Discontinuous Petrov-Galerkin Method
Discontinuous Petrov-Galerkin Method
293–384
doi:10.1017/S0962492924000102
Jay Gopalakrishnan
PO Box 751 (MTH), Portland State University,
Portland, OR 97207-0751, USA
E-mail: [email protected]
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
294 L. Demkowicz and J. Gopalakrishnan
CONTENTS
1 Introduction 294
2 Optimal test spaces 297
3 Minimization and other viewpoints 300
4 Ideal DPG methods 305
5 Practical DPG methods 316
6 A posteriori error control 334
7 Ultraweak formulations 338
8 Optimal test functions in time integrators 356
9 Duality in DPG formulations 364
10 Pointers to DPG techniques for nonlinear problems 371
11 Further pointers and conclusion 376
References 378
1. Introduction
In variational methods, approximate solutions are sought in ‘trial’ spaces, while
equations are enforced using ‘test’ spaces. Methods with different trial and test
spaces are referred to as Petrov–Galerkin (PG) formulations. A classical result
on such methods, restated below in Theorem 1.1, provides the following useful
insight for designing Petrov–Galerkin methods: while one must choose trial spaces
with good approximation properties, test spaces may be chosen solely for stabil-
ity. Leveraging this insight, discontinuous Petrov–Galerkin (DPG) methods were
originally conceived (Demkowicz and Gopalakrishnan 2010, 2011b) as Petrov–
Galerkin methods that obtain stability automatically by local test space design
using discontinuous functions. The goal of this review is to provide an introduc-
tion to these methods and present selected recent advances. We review established
DPG techniques, give a few new avenues to existing results, and also present a few
new results. We describe the mathematical foundations for the popular features that
make the DPG method a powerful tool for solving boundary value problems, in-
cluding the ease with which automatic adaptivity can be enabled and stable solvers
for complex problems can be built.
Let us begin by describing a standard difficulty in PG formulations of boundary
value problems that the DPG method addresses. The ‘wellposedness’ (or continu-
ous dependence of solutions on data) of PG formulations need not automatically
yield stability of their discretizations, unlike coercive Galerkin formulations. In
simpler formulations with equal trial and test spaces, once wellposedness of the
variational formulation (usually set in infinite-dimensional Sobolev spaces) is es-
tablished through coercivity, stability of the computable Galerkin discretization
(using finite-dimensional subspaces) follows. The situation for non-coercive PG
methods is more complicated.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 295
By the well-known theory of mixed systems (Babuška 1971, Brezzi 1974, Ern and
Guermond 2021, Nečas 1962), we know that (1.1) holds if and only if there is a
𝛾 > 0 such that
|𝑏(𝑧, 𝑦)|
inf sup ≥ 𝛾, and (1.2a)
0≠𝑧 ∈𝑋 0≠𝑦 ∈𝑌 ∥𝑧∥ 𝑋 ∥𝑦∥𝑌
{𝑦 ∈ 𝑌 : 𝑏(𝑧, 𝑦) = 0 for all 𝑧 ∈ 𝑋 } = {0}, (1.2b)
or equivalently
|𝑏(𝑧, 𝑦)|
inf sup ≥ 𝛾, and (1.3a)
0≠𝑦 ∈𝑌 0≠𝑧 ∈𝑋 ∥𝑦∥𝑌 ∥𝑧∥ 𝑋
{𝑧 ∈ 𝑋 : 𝑏(𝑧, 𝑦) = 0 for all 𝑦 ∈ 𝑌 } = {0}. (1.3b)
a classical result of Babuška (see Babuška 1971, Babuška, Aziz, Fix and Kellogg
1972 or Xu and Zikatanov 2003) can be stated as follows.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
296 L. Demkowicz and J. Gopalakrishnan
and
∥𝑏∥
inf ∥𝑥 − 𝑧 ℎ ∥ 𝑋 .
∥𝑥 − 𝑥 ℎ ∥ 𝑋 ≤ (1.6)
𝛾 ℎ 𝑧ℎ ∈𝑋ℎ
The theorem clarifies the above-mentioned difficulty in inheriting discrete sta-
bility from the wellposedness of the variational problem. Specifically, the standard
difficulty in the analysis of Petrov–Galerkin discretizations of the form (1.5) is that
the inf-sup condition (1.2a) does not generally imply the discrete inf-sup condition
(1.4). Hence, unlike coercive forms 𝑏(·, ·), it is easy to obtain unstable PG meth-
ods even when the variational equation (1.1) is wellposed. A second important
observation is that in Theorem 1.1, the test space 𝑌ℎ is absent from the error es-
timate (1.6). It only appears in the inf-sup condition (1.4), which is responsible for
stability. The approximation rates are determined by the trial space 𝑋ℎ in (1.6).
Letting the trial space carry the burden of achieving good approximation properties
liberates the test space from it. The takeaway is that we can focus solely on stability
when designing test spaces. Hence techniques to design discrete subspaces 𝑌ℎ that
guarantee the discrete inf-sup condition (1.4), with 𝛾 ℎ independent of the finite
dimension, are useful.
