Home Superconvergent DPG Methods for Second-Order Elliptic Problems
Article Publicly Available

Superconvergent DPG Methods for Second-Order Elliptic Problems

  • Thomas Führer ORCID logo EMAIL logo
Published/Copyright: April 6, 2019

Abstract

We consider DPG methods with optimal test functions and broken test spaces based on ultra-weak formulations of general second-order elliptic problems. Under some assumptions on the regularity of solutions of the model problem and its adjoint, superconvergence for the scalar field variable is achieved by either increasing the polynomial degree in the corresponding approximation space by one or by a local postprocessing. We provide a uniform analysis that allows the treatment of different test norms. Particularly, we show that in the presence of convection only the quasi-optimal test norm leads to higher convergence rates, whereas other norms considered do not. Moreover, we also prove that our DPG method delivers the best L2 approximation of the scalar field variable up to higher-order terms, which is the first theoretical explanation of an observation made previously by different authors. Numerical studies that support our theoretical findings are presented.

MSC 2010: 65N30; 65N12

1 Introduction

In this work we investigate convergence rates of DPG methods based on an ultra-weak formulation of second-order elliptic problems stated in the form of the general first-order system

(1.1)

(1.1a)u-𝜷u+𝑪𝝈=𝑪𝒇in Ω,
(1.1b)div𝝈+γu=fin Ω,
(1.1c)u=0on Γ:=Ω,

where Ωd, d2, is a polyhedral domain and 𝑪L(Ω)d×d denotes a symmetric, uniformly positive definite matrix valued function, 𝜷L(Ω)d, γL(Ω). Throughout we suppose that the coefficients additionally satisfy

(1.2)L(Ω)12div(𝑪-1𝜷)+γ0,

which implies that for fL2(Ω), 𝒇𝑳2(Ω):=L2(Ω)d our model problem (1.1) admits a unique solution (u,𝝈) with uH01(Ω), 𝝈𝑯(div;Ω):={𝝉𝑳2(Ω):div𝝉L2(Ω)}. To see this, use (1.1a) in (1.1b) which results in a second-order elliptic problem. Testing with vH01(Ω) gives a bilinear form that is, using (1.2), coercive and, thus, solvability can be obtained by classical arguments.

In this work we consider DPG methods with optimal test functions and broken test spaces, which have been introduced by Demkowicz and Gopalakrishnan, see [5, 6] and also [8, 22]. For a unified stability analysis which also covers our model problem we refer to [3]. We analyze ultra-weak formulations of (1.1), which are obtained by multiplying with locally supported functions and integration by parts, see, e.g., [7] for a Poisson model problem. On the one hand, this has the advantage that the field variables can be sought in L2(Ω), since no derivative operator is applied to these unknowns after integration by parts. On the other hand, this requires the introduction of trace variables u^ and σ^ that live on the skeleton (these unknowns impose weak continuity conditions). However, as analyzed in the recent work [21] the use of ultra-weak formulations also allows to define conforming finite element spaces on polygonal meshes.

The motivation of this work is to analyze superconvergence properties for approximations of the scalar field variable u that have been observed in our recent work [10] for a simple reaction-diffusion problem, where 𝑪 is the identity matrix, 𝜷=0, γ=1, and 𝒇=0. Here we generalize and extend [10] to the model problem (1.1) and introduce new ideas that allow the treatment of different test norms. As in [10], the proofs rely on duality arguments and regularity theory for elliptic PDEs. Such arguments are common when proving higher convergence rates, e.g., the classical Aubin-Nitsche trick, or more recently in variants of DG methods, e.g., [4]. Some early works on convergence in mixed finite element methods include [9, 11, 20].

Let us also mention the recent works [14, 15] that deal with dual problems in the context of DPG methods (the DPG* method and goal-oriented problems). Particularly, we point out the reference [1]. There the authors consider a primal DPG method (without the first-order reformulation) for the Poisson problem and analyze convergence rates (with reduced degrees in test spaces). Moreover, they develop duality arguments and prove that the error in the primal variable u converges at a higher rate when measured in a weaker norm.

1.1 Summary of Results

We seek approximations uh𝒫p(𝒯), 𝝈h𝒫p(𝒯)d of the field variables u, 𝝈, where 𝒯 is a mesh of simplices and 𝒫p(𝒯) denotes the space of 𝒯-piecewise polynomials of degree less than or equal to p0, and approximations u^h, σ^h of the traces u^, σ^ in spaces that will be defined later on. For sufficient regular solutions basic a priori analysis arguments give the estimate

𝒖-𝒖hUu-uh+𝝈-𝝈h+(u^-u^h,σ^-σ^h)𝒮=𝒪(hp+1),

where denotes the L2(Ω) norm and 𝒮 is some appropriate norm for the traces. This estimate is optimal, since we seek approximations of u and 𝝈 in polynomial spaces of the same order and their errors are measured in L2(Ω) norms. Nevertheless, it is unsatisfactory to some extent. Consider 𝑪 the identity, 𝜷=0, 𝒇=0 in (1.1). Then 𝝈=u and we seek approximations of u and its gradient 𝝈 in polynomial spaces of the same order, which seems to be suboptimal. Fortunately, there exist at least two possibilities to achieve higher convergence rates under some assumptions on the regularity of solutions of (1.1) and its adjoint problem:

  1. Augmenting the trial space: Instead of seeking approximations uh𝒫p(𝒯) we seek approximations uh+𝒫p+1(𝒯) and show that

    u-uh+=𝒪(hp+2).
  2. Postprocessing: We use a common local postprocessing technique (see, e.g., the early works [11, 20]) to obtain an approximation u~h𝒫p+1(𝒯) and prove that

    u-u~h=𝒪(hp+2).

Based on similar techniques we also provide a proof of the following:

  1. DPG for ultra-weak formulations delivers the L2(Ω) best approximation up to a higher-order term, i.e., for the approximation uh𝒫p(𝒯) it holds that

    u-uhu-Πpu+𝒪(hp+2),

    where Πp denotes the L2(Ω) projection to 𝒫p(𝒯).

The latter observation is quite interesting, because it shows that even though we do not aim for higher convergence rates (by increasing the polynomial degree in the trial space or by postprocessing) we get highly accurate approximations. We stress that this result has been observed in various numerical experiments, particularly also for more complex model problems like Stokes [17], but up to now a rigorous proof has not been given.

If 𝜷=0, we show that these results hold true when using different test norms (one of them is the so-called quasi-optimal test norm or graph norm). Surprisingly (at this point), for 𝜷0 the results are only valid if the quasi-optimal test norm is used, although all test norms under consideration are equivalent. This is also observed in our numerical studies.

1.2 Basic Ideas

For the proofs of the main results, we develop duality arguments and show approximation results (Lemma 8 and Lemma 10). To get the essential idea, consider the abstract formulation: Find 𝒖U such that

b(𝒖,𝒗)=F(𝒗)for all 𝒗V,

where U denotes the trial space and V the test space. With the trial-to-test operator Θ:UV,

(Θ𝒘,𝒗)V=b(𝒘,𝒗)for all 𝒗V,

the ideal DPG method reads: Find 𝒖hUhU such that

b(𝒖h,Θ𝒘h)=F(Θ𝒘h)for all 𝒘hUh.

Then we solve a dual problem: For some given gL2(Ω), we determine 𝒗V and 𝒘=Θ-1𝒗U, both unique, and employ Galerkin orthogonality to obtain

(u-uh,g)=b(𝒖-𝒖h,𝒗)=b(𝒖-𝒖h,Θ𝒘)=b(𝒖-𝒖h,Θ(𝒘-𝒘h))𝒖-𝒖hU𝒘-𝒘hU

for arbitrary 𝒘hUh.

For the case where we want to show that the approximation uh𝒫p(𝒯) is nearly the L2(Ω) best approximation, we have g=Πp(u-uh)=Πpu-uh. Therefore,

g2=(u-uh,g)𝒖-𝒖hU𝒘-𝒘hU𝒖-𝒖hUhg.

The latter estimate is what we have to show. Suppose that it holds. With the estimate for 𝒖-𝒖hU from above, it is straightforward to see that

u-uhu-Πpu+Πpu-uh=u-Πpu+g=u-Πpu+𝒪(hp+2).

Let us come back to the essential estimate

𝒘-𝒘hUhg.

