Cauchy–Schwarz inequality, we finally deduce that

supwH01(D)|ε(w)|wL2(D)dcβεmax1ld(μlL(d),β-1𝕋lL(d)d×d)|u0|H2(D).

We conclude by observing that 𝕋lL(d)d×dcmax1idθilL(d)cβ. ∎

Lemma A.2 (Dual Norm, Neumann Case (i)).

Assume that the homogenized solution u0 belongs to the space H2(D)W1,(D) and that, for any 1ld, the corrector μl belongs to W1,(Rd). Then

(A.6)supwH1(D)|ε(w)|wL2(D)dcβ(ε|u0|H2(D)+|D|12ε12|u0|W1,(D)),

with c independent of ε, D, u0, α, β, and possibly depending on d, and on max1ldμlW1,(Rd).

Proof.

Our starting point is (A.4). The first two terms on the right-hand side are responsible for a contribution of order βε|u0|H2(D), and it only remains to bound the last term. Following the ideas of [37, p. 29], we define, for η>0, the domain Dη:={𝒙D:dist(𝒙,D)<η}. If η is above a critical value (which scales as D), Dη=D, otherwise DηD. We introduce the cut-off function ζηC0(D¯) such that ζη0 on D, defined by ζη(𝒙)=1ηdist(𝒙,D) if 𝒙Dη, and ζη(𝒙)=1 if 𝒙DDη. We have 0ζη1 and max1qdqζηL(D)η-1. We first infer that

εi,l,q=1dDq(ε(𝕋qil)lu0)iw=εi,l,q=1dDηq((1-ζη)ε(𝕋qil)lu0)iw,

since (1-ζη) vanishes identically on DDη and since

i,l,q=1dDq(ζηε(𝕋qil)lu0)iw=0

as can be seen by integration by parts, using the fact that 𝕋qil=-𝕋iql for any integers 1i,qd, and the fact that ζη vanishes identically on D. Then, accounting for the fact that

εq((1-ζη)ε(𝕋qil)lu0)=-εqζηε(𝕋qil)lu0+(1-ζη)ε(q𝕋qil)lu0+ε(1-ζη)ε(𝕋qil)q,l2u0,

we infer that

|εi,l,q=1dDq(ε(𝕋qil)lu0)iw|c[|Dη|12(εη+1)(max1ld𝕋lW1,(d)d×d)|u0|W1,(D)
+ε(max1ld𝕋lL(d)d×d)|u0|H2(D)]wL2(D)d.

Using now the estimate |Dη|η|D|, the fact that max1ld𝕋lW1,(d)d×dcβ, and since the function ηεη+η is minimal for η=ε, we finally infer the bound (A.6). ∎

Remark A.3 (Weaker Regularity Assumption).

Without the regularity assumption u0W1,(D), one can still invoke a Sobolev embedding since u0H2(D). The second term between the parentheses on the right-hand side of (A.6) becomes

c(p)(|D|εD-d)12-1p(|u0|H1(D)+D|u0|H2(D)),

where p=6 for d=3 and p can be taken as large as wanted for d=2 (note that c(p)+ when p+ in that case). The derivation of estimates in this setting is considered in [43]. Therein, the authors claim that their method is able to get rid of the resonance error (without oversampling). We believe there is an issue with the bound [43, equation (27)], which should exhibit the resonant contribution (εD)12-1p|u0|H1(D).

Lemma A.4 (Dual Norm, Neumann Case (ii)).

Assume that D=TTH, where TH is a member of an admissible mesh sequence in the sense of Definition 3.1; set D=HT. Assume that the homogenized solution u0 belongs to H3(D) and that there is κ>0 so that AC0,κ(Rd;Rd×d). Then

supwH1(D)|ε(w)|wL2(D)dcβ((ε+(εD)12)|u0|H2(D)+εD|u0|H3(D)+ε12D-12|u0|H1(D)),

with c independent of ε, D, u0, α, β, and possibly depending on d, γ, and AβC0,κ(Rd;Rd×d).