The next section (Section 2) provides such a technique through the concept
of optimal test functions which attain the supremum in an inf-sup condition. In
Section 3 we shall see that DPG methods can equivalently be thought of as methods
that minimize a residual in a non-standard norm, as well as a non-standard mixed
method with an approximate error representation as a discrete solution component.
One of the key steps that enable local computation of the optimal test functions
is a reformulation of the boundary value problem using a test space of functions
which have no continuity constraints across elements. Such spaces are often
referred to as ‘broken’ spaces and the process is akin to ‘hybridization’. We will
see these terms again when the test space localization technique is introduced
precisely in Section 4. Next, we address the usual practice of computing the
optimal test functions inexactly. The loss of optimality in the stability constant and
techniques to regain discrete stability despite the inexact computation are topics
covered in Section 5. The key ingredient there is a local Fortin operator. The built-
in error estimator contained in all DPG methods is then described in Section 6.
The wide scope of applicability of DPG methods becomes clearer in Section 7,
where we show how to accomplish the above-mentioned localization through a
reformulation in broken graph spaces of very general partial differential equations
(PDE). Application of optimal test functions to create enhanced time integrators is
the subject of Section 8. Duality arguments, certain drawbacks in applying them
to DPG methods, a dual DPG* method, and application to estimating error in goal
functionals are briefly discussed in Section 9. Section 10 contains a collection of
remarks on ongoing efforts to incorporate DPG techniques into nonlinear boundary
value problems, a research area where the last word seems yet to be written. We
conclude in Section 11 with pointers to further works whose details could not be
included in this paper.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 297
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
298 L. Demkowicz and J. Gopalakrishnan
opt
where 𝑌ℎ ⊂ 𝑌 is computed using the 𝑌 -inner product by (2.1)–(2.2).
Theorem 2.4 (Quasioptimality). If (1.3) holds, then the IPG method (2.3) is
uniquely solvable for 𝑥 ℎ and
∥𝑏∥
∥𝑥 − 𝑥 ℎ ∥ 𝑋 ≤ inf ∥𝑥 − 𝑧 ℎ ∥ 𝑋 , (2.4)
𝛾 𝑧ℎ ∈𝑋ℎ
where 𝑥 is the unique exact solution of (1.1).
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 299
This implies that letting 𝑃 𝑝 (𝛺) denote the space of polynomials of degree at most
𝑝, restricted to 𝛺, and setting the discrete trial space to 𝑋ℎ = 𝑃 𝑝 (𝛺) × R, we have
opt
𝑌ℎ = 𝑃 𝑝+1 (𝛺).
The solution 𝑥 ℎ = (𝑢 ℎ , 𝑢ˆ 1,ℎ ) of the resulting IPG method, in view of Theorem 2.4
and (2.9), is interesting in that 𝑢 ℎ equals the 𝐿 2 (𝛺)-projection of 𝑢 onto 𝑃 𝑝 (𝛺).
In the general case, although one cannot expect the method to deliver the best
𝐿 2 -approximation from the trial space, the solution is the best approximation in
some norm, as will be proved in the next section.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
300 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 301
Proof. (a) ⇐⇒ (b) By definition of the IPG method, 𝑥 ℎ solves (2.3) if and only
opt
if 𝑏(𝑥 − 𝑥 ℎ , 𝑦 ℎ ) = 0 for all 𝑦 ℎ ∈ 𝑌ℎ . By the definition of the optimal test space,
this is equivalent to
𝑏(𝑥 − 𝑥 ℎ , 𝑇 𝑧 ℎ ) = 0 for all 𝑧 ℎ ∈ 𝑋ℎ ,
which, in turn, is equivalent to
(𝑇(𝑥 − 𝑥 ℎ ), 𝑇 𝑧 ℎ )𝑌 = 0 for all 𝑧 ℎ ∈ 𝑋ℎ ,
due to (2.2). The result follows since (𝑇 ·, 𝑇 ·)𝑌 is the inner product generating the
|||·||| 𝑋 -norm.
(b) ⇐⇒ (c) In view of (3.2),
|||𝑥 − 𝑧 ℎ ||| 𝑋 = ∥𝑇(𝑥 − 𝑧 ℎ )∥𝑌 = ∥𝑅𝑌−1 𝐵(𝑥 − 𝑧 ℎ )∥𝑌 .
Hence, by the isometry of the Riesz map (3.1), item (b) holds if and only if
∥𝐵(𝑥 − 𝑥 ℎ )∥𝑌 ∗ = inf ∥𝐵(𝑥 − 𝑧 ℎ )∥𝑌 ∗ ,
𝑧ℎ ∈𝑋ℎ
Proof. (a) =⇒ (b) Equation (3.4a) is the same as (3.3), so we only need to prove
(3.4b). To this end,
𝑏(𝑧 ℎ , 𝜀) = (𝑇 𝑧 ℎ , 𝜀)𝑌 = (𝑇 𝑧 ℎ , 𝑅𝑌−1 (ℓ − 𝐵𝑥 ℎ ))𝑌 = (𝑇 𝑧 ℎ , 𝑇(𝑥 − 𝑥 ℎ ))𝑌 ,
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
302 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 303
Let 𝐵 : 𝑋 → 𝑌 ∗ denote the operator generated by 𝑏(·, ·) by ⟨𝐵𝑧, 𝑦⟩𝑌 = 𝑏(𝑧, 𝑦) for
all 𝑧 ∈ 𝑋, 𝑦 ∈ 𝑌 . Identifying the bidual 𝑌 ∗∗ with 𝑌 , the adjoint 𝐵∗ : 𝑌 → 𝑋 ∗ of 𝐵
satisfies
(𝐵∗ 𝑦)(𝑧) = 𝑏(𝑧, 𝑦) for all 𝑧 ∈ 𝑋, 𝑦 ∈ 𝑌 . (3.9)
Using the Riesz maps in 𝑋 and 𝑌 , we then immediately have
𝑏(𝑧, 𝑦) = 𝑅𝑌−1 𝐵𝑧, 𝑦 𝑌 = 𝑧, 𝑅 𝑋−1 𝐵∗ 𝑦 𝑋 , 𝑧 ∈ 𝑋, 𝑦 ∈ 𝑌 .