It holds if we would know that the higher derivatives of 𝒘 exist (in some sense) and can be bounded by the norm of g, so that, formally,

𝒘-𝒘hUhDhigher𝒘hg

by some standard arguments. In our case we have that 𝒗H01(Ω)×𝑯(div;Ω)V is the solution to the adjoint problem of (1.1) and under some assumptions has the higher regularity 𝒗H2(Ω)×𝑯1(𝒯)𝑯(div;Ω), where 𝑯1(𝒯) denotes 𝒯-piecewise Sobolev functions. Recall that 𝒘=Θ-1𝒗. One difficulty is that the inverse of the trial-to-test operator does not map regular functions back to regular functions. However, it turns out (Lemma 8) that 𝒘 can be written as

𝒘=(g,0,0,0)+~𝒘+𝒘,

where components of ~𝒘U are related to the dual solution 𝒗, which is sufficiently regular and 𝒘 is the solution of the (primal) problem (1.1) with data f and 𝒇 depending on the dual solution 𝒗 so that 𝒘 has sufficient regularity as well. Let us point out that this idea used in the proofs is new and allows to treat different test norms. In [10], which deals with a simple reaction-diffusion problem and one specific test norm only, the representation of 𝒘 is obtained by integration by parts using the dual solution 𝒗 and it is not clear if that approach can be generalized to the present setting. Here, in the general case we have to consider the regularity of the dual solution 𝒗 and the regularity of the solution 𝒘 of the primal problem. For the proofs it is also necessary that g is a function in the finite element space, so that we can choose 𝒘h=(g,0,0,0)+¯𝒘h, where ¯𝒘h is the best approximation of ~𝒘+𝒘. Then we show that the above estimates hold true.

Let us note that Θ is defined through the inner product in the test space. Thus, the representation of 𝒘=Θ-1𝒗 from above strongly depends on the choice of the test norm and has to be analyzed for each norm individually (this is done in Lemma 8). Earlier Keith, Vaziri Astaneh and Demkowicz [15] considered the optimal test norm. In our notation this would yield 𝒘=(g,0,0,0), i.e., ~𝒘=0=𝒘.

Moreover, the ideas so far dealt with the ideal DPG method. In this paper we work out all results for the practical DPG method under standard assumptions, i.e., the existence of Fortin operators. This implies that we have to deal with additional discretization errors.

Finally, we note that higher convergence rates for the dual variable 𝝈 can not be obtained with the same arguments since the components of 𝒗=(v,𝝉) with

(𝝈-𝝈h,𝒈)=b(𝒖-𝒖h,𝒗)

are less regular than in the case described above.

1.3 Outline

The remainder of the paper is organized as follows: Section 2 introduces basic notations, states the assumptions, and presents the main results (Theorem 35). The proofs of these theorems are postponed to Section 3, which also includes an a priori convergence estimate (Theorem 6) and the important auxiliary results Lemma 8, 10. In Section 4 we present two numerical experiments. The final Section 5 concludes this work with some remarks.

2 Main Results

2.1 Notation

We make use of the notation , i.e., AB means that there exists a constant C>0, which is independent of relevant quantities, such that ACB. Moreover, AB means that both directions hold, i.e., AB and BA.

2.2 Mesh

Let 𝒯 denote a regular mesh of Ω consisting of simplices T and let 𝒮:={T:T𝒯} denote the skeleton. We suppose that 𝒯 is shape-regular, i.e., there exists a constant κ𝒯>0 such that

maxT𝒯diam(T)d|T|κ𝒯,

where |T| denotes the volume measure of T𝒯. As usual, h:=h𝒯:=maxT𝒯diam(T) denotes the mesh-size.

2.3 Ultra-Weak Formulation

Before we derive the ultra-weak formulation of (1.1) in this subsection, we introduce some notation. Let T𝒯. We denote by (,)T the L2(T) scalar product and with T the induced norm. On boundaries T, the L2(T) scalar product is denoted by ,T and extended to the duality between the spaces H1/2(T) and H-1/2(T). Furthermore, we define the piecewise trace operators

γ0,𝒮:H1(Ω)T𝒯H1/2(T),(γ0,𝒮v)|T=v|T,
γ𝒏,𝒮:𝑯(div;Ω)T𝒯H-1/2(T),(γ𝒏,𝒮𝝉)|T=𝝉𝒏T|T,

where 𝒏T denotes the normal on T pointing from T to its complement. With these operators we define the trace spaces

H01/2(𝒮):=γ0,𝒮(H01(Ω)),
H-1/2(𝒮):=γ𝒏,𝒮(𝑯(div;Ω)).

These Hilbert spaces are equipped with minimum energy extension norms

u^1/2,𝒮:=inf{uH1(Ω):γ0,𝒮u=u^},
σ^-1/2,𝒮:=inf{𝝈𝑯(div;Ω):γ𝒏,𝒮𝝈=σ^}.

We use the broken test spaces

H1(𝒯):={vL2(Ω):v|TH1(T) for all T𝒯},
𝑯(div;𝒯):={𝝉𝑳2(Ω):𝝉|T𝑯(div;T) for all T𝒯}

and define the piecewise differential operators 𝒯:H1(𝒯)𝑳2(Ω), div𝒯:𝑯(div;𝒯)L2(Ω) on each T𝒯 by

𝒯v|T:=(v|T),
div𝒯𝝉|T:=div(𝝉|T).

Moreover, we define the following dualities for all u^H01/2(𝒮), 𝝉𝑯(div;𝒯), σ^H-1/2(𝒮), vH1(𝒯):

u^,𝝉𝒏𝒮:=T𝒯u^|T,𝝉𝒏T|TT,
σ^,v𝒮:=T𝒯σ^|T,v|TT.

These dualities measure the jumps of 𝒗=(v,𝝉)H1(𝒯)×𝑯(div;𝒯), i.e.,

(2.1)

(2.1a)vH01(Ω)σ^,v𝒮=0for all σ^H-1/2(𝒮),
(2.1b)𝝉𝑯(div;Ω)u^,𝝉𝒏𝒮=0for all u^H01/2(𝒮),

see, e.g., [3, Theorem 2.3].

The ultra-weak formulation is then derived from (1.1) by testing (1.1a) with 𝝉𝑯(div;𝒯), (1.1b) with vH1(𝒯), and piecewise integration by parts, i.e.,

-(u,div𝒯𝝉)+γ0,𝒮u,𝝉𝒏𝒮-(𝜷u,𝝉)+(𝑪𝝈,𝝉)=(𝑪𝒇,𝝉),
-(𝝈,𝒯v)+γ𝒏,𝒮𝝈,v𝒮+(γu,v)=(f,v).

Here, (,):=(,)Ω is the L2(Ω) scalar product with norm . Set

U:=L2(Ω)×𝑳2(Ω)×H01/2(𝒮)×H-1/2(𝒮),
V:=H1(𝒯)×𝑯(div;𝒯)

and define F:V and b:U×V by

F(𝒗):=(f,v)+(𝒇,𝑪𝝉),
b(𝒖,𝒗):=(u,-div𝒯𝝉-𝜷𝝉+γv)+(𝝈,𝑪𝝉-𝒯v)+u^,𝝉𝒏𝒮+σ^,v𝒮

for all 𝒖=(u,𝝈,u^,σ^)U, 𝒗=(v,𝝉)V. The ultra-weak formulation then reads: Find 𝒖U such that

(2.2)b(𝒖,𝒗)=F(𝒗)for all 𝒗V.

2.4 DPG Method and Approximation

In U we use the canonical norm,

𝒖U2:=u2+𝝈2+u^1/2,𝒮2+σ^-1/2,𝒮2for 𝒖=(u,𝝈,u^,σ^)U.

For the test space V we define the three different norms

(2.3)

(2.3a)𝒗V,qopt2:=-div𝒯𝝉-𝜷𝝉+γv2+𝑪1/2𝝉-𝑪-1/2𝒯v2+𝑪1/2𝝉2+v2,
(2.3b)𝒗V,12:=𝑪-1/2𝒯v2+v2+div𝒯𝝉2+𝑪1/2𝝉2,
(2.3c)𝒗V,22:=𝒯v2+v2+div𝒯𝝉2+𝝉2

for 𝒗=(v,𝝉)V and denote by (,)V, the corresponding scalar products. Note that all norms in (2.3) are equivalent with equivalence constants depending on the coefficients 𝑪, 𝜷, γ. However, our main results hold for the quasi-optimal test norm V,qopt under mild assumptions on the coefficient 𝜷, whereas they hold for V,1, V,2 only if 𝜷=0, i.e., for symmetric problems.

We stress that b:U×V is a bounded bilinear form and satisfies the infsup conditions with mesh independent constant. This can be proved with the theory developed in [3]. For our model problem we explicitly refer to [3, Example 3.7] for the details. There it is assumed that div(𝑪-1𝜷)=0 and γ0. We note that their analysis can also be done with our more general assumption (1.2).

The DPG method seeks an approximation 𝒖hUhU of the solution 𝒖U using the optimal test space Θ(Uh), where Θ:UV is defined by

(2.4)(Θ𝒘,𝒗)V=b(𝒘,𝒗)for all 𝒘U,𝒗V.