Proof.

We proceed as in the proof of Lemma A.1. Concerning the regularity of θil, we now have θilC0,ι(d) for some ι>0 since the Hölder continuity of 𝔸 on d implies the Hölder continuity of μl and μl on d for any 1ld; cf., e.g., [29, Theorem 8.22 and Corollary 8.36]. Following [37, pp. 6–7] and [39, pp. 131–132], we infer that the skew-symmetric matrix 𝕋l is such that 𝕋lC1(d)d×d. Our starting point is (A.5). The first two terms on the right-hand side are responsible for a contribution of order βε|u0|H2(D), and it only remains to bound the last term. We have

εi,l,q=1dDq(ε(𝕋qil)lu0)nD,iw=εi,l,q=1dDε(𝕋qil)q,l2u0nD,iw+i,l,q=1dDε(q𝕋qil)lu0nD,iw
=:𝔗1+𝔗2.

Using the Cauchy–Schwarz inequality and the trace inequality (3.2), the first term on the right-hand side can be estimated as

|𝔗1|cβεD-1(|u0|H2(D)+D|u0|H3(D))(wL2(D)+DwL2(D)d),

since max1ld𝕋lC0(d)d×dcβ. Observing that Dw=0, we can use the Poincaré inequality (3.3) to infer that

|𝔗1|cβε(|u0|H2(D)+D|u0|H3(D))wL2(D)d.

To estimate the second term on the right-hand side, we adapt the ideas from [39, Lemma 4.6]. Considering the matching simplicial sub-mesh of D, let us collect in the set 𝔉D all the sub-faces composing the boundary of D. Then we can write

𝔗2=σ𝔉Dl=1dq=1dq<idσε(𝕋qil)𝝉σqilu0w,

where the vectors 𝝉σqi are such that 𝝉σqi21 and 𝝉σqi𝒏Dσ=0. Then using a straightforward adaptation of the result in [39, Lemma 4.6], and since max1ld𝕋lC1(d)d×dcβ, we infer that

|σε(𝕋qil)𝝉σqilu0w|cβε12HS-32(|u0|H1(S)+HS|u0|H2(S))(wL2(S)+HSwL2(S)d),

where S is the simplicial sub-cell of D having σ as face. Collecting the contributions of all the sub-faces σ𝔉D and using the mesh regularity assumptions on D, we infer that

|𝔗2|cβε12D-32(|u0|H1(D)+D|u0|H2(D))(wL2(D)+DwL2(D)d).

Finally, invoking the Poincaré inequality (3.3) since w has zero mean-value in D yields

|𝔗2|cβε12D-12(|u0|H1(D)+D|u0|H2(D))wL2(D)d.

Collecting the above bounds on 𝔗1 and 𝔗2 concludes the proof. ∎

A.2 Global Energy-Norm Estimate

Lemma A.5 (Energy-Norm Estimate).

Assume that the homogenized solution u0 belongs to H2(Ω)W1,(Ω), and that, for any 1ld, the corrector μl belongs to W1,(Rd). Then

𝔸ε12(uε-ε1(u0))L2(Ω)dcβ12(ρ12ε|u0|H2(Ω)+|Ω|12ε12|u0|W1,(Ω)),

with c independent of ε, Ω, u0, α, β, and possibly depending on d, and on max1ldμlW1,(Rd).

Proof.