(3.10)
This readily implies the twin identities
|||𝑧||| 𝑋 = ∥𝑅𝑌−1 𝐵𝑧∥𝑌 , ||||𝑦||||𝑌 = ∥𝑅 𝑋−1 𝐵∗ 𝑦∥ 𝑋 (3.11)
for all 𝑧 ∈ 𝑋 and 𝑦 ∈ 𝑌 , by the definitions of energy norm and optimal test norm.
Now we show that one may equivalently shorten the definition of the generalized
duality pairing by omitting one of the two equalities in (3.7).
Proposition 3.6. The identity |||𝑧||| 𝑋 = ∥𝑧∥ 𝑋 holds for all 𝑧 ∈ 𝑋 if and only if
||||𝑦||||𝑌 = ∥𝑦∥𝑌 for all 𝑦 ∈ 𝑌 . Therefore, whenever either equality holds, we have
∥𝑏∥ = 1 and 𝛾 = 1.
Proof. If |||𝑧||| 𝑋 = ∥𝑧∥ 𝑋 for all 𝑧 ∈ 𝑋, then (3.6) and (3.11) imply
The last supremum equals ∥𝑦∥𝑌 since 𝑅𝑌−1 𝐵 : 𝑋 → 𝑌 is a bijection. The converse
is proved similarly using the other identity in (3.11). The last assertion on ∥𝑏∥ and
𝛾 immediately follows from (3.8).
Clearly one direction of Proposition 3.6 answers the question posed at the begin-
ning of this subsection. If we use the optimal test norm for 𝑌 , then the energy norm
coincides with the given ∥ · ∥ 𝑋 -norm, and Theorem 3.2(b) shows that the solution
of the IPG method is guaranteed to be the best approximation in the given 𝑋-
norm. However, as we shall see later, the optimal test norm is often not practically
computable easily in the multi-dimensional examples we have in mind.
Next, we contrast the previously introduced trial-to-test operator which produces
optimal test functions with an earlier trial-to-test operator given in Barrett and
Morton (1984). To this end, let us introduce an adaptation of their ideas to our
current Petrov–Galerkin setting. (They used equal trial and test spaces.) We define
the ‘Barrett–Morton trial-to-test operator’ 𝑇 BM : 𝑋 → 𝑌 by
𝑏(𝑤, 𝑇 BM 𝑧) = (𝑤, 𝑧)𝑋 for all 𝑤, 𝑧 ∈ 𝑋. (3.12)
Using the inverse of 𝐵∗ , an equivalent characterization of 𝑇 BM is
𝑇 BM = (𝐵∗ )−1 ◦ 𝑅 𝑋 .
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
304 L. Demkowicz and J. Gopalakrishnan
Comparing (3.12) with (2.2), we find two different trial-to-test mappings. The
difference between our 𝑇 = 𝑅𝑌−1 ◦ 𝐵 (see (3.2)) and 𝑇 BM is illustrated in the
following diagram:
𝑌∗
𝐵
𝑋
𝑅𝑋 𝑅𝑌
𝑋∗ 𝑌
𝐵∗
This implies the remarkable property that the solution 𝑥 BM ℎ ∈ 𝑋ℎ of the method
(3.13) equals the 𝑋-orthogonal projection of the exact solution 𝑥 and explains the
potential interest in the method (3.13). However, inverting 𝐵∗ to compute test
space basis functions is generally too expensive. In contrast, we will show in later
sections that the inversion of 𝑅𝑌 to compute 𝑇 can be realized locally if the problem
is reformulated adequately.
Nevertheless, at this point it is useful to note one scenario where 𝑇 and 𝑇 BM
coincide. This occurs when 𝑏 is a generalized duality pairing.
Proposition 3.7. If ∥𝑧∥ 𝑋 = |||𝑧||| 𝑋 for all 𝑧 ∈ 𝑋, then 𝑇 = 𝑇 BM .