Then 𝒖hUh is the solution of

b(𝒖h,𝒗h)=F(𝒗h)for all 𝒗hΘ(Uh).

An essential feature of DPG is that infsup stability directly transfers to the discrete problem. However, in practice we replace Θ by a discrete version Θh:UhVhV defined by

(Θh𝒘h,𝒗h)V=b(𝒘h,𝒗h)for all 𝒘hUh,𝒗hVh.

Then the practical DPG method reads: Find 𝒖hUh such that

(2.5)b(𝒖h,Θh𝒘h)=F(Θh𝒘h)for all 𝒘hUh.

In this work we deal with the piecewise polynomial trial spaces

Uhp:=𝒫p(𝒯)×𝒫p(𝒯)d×𝒫c,0p+1(𝒮)×𝒫p(𝒮),
Uhp+:=𝒫p+1(𝒯)×𝒫p(𝒯)d×𝒫c,0p+1(𝒮)×𝒫p(𝒮)

and the piecewise polynomial test spaces

Vhk:=𝒫k1(𝒯)×𝒫k2(𝒯)d.

Here, we set

𝒫p(T):={vL2(T):v is polynomial of degree p},
𝒫p(𝒯):={vL2(Ω):v|T𝒫p(T),T𝒯},𝒫c,0p+1(𝒯):=𝒫p+1(𝒯)H01(Ω),
𝒫c,0p+1(𝒮):=γ0,𝒮(𝒫c,0p+1(𝒯)),𝒫p(𝒮):=γ𝒏,𝒮(𝒯p(𝒯)),

where

𝒯p(𝒯)={𝝉𝑯(div;Ω):𝝉|T(𝒙)=𝒂+𝒙b,𝒂𝒫p(T)d,b~𝒫p(T),T𝒯}

is the space of Raviart–Thomas functions (here ~𝒫p(T) denotes the space of homogeneous polynomials of degree p).

We also use the space C1(𝒯):={vL(Ω):v|TC1(T¯),T𝒯}.

2.5 Fortin Operators

It is well known, see, e.g., [12], that (2.5) satisfies infsup conditions (and therefore admits a unique solution) if there exists a Fortin operator ΠF:VVh such that

(2.6)ΠF𝒗VCF𝒗Vandb(𝒖h,𝒗)=b(𝒖h,ΠF𝒗)for all 𝒗V,𝒖hUh.

Throughout, we suppose that a Fortin operator exists for the discrete polynomial trial and test spaces under consideration and that CF depends only on 𝑪, 𝜷, γ, p0, and the shape-regularity of 𝒯. Let us note that for general coefficients 𝑪, 𝜷, γ the existence of such operators is not known, except for some special cases, i.e., the Poisson model problem where 𝑪 is the identity and 𝜷=0=γ. Fortin operators for the latter problem on simplicial meshes have been constructed and analyzed in [3, 12]. We refer also to [16] for the construction and analysis of Fortin operators for second-order problems.

Supposing the existence of an Fortin operator, i.e., (2.6), we have:

Proposition 1.

Problems (2.2), (2.5) admit unique solutions 𝐮=(u,𝛔,u^,σ^)U, 𝐮hUh and

𝒖-𝒖hUCoptmin𝒘hUh𝒖-𝒘hU.

The constant Copt>0 depends only on Ω, 𝐂, 𝛃, γ, pN0, and the shape-regularity of T.

2.6 Adjoint Problem and Regularity Assumptions

We define the adjoint problem (in the sense of L2(Ω)-adjoints) of (1.1) as

(2.7)

(2.7a)-div𝝉-𝜷𝝉+γv=gin Ω,
(2.7b)𝑪𝝉-v=𝑪𝒈in Ω,
(2.7c)u=0on Γ.

Supposing (1.2) this problem admits a unique solution (v,𝝉)H01(Ω)×𝑯(div;Ω) for gL2(Ω), 𝒈𝑳2(Ω).

For our results we make use of the following assumptions:

Assumption.

We suppose that the coefficients 𝑪, 𝜷, γ and the domain Ω are such that for f,gL2(Ω), 𝒇,𝒈𝑯1(𝒯)𝑯(div;Ω) the unique solutions (u,𝝈)H01(Ω)×𝑯(div;Ω) resp. (v,𝝉)H01(Ω)×𝑯(div;Ω) of (1.1) resp. (2.7) satisfy

(2.8)

(2.8a)uH2(Ω)+𝝈𝑯1(𝒯)C(f+𝒇𝑯1(𝒯)),
(2.8b)vH2(Ω)+𝝉𝑯1(𝒯)C(g+𝒈𝑯1(𝒯)).

Here, Hs(Ω) is the usual notation for norms in the Sobolev space Hs(Ω) (s>0), and is the L2(Ω) norm and 𝑯s(𝒯) the broken Sobolev norm for vector valued functions. We note that the constant C>0 strongly depends on the material data, i.e., the coefficients 𝑪, 𝜷, γ. For example, a small constant diffusion implies a blow-up of C.

Remark 2.

The regularity estimates (2.8) are satisfied if d=2, 𝑪 is the identity matrix, 𝜷C1(𝒯)d𝑯(div;Ω) and Ω is convex. This can be seen as follows: The first component uH01(Ω) of the solution of (1.1) satisfies

-Δu=f-div𝒇-(div𝜷)u+𝜷u-γuL2(Ω).

Then uH2(Ω) and uH2(Ω) is bounded by the L2(Ω) norm of the right-hand side, since Ω is a convex polyhedral domain, see [13]. Finally, the second equation of the model problem (1.1) shows

𝝈𝑯1(𝒯)=𝒇-u+𝜷u𝑯1(𝒯)𝒇𝑯1(𝒯)+uH2(Ω)f+𝒇𝑯1(𝒯).

Similarly, one shows (2.8b) (even a less regular coefficient 𝜷 suffices for the adjoint problem).

2.7 Assumptions on Coefficients and Test Norms

Besides the assumptions on the coefficients and the domain to ensure unique solvability of problems (1.1) and (2.7) and estimates (2.8), we also need some additional assumptions on the coefficients that are listed in Table 1. We emphasize that 𝜷=0 in Cases (b) and (c) is also necessary in general. In particular, in Section 4 we provide a simple example where 𝜷0 and the choice 𝒗V=𝒗V,1 or 𝒗V=𝒗V,2 does not lead to higher convergence rates, whereas 𝒗V=𝒗V,qopt does.

Table 1

Additional assumptions (besides (1.2) and (2.8)) on the coefficients for the three test norms under consideration.

CaseTest norm V𝑪𝜷γ
(a)V,qoptC1(𝒯)d×dC1(𝒯)dC1(𝒯)
(b)V,1C1(𝒯)d×d0C1(𝒯)
(c)V,2C0,1(Ω¯)d×dC1(𝒯)d×d0C1(𝒯)

2.8 L2(Ω) Projection

Our first main result shows that the DPG method with ultra-weak formulation delivers up to a higher-order term the L2(Ω) best approximation for the scalar field variable. To that end let Πp:L2(Ω)𝒫p(𝒯) denote the L2(Ω) projector.

Theorem 3.

Consider one of Cases (a), (b), or (c). Let 𝐮=(u,𝛔,u^,σ^)U be the solution of (2.2) for some given fL2(Ω), 𝐟𝐋2(Ω) and suppose uHp+2(Ω), 𝛔𝐇p+1(T). Let 𝐮h=(uh,𝛔h,u^h,σ^h)Uh:=Uhp be the solution of the practical DPG method (2.5). Suppose Pc,01(T)×RTp(T)Vhk. It holds that

u-Πpuu-uhu-Πpu+Chp+2(uHp+2(Ω)+𝝈𝑯p+1(𝒯)).

The constant C>0 depends only on Ω, 𝐂, 𝛃, γ, pN0, and shape-regularity of T.

2.9 Higher Convergence Rate by Increasing Polynomial Degree

Our second main result shows that higher convergence rates for the scalar field variable are obtained by increasing the polynomial degree in the approximation space.

Theorem 4.

Consider one of Cases (a), (b), or (c). Let 𝐮=(u,𝛔,u^,σ^)U be the solution of (2.2) for some given fL2(Ω), 𝐟𝐋2(Ω) and suppose uHp+2(Ω), 𝛔𝐇p+1(T). Let 𝐮h+=(uh+,𝛔h,u^h,σ^h)Uh:=Uhp+ be the solution of the practical DPG method (2.5). Suppose Pc,01(T)×RTp+1(T)Vhk. It holds that

u-uh+Chp+2(uHp+2(Ω)+𝝈𝑯p+1(𝒯)).

The constant C>0 depends only on Ω, 𝐂, 𝛃, γ, pN0, and shape-regularity of T.