The regularity assumptions on u0 and the correctors imply (uε-ε1(u0))H1(Ω); however, we do not have (uε-ε1(u0))H01(Ω). Following the ideas in [37, p. 28], we define, for η>0, the domain Ωη:={𝒙Ω:dist(𝒙,Ω)<η}. If η is above a critical value, Ωη=Ω, otherwise ΩηΩ. We introduce the cut-off function ζηC0(Ω¯) such that ζη0 on Ω, defined by ζη(𝒙)=1ηdist(𝒙,Ω) if 𝒙Ωη, and ζη(𝒙)=1 if 𝒙ΩΩη. We have 0ζη1 and max1idiζηL(Ω)η-1. The function ζη allows us to define a corrected first-order two-scale expansion ε1,0(u0):=u0+εζηl=1dε(μl)lu0 such that (uε-ε1,0(u0))H01(Ω). We start with the triangle inequality:

(A.7)𝔸ε12(uε-ε1(u0))L2(Ω)d𝔸ε12(uε-ε1,0(u0))L2(Ω)d+𝔸ε12(ε1(u0)-ε1,0(u0))L2(Ω)d.

Let us focus on the first term on the right-hand side of (A.7). We have

𝔸ε12(uε-ε1,0(u0))L2(Ω)d2=Ω𝔸ε(uε-ε1(u0))(uε-ε1,0(u0))
+Ω𝔸ε(ε1(u0)-ε1,0(u0))(uε-ε1,0(u0)).

Since (uε-ε1,0(u0))H01(Ω), we infer that

𝔸ε12(uε-ε1,0(u0))L2(Ω)dα-12supwH01(Ω)|Ω𝔸ε(uε-ε1(u0))w|wL2(Ω)d
(A.8)+𝔸ε12(ε1(u0)-ε1,0(u0))L2(Ω)d.

Since Ω𝔸εuεw=Ω𝔸0u0w for any wH01(Ω) in view of (1.1) and (2.4), estimates (A.7) and (A.8) lead to

(A.9)𝔸ε12(uε-ε1(u0))L2(Ω)dα-12supwH01(Ω)|ε(w)|wL2(Ω)d+2β12(ε1(u0)-ε1,0(u0))L2(Ω)d,

recalling that ε(w)=Ω(𝔸εε1(u0)-𝔸0u0)w. Since we can bound the first term on the right-hand side of (A.9) using Lemma A.1 (with D=Ω), it remains to estimate the second term. Owing to the definition of ζη, we infer that

(ε1(u0)-ε1,0(u0))L2(Ω)d=ε((1-ζη)l=1dε(μl)lu0)L2(Ωη)d.

For any integer 1id, we have

i((1-ζη)l=1dε(μl)lu0)=-iζηl=1dε(μl)lu0+(1-ζη)εl=1dε(iμl)lu0+(1-ζη)l=1dε(μl)i,l2u0,

and using the properties of the cut-off function ζη, we infer that

ε((1-ζη)l=1dε(μl)lu0)L2(Ωη)dc(|Ωη|12(εη+1)|u0|W1,(Ω)+ε|u0|H2(Ω)).

Since |Ωη||Ω|η, and choosing η=ε to minimize the function ηεη+η, we can conclude the proof (note that ρ1 by definition). ∎

Acknowledgements

The authors are thankful to Alexei Lozinski (LMB, Université de Franche-Comté) for fruitful discussions on the topic.

References

[1] A. Abdulle, W. E, B. Engquist and E. Vanden-Eijnden, The heterogeneous multiscale method, Acta Numer. 21 (2012), 1–87. 10.1017/S0962492912000025Search in Google Scholar

[2] G. Allaire, Shape Optimization by the Homogenization Method, Appl. Math. Sci. 146, Springer, New York, 2002. 10.1007/978-1-4684-9286-6Search in Google Scholar

[3] G. Allaire and R. Brizzi, A multiscale finite element method for numerical homogenization, SIAM Multiscale Model. Simul. 4 (2005), no. 3, 790–812. 10.1137/040611239Search in Google Scholar

[4] R. Araya, C. Harder, D. Paredes and F. Valentin, Multiscale hybrid-mixed method, SIAM J. Numer. Anal. 51 (2013), no. 6, 3505–3531. 10.1137/120888223Search in Google Scholar