Proof. By (3.11), |||𝑧||| 𝑋 = ∥𝑅𝑌−1 𝐵𝑧∥𝑌 = ∥𝑇 𝑧∥𝑌 . Hence, whenever ∥𝑧∥ 𝑋 = |||𝑧||| 𝑋
for all 𝑧 ∈ 𝑋, by polarization, we have
(𝑤, 𝑧)𝑋 = (𝑇 𝑤, 𝑇 𝑧)𝑌 = 𝑏(𝑤, 𝑇 𝑧) for all 𝑧, 𝑤 ∈ 𝑋,
where we have used (2.2) in the last equality. Comparing this with (3.12), we find
that 𝑏(𝑤, 𝑇 𝑧) = 𝑏(𝑤, 𝑇 BM 𝑧) for all 𝑤, 𝑧 ∈ 𝑋. Hence 𝑇 = 𝑇 BM by (1.2b).
Thus, when 𝑏 is a generalized duality pairing, the IPG method coincides with
the method (3.13) and the discrete solution equals the 𝑋-orthogonal projection of
the exact solution.
Example 3.8. It can be easily seen that the bilinear form 𝑏 in (2.8) of Example 2.6
is a generalized duality pairing. Hence the analytically solved expression for 𝑇,
given there in (2.10), coincides with 𝑇 BM .
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 305
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
306 L. Demkowicz and J. Gopalakrishnan
Our interest in using such a product space for the test variable is the resulting
localization of the trial-to-test operator 𝑇. Note that to compute a basis for the
optimal test space, we must solve (2.2) to compute 𝑇 𝑧 for each 𝑧 in a basis of 𝑋ℎ .
That equation, (𝑇 𝑧, 𝑦)𝑌 = 𝑏(𝑧, 𝑦), decouples into independent equations on each
element, if 𝑌 has the form (4.1). Localization of 𝑇 refers to the fact that the part
of 𝑇 𝑧 on an element 𝐾, namely (𝑇 𝑧)𝐾 , can be computed, independently of other
elements, by solving
((𝑇 𝑧)𝐾 , 𝑦 𝐾 )𝑌 (𝐾) = 𝑏(𝑧, 𝑦 𝐾 ) for all 𝑦 𝐾 ∈ 𝑌 (𝐾). (4.3)
The adjective discontinuous in the name ‘DPG’ refers to the fact that test functions
in 𝑌 of the form (4.1) admit discontinuous functions with no continuity constraints
across element interfaces. For example, in many applications, we set 𝑌 to
𝐻 1 (𝛺 ℎ ) ≔ {𝑣 ∈ 𝐿 2 (𝛺) : 𝑣| 𝐾 ∈ 𝐻 1 (𝐾) for all 𝐾 ∈ 𝛺 ℎ },
which can be identified with the Cartesian product
Ö
𝐻 1 (𝛺 ℎ ) ≡ 𝐻 1 (𝐾), (4.4)
𝐾 ∈𝛺ℎ
and contains functions that are discontinuous across element interfaces. Collo-
quially, we say that 𝐻 1 (𝛺 ℎ ) is a broken Sobolev space, obtained by breaking the
inter-element continuity constraints of 𝐻 1 (𝛺). DPG methods are built using broken
Sobolev spaces as test spaces.
Example 4.2 (Laplace equation: primal DPG formulation). Let 𝑓 ∈ 𝐿 2 (𝛺) and
𝑢 satisfy
−Δ𝑢 = 𝑓 in 𝛺, (4.5a)
𝑢=0 on 𝜕𝛺. (4.5b)
The standard variational formulation for this problem finds 𝑢 in 𝐻˚ 1 (𝛺) such that
(grad 𝑢, grad 𝑣)𝛺 = ( 𝑓 , 𝑣)𝛺 for all 𝑣 ∈ 𝐻˚ 1 (𝛺). (4.6)
A different variational formulation is obtained if we multiply (4.5a) by a possibly
discontinuous test function 𝑦 ∈ 𝐻 1 (𝛺 ℎ ) (defined in (4.4)) and integrate by parts,
element by element. On a single element 𝐾 ∈ 𝛺 ℎ , we have
∫ ∫ ∫
grad 𝑢 · grad 𝑦 − (𝑛 · grad 𝑢)𝑦 = 𝑓 𝑦. (4.7)
𝐾 𝜕𝐾 𝐾
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 307
where (·, ·)𝐷 , for any domain 𝐷, denotes the 𝐿 2 (𝐷)-inner product and ⟨ℓ, ·⟩ 𝐻 1/2 (𝜕𝐾)
denotes the action of a conjugate linear functional ℓ ∈ 𝐻 −1/2 (𝜕𝐾) on a function in
𝐻 1/2 (𝜕𝐾). Define the element-by-element trace operator
Ö
tr𝑛 : 𝐻(div, 𝛺) → 𝐻 −1/2 (𝜕𝐾), tr𝑛 𝑟 | 𝜕𝐾 = 𝑟 · 𝑛| 𝜕𝐾 . (4.9)
𝐾 ∈𝛺ℎ
Here and throughout, 𝑛 denotes the unit outward normal vector of a domain under
consideration, which is usually clear from the context, e.g. above 𝑛 is the outward