2.10 Higher Convergence Rate by Postprocessing

Our third and final main result shows that higher convergence rates for the scalar field variable are obtained by postprocessing the solution: Let 𝒖h=(uh,𝝈h,u^h,σ^h)Uh:=Uhp be the solution of (2.5). We define u~h𝒫p+1(𝒯) on each element T𝒯 as the solution of the local Neumann problem

(2.9)

(2.9a)(u~h,vh)T=(𝑪𝒇-𝑪𝝈h+𝜷uh,vh)Tfor all vh𝒫p+1(T),
(2.9b)(u~h,1)T=(uh,1)T.

Let us note that this type of postprocessing is common in literature and can already be found in the early works [11, 20].

Theorem 5.

Consider one of Cases (a), (b), or (c). Let 𝐮=(u,𝛔,u^,σ^)U be the solution of (2.2) for some given fL2(Ω), 𝐟𝐋2(Ω) and suppose uHp+2(Ω), 𝛔𝐇p+1(T). Let 𝐮h=(uh,𝛔h,u^h,σ^h)Uh:=Uhp be the solution of the practical DPG method (2.5) and define u~hPp+1(T) by (2.9). Suppose Pc,01(T)×RTp(T)Vhk. It holds that

u-u~hChp+2(uHp+2(Ω)+𝝈𝑯p+1(𝒯)).

The constant C>0 depends only on Ω, 𝐂, 𝛃, γ, pN0, and shape-regularity of T.

3 Proofs

In this section we prove the results stated in Theorems 3, 4, and 5. First, in Section 3.1 we collect some standard results on projection operators and consider approximation results with respect to U. Second, Section 3.2 recalls the equivalent mixed formulation of the practical DPG method. Then Section 3.3 provides auxiliary results that allow to prove the main results in a uniform fashion. Finally, in Sections 3.4, 3.5, 3.6 we give the proofs of our main results.

3.1 Projection Operators and Approximation Results

Throughout let p0. Let Πp:L2(Ω)𝒫p(𝒯) denote the L2(Ω) projector. For 𝝉𝑳2(Ω) the term Πp𝝉 is understood as the application of Πp to each component. We have the (local) approximation properties

(3.1)

(3.1a)u-ΠpuCphp+1|u|Hp+1(𝒯)and𝝈-Πp𝝈Cphp+1|𝝈|𝑯p+1(𝒯),

where ||Hn(𝒯):=D𝒯n with D𝒯n denoting the 𝒯-elementwise n-th derivative operator. Let

Πp+1:H01(Ω)𝒫c,0p+1(𝒯)

denote the Scott–Zhang projection operator [19] or any other operator with the property

(3.1b)u-Πp+1uH1(Ω)Cphp+1uHp+2(Ω).

Moreover, let Πdivp:𝑯(div;Ω)𝑯1(𝒯)𝒯p(𝒯) denote the Raviart–Thomas operator, which satisfies

(3.1c)𝝈-Πdivp𝝈Cphk+1|𝝈|𝑯k+1(𝒯)for k[0,p],

and the commutativity property

divΠdivp𝝈=Πpdiv𝝈.

Note that Πdivp is well defined for functions 𝝈𝑯(div;Ω)𝑯1(𝒯): First, normal traces of 𝝈𝑯1(𝒯) are well defined on each facet of T, T𝒯, in the sense of L2(T), i.e., 𝝈𝒏TL2(T) and, second, 𝝈𝑯(div;Ω) implies unisolvency of normal traces. The constant Cp>0 in (3.1) depends only on p0 and shape-regularity of 𝒯.

The following result is an adaptation of [10, Theorem 5 and Corollary 6].

Theorem 6.

Let pN0 and let wHp+2(Ω), 𝛘𝐇p+1(T)𝐇(div;Ω). Define 𝐰:=(w,𝛘,γ0,Sw,γ𝐧,S𝛘)U. If Uh{Uhp,Uhp+}, then

min𝒘hUh𝒘-𝒘hUChp+1(wHp+2(Ω)+𝝌𝑯p+1(𝒯)).

The constant C>0 depends only on p and shape-regularity of T.

Proof.

Define

𝒘h:=(Πpw,Πp𝝌,γ0,𝒮Πp+1w,γ𝒏,𝒮Πdivp𝝌)Uh.

We estimate the terms in

𝒘-𝒘hU2=w-Πpw2+𝝌-Πp𝝌2+γ0,𝒮(w-Πp+1w)1/2,𝒮2+γ𝒏,𝒮(𝝌-Πdivp𝝌)-1/2,𝒮2.

First, we follow [10, Proof of Theorem 5] to estimate γ𝒏,𝒮(𝝌-Πdivp𝝌)-1/2,𝒮: To this end, we start with the identity from [3, Theorem 2.3], i.e.,

γ𝒏,𝒮(𝝌-Πdivp𝝌)-1/2,𝒮=sup0vH1(𝒯)γ𝒏,𝒮(𝝌-Πdivp𝝌),v𝒮vH1(𝒯).

Then elementwise integration by parts and the commutativity property yield

γ𝒏,𝒮(𝝌-Πdivp𝝌),v𝒮=(𝝌-Πdivp𝝌,𝒯v)+(div(𝝌-Πdivp𝝌),v)
=(𝝌-Πdivp𝝌,𝒯v)+((1-Πp)div𝝌,v).

Using the L2 projection property and the approximation properties (3.1a) and (3.1c), we estimate the last two terms by

|(𝝌-Πdivp𝝌,𝒯v)|+|((1-Πp)div𝝌,v)|=|(𝝌-Πdivp𝝌,𝒯v)|+|((1-Πp)div𝝌,(1-Πp)v)|
hp+1|𝝌|𝑯p+1(𝒯)𝒯v+hp|div𝝌|Hp(𝒯)h𝒯v
hp+1|𝝌|𝑯p+1(𝒯)𝒯v.

Putting the last estimates together this shows that

γ𝒏,𝒮(𝝌-Πdivp𝝌)-1/2,𝒮hp+1𝝌𝑯p+1(𝒯).

Next, observe that γ0,𝒮()1/2,𝒮H1(Ω) by definition of the norms. Finally, applying the approximation properties (3.1a)–(3.1b) and putting altogether finishes the proof. ∎

Remark 7.

As pointed out in [10] the estimate γ𝒏,𝒮(𝝌-Πdivp𝝌)-1/2,𝒮hp+1𝝌𝑯p+1(𝒯) in the proof of Theorem 6 is non-trivial. A direct application of the trace theorem gives

γ𝒏,𝒮(𝝌-Πdivp𝝌)-1/2,𝒮𝝌-Πdivp𝝌+div(𝝌-Πdivp𝝌).

Using the commutativity property and the approximation properties (3.1a), (3.1c) we get

𝝌-Πdivp𝝌+div(𝝌-Πdivp𝝌)hp+1|𝝌|Hp+1(𝒯)+hp+1|div𝝌|Hp+1(𝒯).

Thus, in order to get the same convergence rate as in Theorem 6 we have to assume the higher regularity div𝝌Hp+1(𝒯).

3.2 Mixed Formulation of the Practical DPG Method

The practical DPG method (2.5) can be reformulated as a mixed problem, see, e.g., [1]. Recall that we made the assumption of the existence of a Fortin operator (2.6). The mixed DPG formulation then reads: Find (𝒖h,𝜺hk)Uh×Vhk such that

(3.2)

(3.2a)(𝜺hk,𝒗hk)V+b(𝒖h,𝒗hk)=F(𝒗hk)for all 𝒗hkVhk,
(3.2b)b(𝒘h,𝜺hk)=0for all 𝒘hUh.

The Riesz representation 𝜺hkVhk of the residual (sometimes also called the error representation function) satisfies

𝜺hkV𝒖-𝒖hU,

under assumption (2.6), see [2, Theorem 2.1]. Note that the solution 𝒖h in (3.2) is identical to the solution of (2.5). Recall that the residual on the continuous level vanishes and, therefore, the Riesz representation is zero. Setting 𝜺:=0, we have that (𝒖,𝜺)U×V satisfies the mixed formulation for all test functions (𝒘,𝒗)U×V. In particular, we have Galerkin orthogonality

a((𝒖-𝒖h),(𝜺-𝜺hk),(𝒘h,𝒗hk))=0for all (𝒘h,𝒗hk)Uh×Vhk,

where a((𝒘,𝒗),(δ𝒘,δ𝒗)):=b(𝒘,δ𝒗)+(𝒗,δ𝒗)V-b(δ𝒘,𝒗) for all 𝒘,δ𝒘U, 𝒗,δ𝒗V.

3.3 Auxiliary Results

Recall the adjoint problem (2.7) with gL2(Ω), 𝒈=0,

(3.3)-div𝝉-𝜷𝝉+γv=g,v-𝑪𝝉=0,v|Γ=0.