[5] D. N. Arnold, F. Brezzi, B. Cockburn and L. D. Marini, Unified analysis of discontinuous Galerkin methods for elliptic problems, SIAM J. Numer. Anal. 39 (2002), no. 5, 1749–1779. 10.1137/S0036142901384162Search in Google Scholar

[6] B. Ayuso de Dios, K. Lipnikov and G. Manzini, The nonconforming virtual element method, ESAIM Math. Model. Numer. Anal. (M2AN) 50 (2016), no. 3, 879–904. 10.1051/m2an/2015090Search in Google Scholar

[7] F. Bassi, L. Botti, A. Colombo, D. A. Di Pietro and P. Tesini, On the flexibility of agglomeration-based physical space discontinuous Galerkin discretizations, J. Comput. Phys. 231 (2012), no. 1, 45–65. 10.1016/j.jcp.2011.08.018Search in Google Scholar

[8] M. Bebendorf, A note on the Poincaré inequality for convex domains, Z. Anal. Anwend. 22 (2003), no. 4, 751–756. 10.4171/ZAA/1170Search in Google Scholar

[9] L. Beirão da Veiga, F. Brezzi, A. Cangiani, G. Manzini, L. D. Marini and A. Russo, Basic principles of virtual element methods, Math. Models Meth. Appl. Sci. (M3AS) 23 (2013), 199–214. 10.1142/S0218202512500492Search in Google Scholar

[10] A. Cangiani, E. H. Georgoulis and P. Houston, hp-version discontinuous Galerkin methods on polygonal and polyhedral meshes, Math. Models Meth. Appl. Sci. (M3AS) 24 (2014), no. 10, 2009–2041. 10.1007/978-3-319-67673-9Search in Google Scholar

[11] E. T. Chung, S. Fu and Y. Yang, An enriched multiscale mortar space for high contrast flow problems, Commun. Comput. Phys. 23 (2018), no. 2, 476–499. 10.4208/cicp.OA-2016-0147Search in Google Scholar

[12] M. Cicuttin, D. A. Di Pietro and A. Ern, Implementation of discontinuous skeletal methods on arbitrary-dimensional, polytopal meshes using generic programming, J. Comput. Appl. Math. (2017), 10.1016/j.cam.2017.09.017. 10.1016/j.cam.2017.09.017Search in Google Scholar

[13] B. Cockburn, Static condensation, hybridization, and the devising of the HDG methods, Building Bridges: Connections and Challenges in Modern Approaches to Numerical Partial Differential Equations, Lect. Notes Comput. Sci. Eng. 114, Springer, Cham (2016), 129–177. 10.1007/978-3-319-41640-3_5Search in Google Scholar

[14] B. Cockburn, D. A. Di Pietro and A. Ern, Bridging the hybrid high-order and hybridizable discontinuous Galerkin methods, ESAIM Math. Model. Numer. Anal. (M2AN) 50 (2016), no. 3, 635–650. 10.1051/m2an/2015051Search in Google Scholar

[15] B. Cockburn, J. Gopalakrishnan and R. Lazarov, Unified hybridization of discontinuous Galerkin, mixed, and continuous Galerkin methods for second-order elliptic problems, SIAM J. Numer. Anal. 47 (2009), no. 2, 1319–1365. 10.1137/070706616Search in Google Scholar

[16] D. A. Di Pietro and A. Ern, Mathematical Aspects of Discontinuous Galerkin Methods, Math. Appl. 69, Springer, Berlin, 2012. 10.1007/978-3-642-22980-0Search in Google Scholar

[17] D. A. Di Pietro and A. Ern, A hybrid high-order locking-free method for linear elasticity on general meshes, Comput. Methods Appl. Mech. Engrg. 283 (2015), 1–21. 10.1016/j.cma.2014.09.009Search in Google Scholar