unit normal on each element boundary 𝜕𝐾. (On an interior interface shared by two
elements 𝐾± , the 𝑛 from 𝐾± will have opposite signs.) We endow the image of the
trace map with a quotient norm,
𝐻 −1/2 (𝜕𝛺 ℎ ) = range(tr𝑛 ),
∥ 𝑟ˆ𝑛 ∥ 𝐻 −1/2 (𝜕𝛺ℎ ) = inf ∥𝑞∥ 𝐻(div,𝛺) , (4.10)
𝑞 ∈tr𝑛−1 { 𝑟ˆ𝑛 }
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
308 L. Demkowicz and J. Gopalakrishnan
hybridization; see e.g. Raviart and Thomas (1977b), Brezzi and Fortin (1991) or
Cockburn and Gopalakrishnan (2004). We have used the adjective ‘hybrid’ in the
above example following this tradition. To analyse hybrid formulations like those
ˆ and 𝑌 denote
of Example 4.2, we formulate a result in a general setting. Let 𝑋0 , 𝑋,
Hilbert spaces over C, put 𝑋 = 𝑋0 × 𝑋, and let 𝑏 0 : 𝑋0 ×𝑌 → C and 𝑏ˆ : 𝑋ˆ ×𝑌 → C
ˆ
denote continuous sesquilinear forms. Then
𝑌0 = {𝑦 ∈ 𝑌 : 𝑏( ˆ 𝑦) = 0 for all 𝑥ˆ ∈ 𝑋ˆ }
ˆ 𝑥, (4.12a)
is a closed subspace of 𝑌 . Suppose there are positive constants 𝛾0 and 𝛾ˆ such that
|𝑏 0 (𝑥, 𝑦)|
𝛾0 ∥𝑥∥ 𝑋0 ≤ sup for all 𝑥 ∈ 𝑋0 , and (4.12b)
0≠𝑦 ∈𝑌0 ∥𝑦∥𝑌
| 𝑏(
ˆ 𝑥,
ˆ 𝑦)|
𝛾ˆ ∥ 𝑥∥
ˆ 𝑋ˆ ≤ sup for all 𝑥ˆ ∈ 𝑋.
ˆ (4.12c)
0≠𝑦 ∈𝑌 ∥𝑦∥ 𝑌
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 309
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
310 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 311
Having verified the assumptions, Theorem 4.3 now gives the inf-sup condition
(grad 𝑢, grad 𝑦)ℎ + ⟨𝑞ˆ 𝑛 , 𝑦⟩ℎ
∥(𝑢, 𝑞ˆ 𝑛 )∥ 𝐻˚ 1 (𝛺)×𝐻 −1/2 (𝜕𝛺ℎ ) ≲ sup , (4.23)
0≠𝑦 ∈ 𝐻 1 (𝛺ℎ ) ∥𝑦∥ 𝐻 1 (𝛺ℎ )
thus proving the wellposedness of the primal DPG formulation for the Laplace
equation.
Example 4.5 (Maxwell equations). We now develop and analyse a primal DPG
method for the cavity problem in electromagnetics. Let the cavity 𝛺 be an open
bounded contractible domain in R3 , on the boundary of which the so-called perfect
electric conducting boundary condition is placed. Assuming that all time variations
are harmonic of frequency 𝜔 > 0, Maxwell equations in the cavity are
−ˆ𝚤 𝜔𝜇𝐻 + curl 𝐸 = 0 in 𝛺, (4.24a)
−ˆ𝚤 𝜔𝜖 𝐸 − curl 𝐻 = −𝐽 in 𝛺, (4.24b)
𝑛×𝐸 =0 on 𝜕𝛺. (4.24c)
The functions 𝐸, 𝐻, 𝐽 : 𝛺 → C3 represent electric field, magnetic field and im-
posed current, respectively, and 𝚤ˆ denotes the imaginary unit. We assume that
the electromagnetic material properties 𝜖 and 𝜇 are bounded uniformly positive
functions on 𝛺. The number 𝜔 > 0 denotes a fixed wavenumber. Eliminating 𝐻
from (4.24a) and (4.24b), we obtain the second-order (non-elliptic) equation
curl 𝜇 −1 curl 𝐸 − 𝜔2 𝜖 𝐸 = 𝑓 , (4.25)
where 𝑓 = 𝚤ˆ𝜔𝐽. Let
𝐻(curl, 𝛺) = {𝐹 ∈ 𝐿 2 (𝛺)3 : curl 𝐹 ∈ 𝐿 2 (𝛺)3 }
˚
and let 𝐻(curl, 𝛺) denote the subspace of vector fields in 𝐻(curl, 𝛺) with zero
tangential trace on 𝜕𝛺. A standard variational formulation for this problem is
obtained by multiplying (4.25) by a test function 𝐹 ∈ 𝐻(curl,
˚ 𝛺), integrating by
parts and using the boundary condition (4.24c): find 𝐸 ∈ 𝐻(curl,
˚ 𝛺) satisfying
(𝜇 −1 curl 𝐸, curl 𝐹)𝛺 − 𝜔2 (𝜖 𝐸, 𝐹)𝛺 = ⟨ 𝑓 , 𝐹⟩ (4.26)
for any given 𝑓 ∈ 𝐻(curl,
˚ 𝛺)′ . It is well known (see Monk 2003) that (4.26) has
a unique solution for every 𝑓 ∈ 𝐻(curl,
˚ 𝛺)′ whenever 𝜔 is not a resonance of the
cavity 𝛺, an assumption we place throughout this example.