Note that 𝒗=(v,𝝉)H01(Ω)×𝑯(div;Ω)V. In particular, there exists a unique 𝒘U with Θ𝒘=𝒗, since Θ:UV is an isomorphism. Note that by the definition of the trial-to-test operator (2.4), the element 𝒘 depends on the choice of scalar products in V. This is investigated in the following result.

Lemma 8.

Let gL2(Ω) and let 𝐯:=(v,𝛕)H01(Ω)×𝐇(div;Ω) denote the solution of (3.3). The unique element 𝐰U with Θ𝐰=𝐯 has the following representation depending on the cases from Section 2.7:

  1. Case (a) (V=V,qopt):

    𝒘=(g,0,0,0)+(u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*),

    where (u*,𝝈*)H01(Ω)×𝑯(div;Ω) solves (1.1) with f=v and 𝒇=𝝉.

  2. Case (b) (V=V,1):

    𝒘=(g-γv,0,0,γ𝒏,𝒮𝝉)+(u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*),

    where (u*,𝝈*)H01(Ω)×𝑯(div;Ω) solves (1.1) with f=γ(γv-g)-div𝝉+v and 𝒇=𝝉.

  3. Case (c) (V=V,2):

    𝒘=(g-γv,0,0,γ𝒏,𝒮(𝑪𝝉))+(u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*),

    where (u*,𝝈*)H01(Ω)×𝑯(div;Ω) solves (1.1) with f=γ(γv-g)-div(𝑪𝝉)+v and 𝒇=𝑪-1𝝉.

Moreover,

(3.4)vH2(Ω)+𝝉𝑯1(𝒯)+u*H2(Ω)+𝝈*𝑯1(𝒯)Cg.

For Case (c) it also holds that 𝛕,𝛔*𝐇1(Ω).

Proof.

We consider the three cases. Case (a). Recall that

(Θ𝒘,(μ,𝝀))V=b(𝒘,(μ,𝝀))for all (μ,𝝀)V.

With the inner product in V and div𝒯𝝉=div𝝉, 𝒯v=v we have for (μ,𝝀)V that

(𝒗,(μ,𝝀))V=(-div𝝉-𝜷𝝉+γv,-div𝒯𝝀-𝜷𝝀+γμ)
+(𝑪1/2𝝉-𝑪-1/2v,𝑪1/2𝝀-𝑪-1/2𝒯μ)+(𝑪𝝉,𝝀)+(v,μ)
=(g,-div𝒯𝝀-𝜷𝝀+γμ)+(𝑪𝝉,𝝀)+(v,μ)
=b((g,0,0,0),(μ,𝝀))+(𝑪𝝉,𝝀)+(v,μ).

Let (u*,𝝈*)H01(Ω)×𝑯(div;Ω) solve the (primal) problem (1.1) with f=vL2(Ω) and 𝒇=𝝉𝑯(div;Ω). In particular, (u*,𝝈*) solves the ultra-weak formulation (2.2), i.e.,

b((u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*),(μ,𝝀))=(𝑪𝝉,𝝀)+(v,μ)for all (μ,𝝀)V.

Defining 𝒘:=(g,0,0,0)+(u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*) and putting altogether shows

(𝒗,(μ,𝝀))V=b((g,0,0,0),(μ,𝝀))+b((u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*),(μ,𝝀))
=b(𝒘,(μ,𝝀)).

Thus, Θ𝒘=𝒗.

Case (b). The scalar product in this case is given by

((v,𝝉),(μ,𝝀))V=(div𝝉,div𝒯𝝀)+(𝑪𝝉,𝝀)+(𝑪-1v,𝒯μ)+(v,μ).

Recall that 𝜷=0 and note that div𝝉=-g+γv by (3.3). Therefore,

(div𝝉,div𝒯𝝀)=(g-γv,-div𝒯𝝀)=(g-γv,-div𝒯𝝀+γμ)+(γ(γv-g),μ)
=b((g-γv,0,0,0),(μ,𝝀))+(γ(γv-g),μ).

With 𝑪𝝉=v and piecewise integration by parts we obtain

(𝑪-1v,𝒯μ)=(𝝉,𝒯μ)=γ𝒏,𝒮𝝉,μ𝒮+(-div𝝉,μ)
=b((0,0,0,γ𝒏,𝒮𝝉),(μ,𝝀))+(-div𝝉,μ).

Thus,

((v,𝝉),(μ,𝝀))V=(div𝝉,div𝒯𝝀)+(𝑪𝝉,𝝀)+(𝑪-1v,𝒯μ)+(v,μ)
=b((g-γv,0,0,γ𝒏,𝒮𝝉),(μ,𝝀))+(γ(γv-g)-div𝝉+v,μ)+(𝑪𝝉,𝝀).

Defining

𝒘:=(g-γv,0,0,γ𝒏,𝒮𝝉)+(u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*),

where (u*,𝝈*) solves (1.1) with data f=γ(γv-g)-div𝝉+v, 𝒇=𝝉, shows

((v,𝝉),(μ,𝝀))V=b(𝒘,(μ,𝝀))for all (μ,𝝀)V.

Case (c). The proof is similar as for Case (b). Thus, we only give details on the important differences. We have to take care of the terms involving the matrix 𝑪. Note that by the assumptions on 𝑪 it holds 𝑪-1𝝉𝑯(div;Ω) and 𝑪𝝉𝑯(div;Ω) as well. We have

(𝝉,𝝀)=(𝑪𝑪-1𝝉,𝝀),

and using 𝑪𝝉=v and integration by parts,

(v,𝒯μ)=(𝑪𝝉,𝒯μ)=γ𝒏,𝒮(𝑪𝝉),μ𝒮-(div(𝑪𝝉),μ).

Defining

𝒘:=(g-γv,0,0,γ𝒏,𝒮(𝑪𝝉))+(u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*),

where (u*,𝝈*) solves (1.1) with data f=γ(γv-g)-div(𝑪𝝉)+v, 𝒇=𝑪-1𝝉, shows

((v,𝝉),(μ,𝝀))V=(div𝝉,div𝒯𝝀)+(𝝉,𝝀)+(v,𝒯μ)+(v,μ)
=b(𝒘,(μ,𝝀))for all (μ,𝝀)V.

Finally, note that for all three cases it is straightforward to prove

f+𝒇𝑯1(𝒯)g.

Then (2.8) shows estimate (3.4). Moreover, in Case (c) we have 𝝉=𝑪-1v𝑯1(Ω) and 𝒇=𝑪-1𝝉𝑯1(Ω), thus, 𝝈*=𝑪-1𝝉-𝑪-1u*𝑯1(Ω). This finishes the proof. ∎

Remark 9.

As already mentioned in the introduction, in the recent work [15] the optimal test norm

𝒗V,opt:=sup0𝒖Ub(𝒖,𝒗)𝒖U

is considered and the problem of finding 𝒘U such that G(𝒖)=b(𝒖,Θ𝒘), where GU, is analyzed. In view of our previous results we have G(𝒖)=(g,u) and following [15] one would find 𝒘=(g,0,0,0). Since the optimal test norm is not feasible in computations, we only consider the test norms from Lemma 8 in the remainder of the work.

Lemma 10.

Consider one of Cases (a)(c). Let 𝐮=(u,𝛔,u^,σ^)U denote the solution of problem (2.2) and let 𝐮h=(uh,𝛔h,u^h,σ^h)Uh{Uhp,Uhp+} denote the solution of (2.5). Suppose (g,0,0,0)Uh, i.e., gPp(T) if Uh=Uhp resp. gPp+1(T) if Uh=Uhp+. Moreover, suppose that

  1. 𝒫c,01(𝒯)×𝒯p(𝒯)Vhk if Uh=Uhp,

  2. 𝒫c,01(𝒯)×𝒯p+1(𝒯)Vhk if Uh=Uhp+.

It holds that

|(u-uh,g)|Ch𝒖-𝒖hUg.

The constant C>0 only depends on Ω, 𝐂, 𝛃, γ, pN0, and shape-regularity of T.

Proof.

Let 𝒗=(v,𝝉)V denote the solution of the adjoint problem (3.3) with the given gL2(Ω). Let 𝒘=Θ-1𝒗U denote the element from Lemma 8. Since (v,𝝉)H01(Ω)×𝑯(div;Ω), the identities in (2.1) and the adjoint problem (3.3) imply that

(u-uh,g)=b(𝒖-𝒖h,𝒗).

With the bilinear form a(,) of the mixed formulation of DPG (Section 3.2) and the fact that

b(𝒘,δ𝒗)=(𝒗,δ𝒗)V=(δ𝒗,𝒗)Vfor all δ𝒗V,

we infer

(u-uh,g)=b(𝒖-𝒖h,𝒗)=a((𝒖-𝒖h,𝜺-𝜺h),(𝒘,𝒗)).