[18] D. A. Di Pietro and A. Ern, Hybrid high-order methods for variable-diffusion problems on general meshes, C. R. Acad. Sci. Paris Ser. I 353 (2015), 31–34. 10.1016/j.crma.2014.10.013Search in Google Scholar

[19] D. A. Di Pietro, A. Ern and S. Lemaire, An arbitrary-order and compact-stencil discretization of diffusion on general meshes based on local reconstruction operators, Comput. Methods Appl. Math. 14 (2014), no. 4, 461–472. 10.1515/cmam-2014-0018Search in Google Scholar

[20] D. A. Di Pietro, A. Ern and S. Lemaire, A review of hybrid high-order methods: Formulations, computational aspects, comparison with other methods, Building Bridges: Connections and Challenges in Modern Approaches to Numerical Partial Differential Equations Lect. Notes Comput. Sci. Eng. 114, Springer, Cham (2016), 205–236. 10.1007/978-3-319-41640-3_7Search in Google Scholar

[21] W. E and B. Engquist, The heterogeneous multiscale methods, Commun. Math. Sci. 1 (2003), 87–132. 10.4310/CMS.2003.v1.n1.a8Search in Google Scholar

[22] Y. Efendiev, J. Galvis and T. Y. Hou, Generalized multiscale finite element methods (GMsFEM), J. Comput. Phys. 251 (2013), 116–135. 10.1016/j.jcp.2013.04.045Search in Google Scholar

[23] Y. Efendiev and T. Y. Hou, Multiscale Finite Element Methods – Theory and Applications, Surv. Tutor. Appl. Math. Sci. 4, Springer, New York, 2009. Search in Google Scholar

[24] Y. Efendiev, T. Y. Hou and X.-H. Wu, Convergence of a nonconforming multiscale finite element method, SIAM J. Numer. Anal. 37 (2000), no. 3, 888–910. 10.1137/S0036142997330329Search in Google Scholar

[25] Y. Efendiev, R. Lazarov, M. Moon and K. Shi, A spectral multiscale hybridizable discontinuous Galerkin method for second order elliptic problems, Comput. Methods Appl. Mech. Engrg. 292 (2015), 243–256. 10.1016/j.cma.2014.09.036Search in Google Scholar

[26] Y. Efendiev, R. Lazarov and K. Shi, A multiscale HDG method for second order elliptic equations. Part I. Polynomial and homogenization-based multiscale spaces, SIAM J. Numer. Anal. 53 (2015), no. 1, 342–369. 10.1137/13094089XSearch in Google Scholar

[27] A. Ern and J.-L. Guermond, Finite element quasi-interpolation and best approximation, ESAIM Math. Model. Numer. Anal. (M2AN) 51 (2017), no. 4, 1367–1385. 10.1051/m2an/2016066Search in Google Scholar

[28] A. Ern and M. Vohralík, Stable broken H1 and H(div) polynomial extensions for polynomial-degree-robust potential and flux reconstruction in three space dimensions, preprint (2016), https://hal.inria.fr/hal-01422204. Search in Google Scholar

[29] D. Gilbarg and N. S. Trudinger, Elliptic Partial Differential Equations of Second Order, Classics Math., Springer, Berlin, 2001. 10.1007/978-3-642-61798-0Search in Google Scholar

[30] V. Girault and P.-A. Raviart, Finite Element Methods for Navier–Stokes Equations. Theory and Algorithms, Springer Ser. Comput. Math. 5, Springer, Berlin, 1986. 10.1007/978-3-642-61623-5Search in Google Scholar

[31] A. Gloria, Numerical homogenization: Survey, new results, and perspectives, ESAIM Proc. 37 (2012), 50–116. 10.1051/proc/201237002Search in Google Scholar

[32] P. Henning and D. Peterseim, Oversampling for the multiscale finite element method, SIAM Multiscale Model. Simul. 11 (2013), no. 4, 1149–1175. 10.1137/120900332Search in Google Scholar