The primal DPG method for the electric cavity problem is obtained by multiply-
ing (4.25) by a test function 𝐹 in the ‘broken’ space
Ö
𝐻(curl, 𝛺 ℎ ) = 𝐻(curl, 𝐾)
𝐾 ∈𝛺ℎ
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
312 L. Demkowicz and J. Gopalakrishnan
To set the element boundary term in the right space, a space akin to (4.10), let us
recall a few pertinent results on tangential traces on Lipschitz boundaries.
The tangential trace maps
𝐾
𝐸 ↦→ tr𝑛× 𝐸 : = (𝑛 × 𝐸)| 𝜕𝐾 , 𝐸 ↦→ tr⊤𝐾 𝐸 ≔ 𝑛 × (𝐸 × 𝑛)| 𝜕𝐾 ≡ 𝐸 ⊤ | 𝜕𝐾 ,
both well-defined for smooth vectors fields 𝐸 on any mesh element 𝐾 ∈ 𝛺 ℎ , can
be extended to continuous linear maps
𝐾
tr𝑛× : 𝐻(curl, 𝐾) → 𝐻 −1/2 (divF , 𝜕𝐾), tr⊤𝐾 : 𝐻(curl, 𝐾) → 𝐻 −1/2 (curlF , 𝜕𝐾)
by the work of Buffa, Costabel and Sheen (2002), which contains the definitions of
the codomain spaces and the surface derivatives (divF and curlF ) above. Moreover,
their results imply that the integration-by-parts formula
(curl 𝐸, 𝐹)𝐾 − (𝐸, curl 𝐹)𝐾 = (𝑛 × 𝐸, 𝐹)𝜕𝐾
for smooth vector fields 𝐸 and 𝐹 on 𝐾 can be extended to 𝐸, 𝐹 ∈ 𝐻(curl, 𝛺),
with the understanding that the right-hand side above becomes a duality pairing
⟨tr𝑛×
𝐾 𝐸, tr𝐾 𝐹⟩ −1/2 (div , 𝜕𝐾) and 𝐻 −1/2 (curl , 𝜕𝐾). Re-
⊤ 𝐻 −1/2 (curlF ,𝜕𝐾) between 𝐻 F F
using the notation of (·, ·)ℎ and ⟨·, ·⟩ℎ in (4.8) by extending inner products to vector
fields in the obvious way and letting
∑︁
𝐾
⟨𝑛 × 𝐸, 𝐹⟩ℎ = ⟨tr𝑛× 𝐸, tr⊤𝐾 𝐹⟩ 𝐻 −1/2 (curlF ,𝜕𝐾) ,
𝐾 ∈𝛺ℎ
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 313
and
𝐻 −1/2 (curlF , 𝜕𝛺 ℎ ) ≔ range(tr⊤ ),
∥ 𝐸ˆ ⊤ ∥ 𝐻 −1/2 (curlF ,𝜕𝛺ℎ ) ≔ inf ∥𝐸 ∥ 𝐻(curl,𝛺) , (4.31)
−1 {𝐸 }
𝐸 ∈tr⊤ ⊤
Complementing already defined trace operators tr𝑛 , tr𝑛× and tr⊤ (in (4.9), (4.29)
and (4.30), respectively), define standard 𝐻 1 trace operator, applied elementwise,
by
Ö
tr : 𝐻 1 (𝛺) → 𝐻 1/2 (𝜕𝐾), (tr 𝑢)| 𝜕𝐾 = 𝑢| 𝜕𝐾 , (4.34)
𝐾 ∈𝛺ℎ
and let
𝐻 1/2 (𝜕𝛺 ℎ ) ≔ range(tr), ∥ 𝑢∥
ˆ 𝐻 1/2 (𝜕𝛺ℎ ) ≔ inf ∥𝑢∥ 𝐻 1 (𝛺) ,
𝑢∈tr −1 { 𝑢}
ˆ
where the quotient norm is a standard norm obtained by a ‘minimal energy ex-
tension’ as an infimum of the norm of all extensions of 𝑢ˆ in the preimage set
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
314 L. Demkowicz and J. Gopalakrishnan
tr −1 {𝑢}
ˆ = {𝑢 ∈ 𝐻 1 (𝛺) : 𝑢| 𝜕𝐾 = 𝑢|
ˆ 𝜕𝐾 }. The identity (4.35b) below shows that
this infimum equals a supremum; in fact all identities of (4.35) are of a similar
‘inf = sup’ type.