Here, 𝜺=0 and 𝜺hVhk is the error function which satisfies 𝜺hV𝒖-𝒖hU (see Section 3.2). This, Galerkin orthogonality and boundedness of the bilinear form a(,) show for arbitrary (𝒘h,𝒗h)(Uh,Vhk) that

(u-uh,g)=a((𝒖-𝒖h,𝜺-𝜺h),(𝒘,𝒗))
=a((𝒖-𝒖h,𝜺-𝜺h),(𝒘-𝒘h,𝒗-𝒗h))
𝒖-𝒖hU(𝒘-𝒘hU+𝒗-𝒗hV).

It remains to prove 𝒘-𝒘hU+𝒗-𝒗hVhg. We estimate 𝒗-𝒗hV for all three cases simultaneously and handle the estimation of 𝒘-𝒘hU for the three cases separately, since the representation of 𝒘 by Lemma 8 depends on the choice of norms in V.

We start with the estimation of 𝒗-𝒗hV: We first consider Uh=Uhp. Note that 𝒫c,01(𝒯)×𝒯p(𝒯)Vhk. Choose 𝒗h=(Π1v,Πdivp𝝉)Vhk. Recall that all norms under consideration are equivalent, i.e.,

V,qoptV,1V,2.

Then, using the approximation properties (3.1) together with (3.4), we get

𝒗-𝒗hV𝒗-𝒗hV,2v-Π1vH1(Ω)+𝝉-Πdivp𝝉𝑯(div;Ω)
hg+div(𝝉-Πdivp𝝉).

Then, for the remaining term the commutativity property of the Raviart–Thomas projection, the adjoint problem (3.3) and g𝒫p(𝒯) yield

div(𝝉-Πdivp𝝉)=(1-Πp)div𝝉
=(1-Πp)(-g-𝜷𝝉+γv)
=(1-Πp)(γv-𝜷𝝉).

Using the approximation properties of Π0, γC1(𝒯), 𝜷C1(𝒯)d, and (3.4) shows

(1-Πp)(γv-𝜷𝝉)(1-Π0)(γv-𝜷𝝉)
h𝒯(γv-𝜷𝝉)
hg.

Therefore, we obtain 𝒗-𝒗hVhg. If Uh=Uhp+, then we choose 𝒗h=(Π1,Πdivp+1𝝉)Vhk. With the same lines of proof we also infer 𝒗-𝒗hVhg.

It only remains to estimate 𝒘-𝒘hU. We distinguish between the three different cases.

Case (a). By Lemma 8 we have 𝒘=(g,0,0,0)+~𝒘, where ~𝒘=(u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*). We choose

𝒘h=(g,0,0,0)+~𝒘h,

where ~𝒘hUh0Uh is the best-approximation of (u*,𝝈*,γ0,𝒮u*,γ𝒏,𝒮𝝈*) with respect to U. From Theorem 6 and (3.4) it follows that

𝒘-𝒘hU=~𝒘-~𝒘hUhg.

Case (b). By Lemma 8 we have 𝒘=(g-γv,0,0,γ𝒏,𝒮𝝉)+~𝒘 and choose

𝒘h=(g-Π0γv,0,0,γ𝒏,𝒮Πdiv0𝝉)+~𝒘h,

where ~𝒘hUh0 is the best approximation of ~𝒘 with respect to U. Note that the same arguments as before lead to ~𝒘-~𝒘hUhg. Therefore,

𝒘-𝒘hU(1-Π0)γv+γ𝒏,𝒮(𝝉-Πdiv0𝝉)-1/2,𝒮+~𝒘-~𝒘hUhg,

where we used (3.1) and the approximation property of γ𝒏,𝒮Πdivp in the H-1/2(𝒮) norm (see the proof of Theorem 6) together with (3.4).

Case (c). The proof follows as for Case (b). Therefore, we omit the details. ∎

3.4 Proof of Theorem 3

The best approximation property of Πp and the triangle inequality show that

u-Πpuu-uhu-Πpu+Πp(u-uh).

With g:=Πpu-uh𝒫p(𝒯) observe that

g2=(g,g)=(Πp(u-uh),g)=(u-uh,g).

We apply Lemma 10, and the approximation result from Theorem 6 to see

g2=(u-uh,g)h𝒖-𝒖hUghhp+1(uHp+2(Ω)+𝝈𝑯p+1(𝒯))g.

Dividing by g we infer

u-uhu-Πpu+gu-Πpu+Chp+2(uHp+2(Ω)+𝝈𝑯p+1(𝒯)),

which finishes the proof.

3.5 Proof of Theorem 4

The proof is similar to the one for Theorem 3. We consider

u-uh+u-Πp+1u+Πp+1u-uh+.

Define g:=Πp+1u-uh+𝒫p+1(𝒯). To estimate the second term, we argue as in the proof of Theorem 3 to obtain ghp+2(uHp+2(Ω)+𝝈𝑯p+1(𝒯)). The first term is estimated with the approximation property (3.1a) of the L2 projection, i.e.,

u-Πp+1uhp+2uHp+2(Ω).

This finishes the proof.

3.6 Proof of Theorem 5

Note that (2.9b) is equivalent to Π0u~h=Π0uh. This yields

u-u~h(1-Π0)(u-u~h)+Π0(u-u~h)
h𝒯(u-u~h)+Π0(u-uh),

where we have used the local approximation property of Π0. We define g:=Π0(u-uh). Applying Lemma 10 and Theorem 6 shows

g2=(Π0(u-uh),g)
=(u-uh,g)
h𝒖-𝒖hUg
hp+2(uHp+2(Ω)+𝝈𝑯p+1(𝒯))g.

It remains to estimate 𝒯(u-u~h). The proof follows standard arguments from finite element analysis and is included for completeness. To that end define u¯h𝒫p+1(𝒯) as the solution of the auxiliary Neumann problem

(u¯h,vh)T=(𝑪𝒇-𝑪𝝈+𝜷u,vh)Tfor all vh𝒫p+1(T),
(u¯h,1)T=0

for all T𝒯. Then

𝒯(u¯h-u~h)2=(-𝑪(𝝈-𝝈h)+𝜷(u-uh),𝒯(u¯h-u~h))
𝒖-𝒖hU𝒯(u¯h-u~h)
hp+1(uHp+2(Ω)+𝝈𝑯p+1(𝒯))𝒯(u¯h-u~h).

To estimate 𝒯(u-u¯h), note that there holds Galerkin orthogonality

(𝒯(u-u¯h),𝒯vh)=0for all vh𝒫p+1(𝒯).

Hence, standard approximation results show

𝒯(u-u¯h)=minvh𝒫p+1(𝒯)𝒯(u-vh)hp+1uHp+2(Ω).

Putting altogether gives

u-u~hh𝒯(u-u~h)+g
h(𝒯(u-u¯h)+𝒯(u¯h-u~h))+hp+2(uHp+2(Ω)+𝝈𝑯p+1(𝒯))
hp+2(uHp+2(Ω)+𝝈𝑯p+1(𝒯)),

which finishes the proof.

4 Numerical Studies

In this section we present results of two numerical examples. Let Ω=(0,1)2 be a square domain. Throughout we consider the manufactured solution

u(x,y)=sin(πx)sin(πy),(x,y)Ω,

which is smooth and satisfies u|Γ=0.

Let 𝒖h=(uh,𝝈h,u^h,σ^h)Uhp and 𝒖h+=(uh+,𝝈h+,u^h+,σ^h+)Uhp+ denote the solutions of the practical DPG method (2.5) and let u~h𝒫p+1(𝒯) be the postprocessed solution of 𝒖h, see Section 2.10. We present results for p=0,1,2,3, where we use the test space

Vhk:=𝒫p+2(𝒯)×𝒫p+2(𝒯)d.

To verify our main results (Theorem 3, Theorem 4, and Theorem 5), we check the convergence rates of the L2 errors Πpu-uh, u-uh+, and u-u~h. In all examples below we choose 𝑪 to be the identity matrix. Thus, V,1=V,2 and Cases (b), (c) are identical. The other coefficients are chosen such that the regularity assumptions (2.8) are satisfied.

All computations start with the initial triangulation 𝒯1 visualized in Figure 1.

Figure 1 Initial triangulation 𝒯1{\mathcal{T}_{1}} of domain Ω=(0,1)2{\Omega=(0,1)^{2}}.
Figure 1

Initial triangulation 𝒯1 of domain Ω=(0,1)2.

4.1 Example 1

Define T1:=conv{(0,0),(1,0),(12,12)}, T2:=conv{(1,1),(0,1),(12,12)}. In the first example we set 𝜷=0 and

γ(x,y):={1(x,y)T1,12(x,y)T2,0(x,y)Ω(T1T2).