[33] J. S. Hesthaven, S. Zhang and X. Zhu, High-order multiscale finite element method for elliptic problems, SIAM Multiscale Model. Simul. 12 (2014), no. 2, 650–666. 10.1137/120898024Search in Google Scholar

[34] T. Y. Hou and X.-H. Wu, A multiscale finite element method for elliptic problems in composite materials and porous media, J. Comput. Phys. 134 (1997), 169–189.10.1006/jcph.1997.5682Search in Google Scholar

[35] T. Y. Hou, X.-H. Wu and Z. Cai, Convergence of a multiscale finite element method for elliptic problems with rapidly oscillating coefficients, Math. Comp. 68 (1999), no. 227, 913–943. 10.1090/S0025-5718-99-01077-7Search in Google Scholar

[36] T. Y. Hou, X.-H. Wu and Y. Zhang, Removing the cell resonance error in the multiscale finite element method via a Petrov–Galerkin formulation, Commun. Math. Sci. 2 (2004), no. 2, 185–205. 10.4310/CMS.2004.v2.n2.a3Search in Google Scholar

[37] V. V. Jikov, S. M. Kozlov and O. A. Oleinik, Homogenization of Differential Operators and Integral Functionals, Springer, Berlin, 1994. 10.1007/978-3-642-84659-5Search in Google Scholar

[38] A. Konaté, Méthode multi-échelle pour la simulation d’écoulements miscibles en milieux poreux, Ph.D. thesis, Université Pierre et Marie Curie, 2017, https://tel.archives-ouvertes.fr/tel-01558994. Search in Google Scholar

[39] C. Le Bris, F. Legoll and A. Lozinski, MsFEM à la Crouzeix–Raviart for highly oscillatory elliptic problems, Chin. Ann. Math. Ser. B 34 (2013), no. 1, 113–138. 10.1007/978-3-642-41401-5_11Search in Google Scholar

[40] C. Le Bris, F. Legoll and A. Lozinski, An MsFEM-type approach for perforated domains, SIAM Multiscale Model. Simul. 12 (2014), no. 3, 1046–1077. 10.1137/130927826Search in Google Scholar

[41] A. Målqvist and D. Peterseim, Localization of elliptic multiscale problems, Math. Comp. 83 (2014), 2583–2603. 10.1090/S0025-5718-2014-02868-8Search in Google Scholar

[42] L. Mu, J. Wang and X. Ye, A weak Galerkin generalized multiscale finite element method, J. Comput. Appl. Math. 305 (2016), 68–81. 10.1016/j.cam.2016.03.017Search in Google Scholar

[43] D. Paredes, F. Valentin and H. M. Versieux, On the robustness of multiscale hybrid-mixed methods, Math. Comp. 86 (2017), 525–548. 10.1090/mcom/3108Search in Google Scholar

[44] N. Sukumar and A. Tabarraei, Conforming polygonal finite elements, Internat. J. Numer. Methods Engrg. 61 (2004), no. 12, 2045–2066. 10.1002/nme.1141Search in Google Scholar

[45] A. Veeser and R. Verfürth, Poincaré constants for finite element stars, IMA J. Numer. Anal. 32 (2012), no. 1, 30–47. 10.1093/imanum/drr011Search in Google Scholar

[46] E. L. Wachspress, A Rational Finite Element Basis, Math. Sci. Eng. 114, Academic Press, New York, 1975. Search in Google Scholar

[47] J. Wang and X. Ye, A weak Galerkin finite element method for second-order elliptic problems, J. Comput. Appl. Math. 241 (2013), 103–115. 10.1016/j.cam.2012.10.003Search in Google Scholar

Received: 2018-02-15
Accepted: 2018-06-01
Published Online: 2018-06-19
Published in Print: 2019-10-01

© 2020 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 14.1.2025 from https://www.degruyter.com/document/doi/10.1515/cmam-2018-0013/html
Scroll to top button