Theorem 4.6 (Interface duality). The following identities hold for any 𝜎 ˆ 𝑛 in
𝐻 −1/2 (𝜕𝛺 ℎ ), 𝑢ˆ in 𝐻 1/2 (𝜕𝛺 ℎ ), 𝑛×𝐸ˆ in 𝐻 −1/2 (divF , 𝜕𝐾), and 𝐹ˆ⊤ in 𝐻 −1/2 (curl, 𝜕𝐾):
|⟨𝜎
ˆ 𝑛 , 𝑢⟩ℎ |
∥𝜎
ˆ 𝑛 ∥ 𝐻 −1/2 (𝜕𝛺ℎ ) = sup , (4.35a)
0≠𝑢∈ 𝐻 1 (𝛺ℎ ) ∥𝑢∥ 𝐻 1 (𝛺ℎ )
|⟨𝑛 · 𝜎, 𝑢⟩
ˆ ℎ|
∥ 𝑢∥
ˆ 𝐻 1/2 (𝜕𝛺ℎ ) = sup , (4.35b)
0≠𝜎 ∈ 𝐻(div,𝛺ℎ ) ∥𝜎∥ 𝐻(div,𝛺ℎ )
|⟨𝑛 × 𝐸,
ˆ 𝐹⟩ℎ |
∥𝑛 × 𝐸ˆ ∥ 𝐻 −1/2 (divF ,𝜕𝛺ℎ ) = sup , (4.35c)
0≠𝐹 ∈ 𝐻(curl,𝛺ℎ ) ∥𝐹 ∥ 𝐻(curl,𝛺ℎ )
|⟨𝑛 × 𝐸, 𝐹ˆ⊤ ⟩ℎ |
∥ 𝐹ˆ⊤ ∥ 𝐻 −1/2 (curl,𝜕𝛺ℎ ) = sup . (4.35d)
0≠𝐸 ∈ 𝐻(curl,𝛺ℎ ) ∥𝐸 ∥ 𝐻(curl,𝛺ℎ )
Furthermore,
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 315
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
316 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 317
shall refer to as a ‘Fortin operator’, based on similar such operators in the study of
mixed methods (Brezzi and Fortin 1991).
Theorem 5.2 (Fortin operator gives DPG convergence). Suppose (1.3) holds,
𝑋ℎ ⊂ 𝑋 and 𝑌 𝑟 ⊂ 𝑌 . Assume that there is a bounded linear operator 𝛱 : 𝑌 → 𝑌 𝑟 ,
of operator norm ∥𝛱 ∥, such that for all 𝑤 ℎ ∈ 𝑋ℎ and all 𝑣 ∈ 𝑌 ,
𝑏(𝑤 ℎ , 𝑣 − 𝛱 𝑣) = 0. (5.5)
Then the DPG method (5.4) is uniquely solvable for 𝑥 ℎ and
∥𝑏∥ ∥𝛱 ∥
∥𝑥 − 𝑥 ℎ ∥ 𝑋 ≤ inf ∥𝑥 − 𝑧 ℎ ∥ 𝑋 , (5.6)
𝛾 𝑧ℎ ∈𝑋ℎ
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
318 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 319
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
320 L. Demkowicz and J. Gopalakrishnan
where
𝑋0,ℎ = {𝑤 ∈ 𝐻˚ 1 (𝛺) : 𝑤| 𝐾 ∈ 𝑃 𝑝+1 (𝐾) for all 𝐾 ∈ 𝛺 ℎ }, (5.15a)
𝑋ˆ ℎ = tr𝑛 𝑅 ℎ𝑝+1 , (5.15b)
𝑌ℎ = 𝑃 𝑝+𝑁 (𝛺 ℎ ) (5.15c)
for some degree 𝑝 ≥ 0. Clearly, the above set 𝑋0,ℎ is a standard Lagrange finite
element subspace of 𝐻˚ 1 (𝛺) and 𝑌ℎ is a standard DG space. Also, the space 𝑋ˆ ℎ set
above is a subspace of 𝑋ˆ = 𝐻 −1/2 (𝜕𝛺 ℎ ) since 𝑅 ℎ𝑝+1 ⊆ 𝐻(div, 𝛺). An alternative
characterization of 𝑋ˆ ℎ = tr𝑛 𝑅 ℎ𝑝+1 can be given assuming that every 𝐹 ∈ Fℎ is
provided a fixed unit normal 𝑛 𝐹 , which equals the outward pointing unit normal 𝑛
if 𝐹 ⊂ 𝜕𝛺, and equals either 𝑛 or −𝑛 on an interior facet 𝐹 shared by an element
𝐾 with unit outward normal 𝑛. Then it is easy to see from the well-known degrees
of freedom of the Raviart–Thomas space that
𝑋ˆ ℎ = {𝑟ˆ𝑛 : on every 𝐾 ∈ 𝛺 ℎ and each 𝐹 ∈ △ 𝑁 −1 𝐾,
there is a 𝜇 ∈ 𝑃 𝑝 (𝐹) such that 𝑟ˆ𝑛 | 𝜕𝐾 = (𝜇𝑛 𝐹 ) · 𝑛| 𝜕𝐾 }. (5.16)
Consequently, one may choose to implement 𝑋ˆ ℎ without using 𝑅 ℎ𝑝+1 . An imple-
mentation of (5.16) can proceed using only the standard polynomial space 𝑃 𝑝 (𝐹)
on each facet 𝐹 ∈ Fℎ together with some fixed facet orientation given by 𝑛 𝐹 .
Let us examine what the Fortin condition (5.5) entails for this discrete setting.