Moreover, we choose

𝒇(x,y):={(11)x<12,(1-1)x12.

Note that div𝒇=0 and 𝒇𝑯(div;Ω)𝑯1(𝒯1). With the coefficients, 𝒇 and the exact solution at hand, we calculate the right-hand side f and 𝝈 through (1.1). Table 2 resp. Table 3 show errors and convergence rates when using the test norm V,qopt resp. V,1=V,2.

4.2 Example 2

For this example we choose

𝒇=0,γ=0,𝜷(x,y)=(1,1)Tfor (x,y)Ω .

Note that 𝜷 is smooth. Again we calculate f and 𝝈 through (1.1). Table 4 resp. Table 5 show the results for Case (a) (V=V,qopt) resp. Case (b), (c) (V=V,1=V,2). Observe from Table 5 that we do not get higher convergence rates neither for solutions from the augmented space Uhp+ nor for the postprocessed solution. Even for the L2 error of Πpu-uh we do not get higher rates, whereas with the use of the quasi-optimal test norm V,qopt higher rates are obtained. This demonstrates that the assumption 𝜷=0 in Section 2.7 for the Cases (b)–(c) is not an artefact used in the proofs but in general is also necessary to obtain superconvergence results with the norms V,1, V,2.

Table 2

Errors and rates for the problem from Section 4.1 with test norm V=V,qopt.

p#𝒯u-uhrateΠpu-uhrateu-uh+rateu-u~hrate
0161.94e-017.41e-028.37e-021.23e-01
649.37e-021.051.85e-022.002.09e-022.013.21e-021.94
2564.64e-021.014.63e-032.005.20e-032.008.12e-031.98
10242.32e-021.001.16e-032.001.30e-032.002.04e-032.00
40961.16e-021.002.90e-042.003.25e-042.005.09e-042.00
163845.79e-031.007.24e-052.008.13e-052.001.27e-042.00
655362.89e-031.001.81e-052.002.03e-052.003.18e-052.00
1163.47e-023.02e-035.96e-037.89e-03
648.86e-031.975.58e-042.448.72e-042.779.53e-043.05
2562.22e-031.997.92e-052.821.16e-042.921.18e-043.01
10245.56e-042.001.02e-052.951.47e-052.981.48e-053.00
40961.39e-042.001.29e-062.991.84e-062.991.84e-063.00
163843.48e-052.001.62e-073.002.31e-073.002.30e-073.00
2164.51e-032.55e-043.51e-046.14e-04
645.74e-042.981.30e-054.291.98e-054.144.18e-053.88
2567.20e-052.997.67e-074.091.21e-064.042.68e-063.96
10249.01e-063.004.72e-084.027.50e-084.011.68e-073.99
40961.13e-063.002.97e-093.994.69e-094.001.05e-084.00
3162.20e-042.08e-052.01e-055.48e-05
641.39e-053.988.34e-074.648.38e-074.581.67e-065.03
2568.70e-074.002.82e-084.892.86e-084.875.18e-085.01
10245.44e-084.009.08e-104.969.24e-104.951.62e-095.00
Table 3

Errors and rates for the problem from Section 4.1 with test norm V=V,2.

p#𝒯u-uhrateΠpu-uhrateu-uh+rateu-u~hrate
0161.92e-016.88e-027.86e-028.48e-02
649.35e-021.041.73e-021.991.97e-021.992.17e-021.97
2564.64e-021.014.33e-032.004.94e-032.005.44e-031.99
10242.32e-021.001.08e-032.001.23e-032.001.36e-032.00
40961.16e-021.002.71e-042.003.09e-042.003.41e-042.00
163845.79e-031.006.77e-052.007.71e-052.008.51e-052.00
655362.89e-031.001.69e-052.001.93e-052.002.13e-052.00
1163.49e-024.81e-036.96e-036.79e-03
648.87e-031.987.36e-042.719.71e-042.848.82e-042.95
2562.22e-032.009.82e-052.911.26e-042.951.12e-042.98
10245.56e-042.001.25e-052.971.59e-052.991.41e-052.99
40961.39e-042.001.57e-062.991.99e-063.001.76e-063.00
163843.48e-052.001.96e-073.002.49e-073.002.20e-073.00
2164.53e-034.38e-045.07e-045.22e-04
645.74e-042.982.53e-054.113.00e-054.083.25e-054.01
2567.20e-052.991.54e-064.041.85e-064.022.03e-064.00
10249.01e-063.009.58e-084.011.15e-074.011.27e-074.00
40961.13e-063.006.03e-093.997.22e-093.997.94e-094.00
3162.25e-045.14e-055.06e-056.01e-05
641.40e-054.011.75e-064.881.73e-064.871.96e-064.94
2568.71e-074.005.62e-084.965.55e-084.966.20e-084.98
10245.44e-084.001.80e-094.961.78e-094.961.96e-094.98
Table 4

Errors and rates for the problem from Section 4.2 with test norm V=V,qopt.

p#𝒯u-uhrateΠpu-uhrateu-uh+rateu-u~hrate
0161.96e-017.95e-028.85e-021.27e-01
649.41e-021.062.04e-021.962.25e-021.983.36e-021.92
2564.65e-021.025.14e-031.995.64e-031.998.51e-031.98
10242.32e-021.001.29e-032.001.41e-032.002.13e-031.99
40961.16e-021.003.22e-042.003.53e-042.005.34e-042.00
163845.79e-031.008.05e-052.008.82e-052.001.34e-042.00
655362.89e-031.002.01e-052.002.21e-052.003.34e-052.00
1163.47e-022.77e-035.91e-038.02e-03
648.85e-031.975.22e-042.408.59e-042.789.73e-043.04
2562.22e-031.997.47e-052.801.14e-042.921.21e-043.01
10245.56e-042.009.69e-062.951.44e-052.981.51e-053.00
40961.39e-042.001.22e-062.991.81e-062.991.89e-063.00
163843.48e-052.001.53e-073.002.27e-073.002.36e-073.00
2164.51e-032.37e-043.44e-046.25e-04
645.73e-042.981.19e-054.321.95e-054.144.24e-053.88
2567.20e-052.996.97e-074.091.19e-064.042.72e-063.97
10249.01e-063.004.28e-084.027.37e-084.011.71e-073.99
40961.13e-063.002.68e-094.004.60e-094.001.07e-084.00
3162.20e-041.95e-051.98e-055.51e-05
641.39e-053.987.80e-074.648.14e-074.611.68e-065.04
2568.70e-074.002.65e-084.882.78e-084.875.21e-085.01
10245.44e-084.008.73e-104.929.16e-104.921.63e-095.00
Table 5

Errors and rates for the problem from Section 4.2 with test norm V=V,2.

p#𝒯u-uhrateΠpu-uhrateu-uh+rateu-u~hrate
0164.37e-013.98e-014.15e-014.00e-01
642.25e-010.962.06e-010.952.14e-010.962.06e-010.96
2561.14e-010.991.04e-010.981.08e-010.991.04e-010.99
10245.70e-021.005.21e-021.005.41e-021.005.21e-021.00
40962.85e-021.002.61e-021.002.71e-021.002.61e-021.00
163841.43e-021.001.30e-021.001.35e-021.001.30e-021.00
655367.13e-031.006.52e-031.006.77e-031.006.52e-031.00
1166.23e-025.18e-025.69e-021.64e-02
641.63e-021.931.37e-021.921.49e-021.935.70e-031.52
2564.12e-031.983.47e-031.983.77e-031.981.61e-031.83
10241.03e-032.008.67e-042.009.42e-042.004.19e-041.94
40962.57e-042.002.17e-042.002.35e-042.001.06e-041.98
163846.43e-052.005.41e-052.005.88e-052.002.68e-051.99
2167.46e-035.95e-036.56e-039.34e-04
649.32e-043.007.35e-043.028.17e-043.006.23e-053.91
2561.17e-043.009.17e-053.001.02e-043.004.01e-063.96
10241.46e-053.001.15e-053.001.28e-053.002.57e-073.96
40961.82e-063.001.43e-063.001.59e-063.001.66e-083.95
3166.03e-045.62e-045.59e-047.46e-05
643.88e-053.963.63e-053.953.64e-053.943.93e-064.25
2562.44e-063.992.28e-063.992.29e-063.992.41e-074.03
10241.52e-074.001.42e-074.001.43e-074.001.52e-083.99

5 Concluding Remarks

We conclude this work with some remarks. The results and their proofs are presented in a systematic way that allow to extend and transfer them to other types of meshes and different model problems. In principle, the crucial results Lemma 8 and Lemma 10 have to be verified. Consider for instance that 𝒯 is a mesh with polygonal elements. Lemma 8 still holds true in that case since it is independent of the underlying mesh so that only the assertion of Lemma 10 has to be shown. To be more precise: Analyzing the proof one finds out that it only remains to provide the estimate

min𝒘hUh𝒘-𝒘hU+min𝒗kVhk𝒗-𝒗kVhg,

which is an optimal a priori error bound for sufficient regular functions (see Lemma 10 for details on the definition of the functions 𝒘 and 𝒗). In the case of triangular meshes we have proven the estimate by using basic properties of well-known interpolation operators. If operators with the same properties can be defined on meshes with polygonal elements, then, clearly, the estimate holds true as well. We note that the analysis of DPG methods for ultra-weak formulations on general (polygonal) meshes is an ongoing research. For an overview we refer to the recent work [21].

Future research will include other model problems, e.g., linear elasticity. Another possible application of the developed ideas could be to the Stokes problem. Consider its velocity-gradient-pressure formulation: Find (𝒖S,𝝈S,pS) such that

-pS+div𝝈S=𝒇in Ω,
𝝈S-𝒖S=0in Ω,
div𝒖S=0in Ω,
𝒖S=0on Ω.

DPG methods based on ultra-weak formulations are known and thoroughly analyzed [17]. Since regularity theory is also known, our main results (Theorem 35) should carry over (for the velocity variable 𝒖S instead of u) to the Stokes problem following the same lines in the proofs. In particular, the assertion of Theorem 3 has been already observed in numerical experiments [17, Section 3] even for different test norms. We refer also to [18, Section 3] for numerical evidence in the case of incompressible Navier Stokes problems.

Another point we like to mention is that the principal ideas of the proofs and, thus, our main results carry over to the low regularity case, i.e., when we do not have the “full” regularity uH2(Ω), vH2(Ω) for solutions of (1.1) and (2.7) but rather uH1+s(Ω), vH1+s(Ω) for some s(12,1). This is usually the case when Ω is a nonconvex polygonal domain. Nevertheless, we stress that our main results (Theorem 35) hold true with hp+2 replaced by hp+1+s. Therefore, one still obtains higher convergence rates than the overall error 𝒖-𝒖h=𝒪(hp+1). For the particular case of a reaction-diffusion model problem (𝑪 is the identity, 𝜷=0, and γ=1) Theorem 4 and 5 are analyzed in [10] for V=V,1=V,2.

Finally, let us remark the importance of the choice of norms in the test space. Although all test norms under consideration are equivalent and, thus, the corresponding DPG methods have the same stability properties (i.e., the infsup constants resp. boundedness constants are equivalent), only one of the norms under consideration (the quasi-optimal norm V,qopt) yields higher convergence rates for general model problems with 𝜷0. This has to be taken into account in the design of DPG methods.

Award Identifier / Grant number: 11170050

Funding statement: This work was supported by FONDECYT project 11170050.

References

[1] T. Bouma, J. Gopalakrishnan and A. Harb, Convergence rates of the DPG method with reduced test space degree, Comput. Math. Appl. 68 (2014), no. 11, 1550–1561. 10.1016/j.camwa.2014.08.004Search in Google Scholar

[2] C. Carstensen, L. Demkowicz and J. Gopalakrishnan, A posteriori error control for DPG methods, SIAM J. Numer. Anal. 52 (2014), no. 3, 1335–1353. 10.1137/130924913Search in Google Scholar

[3] C. Carstensen, L. Demkowicz and J. Gopalakrishnan, Breaking spaces and forms for the DPG method and applications including Maxwell equations, Comput. Math. Appl. 72 (2016), no. 3, 494–522. 10.1016/j.camwa.2016.05.004Search in Google Scholar

[4] B. Cockburn, B. Dong and J. Guzmán, A superconvergent LDG-hybridizable Galerkin method for second-order elliptic problems, Math. Comp. 77 (2008), no. 264, 1887–1916. 10.1090/S0025-5718-08-02123-6Search in Google Scholar

[5] L. Demkowicz and J. Gopalakrishnan, A class of discontinuous Petrov–Galerkin methods. Part I: The transport equation, Comput. Methods Appl. Mech. Engrg. 199 (2010), no. 23–24, 1558–1572. 10.1016/j.cma.2010.01.003Search in Google Scholar

[6] L. Demkowicz and J. Gopalakrishnan, A class of discontinuous Petrov–Galerkin methods. II. Optimal test functions, Numer. Methods Partial Differential Equations 27 (2011), no. 1, 70–105. 10.1002/num.20640Search in Google Scholar

[7] L. Demkowicz and J. Gopalakrishnan, Analysis of the DPG method for the Poisson equation, SIAM J. Numer. Anal. 49 (2011), no. 5, 1788–1809. 10.1137/100809799Search in Google Scholar

[8] L. Demkowicz, J. Gopalakrishnan and A. H. Niemi, A class of discontinuous Petrov–Galerkin methods. Part III: Adaptivity, Appl. Numer. Math. 62 (2012), no. 4, 396–427. 10.1016/j.apnum.2011.09.002Search in Google Scholar

[9] A. Demlow, Suboptimal and optimal convergence in mixed finite element methods, SIAM J. Numer. Anal. 39 (2002), no. 6, 1938–1953. 10.1137/S0036142900376900Search in Google Scholar

[10] T. Führer, Superconvergence in a DPG method for an ultra-weak formulation, Comput. Math. Appl. 75 (2018), no. 5, 1705–1718. 10.1016/j.camwa.2017.11.029Search in Google Scholar

[11] L. Gastaldi and R. H. Nochetto, Sharp maximum norm error estimates for general mixed finite element approximations to second order elliptic equations, RAIRO Modél. Math. Anal. Numér. 23 (1989), no. 1, 103–128. 10.1051/m2an/1989230101031Search in Google Scholar

[12] J. Gopalakrishnan and W. Qiu, An analysis of the practical DPG method, Math. Comp. 83 (2014), no. 286, 537–552. 10.1090/S0025-5718-2013-02721-4Search in Google Scholar

[13] P. Grisvard, Elliptic Problems in Nonsmooth Domains, Monogr. Stud. Math. 24, Pitman, Boston, 1985. Search in Google Scholar

[14] B. Keith, L. Demkowicz and J. Gopalakrishnan, DPG* method, preprint (2017), https://arxiv.org/abs/1710.05223. Search in Google Scholar

[15] B. Keith, A. Vaziri Astaneh and L. Demkowicz, Goal-oriented adaptive mesh refinement for non-symmetric functional settings, preprint (2017), https://arxiv.org/abs/1711.01996. Search in Google Scholar

[16] S. Nagaraj, S. Petrides and L. F. Demkowicz, Construction of DPG Fortin operators for second order problems, Comput. Math. Appl. 74 (2017), no. 8, 1964–1980. 10.1016/j.camwa.2017.05.030Search in Google Scholar

[17] N. V. Roberts, T. Bui-Thanh and L. Demkowicz, The DPG method for the Stokes problem, Comput. Math. Appl. 67 (2014), no. 4, 966–995. 10.1016/j.camwa.2013.12.015Search in Google Scholar

[18] N. V. Roberts, L. Demkowicz and R. Moser, A discontinuous Petrov–Galerkin methodology for adaptive solutions to the incompressible Navier–Stokes equations, J. Comput. Phys. 301 (2015), 456–483. 10.1016/j.jcp.2015.07.014Search in Google Scholar

[19] L. R. Scott and S. Zhang, Finite element interpolation of nonsmooth functions satisfying boundary conditions, Math. Comp. 54 (1990), no. 190, 483–493. 10.1090/S0025-5718-1990-1011446-7Search in Google Scholar

[20] R. Stenberg, Postprocessing schemes for some mixed finite elements, RAIRO Modél. Math. Anal. Numér. 25 (1991), no. 1, 151–167. 10.1051/m2an/1991250101511Search in Google Scholar

[21] A. Vaziri Astaneh, F. Fuentes, J. Mora and L. Demkowicz, High-order polygonal discontinuous Petrov–Galerkin (PolyDPG) methods using ultraweak formulations, Comput. Methods Appl. Mech. Engrg. 332 (2018), 686–711. 10.1016/j.cma.2017.12.011Search in Google Scholar

[22] J. Zitelli, I. Muga, L. Demkowicz, J. Gopalakrishnan, D. Pardo and V. M. Calo, A class of discontinuous Petrov–Galerkin methods. Part IV: the optimal test norm and time-harmonic wave propagation in 1D, J. Comput. Phys. 230 (2011), no. 7, 2406–2432. 10.1016/j.jcp.2010.12.001Search in Google Scholar

Received: 2018-09-25
Revised: 2019-01-10
Accepted: 2019-03-05
Published Online: 2019-04-06
Published in Print: 2019-07-01

© 2019 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 24.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/cmam-2018-0250/html
Scroll to top button