Let (𝑤 ℎ , 𝑟ˆℎ ) ∈ 𝑋ℎ . Since 𝑟ˆℎ = 𝑟 ℎ · 𝑛 for some 𝑟 ℎ ∈ 𝑅 ℎ𝑝+1 , using the 𝑏(·, ·) in (5.14),
condition (5.5) reads as follows:
(grad 𝑤 ℎ , grad(𝑦 − 𝛱 𝑦))ℎ + ⟨𝑟 ℎ · 𝑛, 𝑦 − 𝛱 𝑦⟩ℎ = 0 (5.17)
for all 𝑤 ℎ ∈ 𝑋ℎ,0 and 𝑟 ℎ ∈ 𝑅 ℎ𝑝+1 . By integration by parts, we see that (5.17) is
implied by
(𝑦 − 𝛱 𝑦, Δ𝑤 ℎ )𝐾 = 0 and
(𝑦 − 𝛱 𝑦, (grad 𝑤 ℎ − 𝑟 ℎ ) · 𝑛)𝜕𝐾 = 0
grad
on every 𝐾 ∈ 𝛺 ℎ . Once 𝛱 is set to 𝛱𝑟 of Theorem 5.4, these two identities follow
from (5.13a) and (5.13b), respectively, and we are ready to apply Theorem 5.2. Let
(𝑢, 𝑞ˆ 𝑛 ) ∈ 𝑋 be the exact solution, and let 𝑞 = grad 𝑢, so that 𝑞ˆ 𝑛 = tr𝑛 (𝑞). If
𝑢 ℎ ∈ 𝑋0,ℎ and 𝑞ˆ 𝑛,ℎ ∈ 𝑋ˆ ℎ together form the solution of the practical DPG method
with discrete spaces set by (5.15), then Theorem 5.2 yields
∥𝑢 − 𝑢 ℎ ∥ 𝐻 1 (𝛺) + ∥ 𝑞ˆ 𝑛 − 𝑞ˆ 𝑛,ℎ ∥ 𝐻 −1/2 (𝜕𝛺ℎ )
≲ inf ∥𝑢 − 𝑤 ℎ ∥ 𝐻 1 (𝛺) + ∥ 𝑞ˆ 𝑛 − 𝑟ˆ𝑛,ℎ ∥ 𝐻 −1/2 (𝜕𝛺ℎ ) .
(𝑤ℎ ,𝑟𝑛,ℎ )∈𝑋ℎ
To obtain convergence rates, the standard approximation rates for the Lagrange
finite element space 𝑋ℎ,0 ,
inf ∥𝑢 − 𝑤 ℎ ∥ 𝐻 1 (𝛺) ≲ ℎ 𝑠 |𝑢| 𝐻 1+𝑠 (𝛺) , 0 ≤ 𝑠 ≤ 𝑝 + 1, (5.18)
𝑤ℎ ∈𝑋ℎ,0
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 321
may be used to bound the first term in the infimum. For the other term we use the
definition of the 𝐻 −1/2 (𝜕𝛺 ℎ )-norm via the minimal extension norm, that is,
inf ∥ 𝑞ˆ 𝑛 − 𝑟ˆ𝑛,ℎ ∥ 𝐻 −1/2 (𝜕𝛺ℎ ) = inf ∥𝑞 − 𝑟 ℎ ∥ 𝐻(div,𝛺) . (5.19)
𝑟𝑛,ℎ ∈ 𝑋ˆ ℎ 𝑟ℎ ∈ 𝑅 ℎ𝑝+1
To obtain convergence rates from this, we use the standard Raviart–Thomas inter-
polant 𝛱 𝑅 𝑞 ∈ 𝑅 ℎ𝑝+1 , which is well-defined when 𝑞 ∈ 𝐻 𝑠 (𝛺) 𝑁 ∩ 𝐻(div, 𝛺) with
𝑠 > 1/2, together with its commutativity property
div 𝛱 𝑅 𝑞 = 𝛱 𝑝 div 𝑞, (5.20)
where 𝛱 𝑝 denotes the 𝐿 2 (𝛺)-orthogonal projection into 𝑃 𝑘 (𝛺 ℎ ), as follows:
inf ∥𝑞 − 𝑟 ℎ ∥ 2𝐻(div,𝛺) ≤ ∥𝑞 − 𝛱 𝑅 𝑞∥ 2𝛺 + ∥ div(𝑞 − 𝛱 𝑅 𝑞)∥ 2𝛺
𝑟ℎ ∈ 𝑅 ℎ𝑝+1
≤ ∥𝑞 − 𝛱 𝑅 𝑞∥ 2𝛺 + ∥(𝐼 − 𝛱 𝑝 ) div 𝑞∥ 2𝛺
≲ ℎ2𝑠 |𝑞| 2𝐻 𝑠 (𝛺) + ℎ2𝑠 | div 𝑞| 2𝐻 𝑠 (𝛺) , (5.21)
by the usual Bramble–Hilbert argument. Combining (5.18) and (5.21), we obtain
∥𝑢 − 𝑢 ℎ ∥ 𝐻 1 (𝛺) + ∥ 𝑞ˆ 𝑛 − 𝑞ˆ 𝑛,ℎ ∥ 𝐻 −1/2 (𝜕𝛺ℎ ) ≲ ℎ 𝑠 |𝑢| 𝐻 1+𝑠 (𝛺) + ℎ 𝑠 |Δ𝑢| 𝐻 𝑠 (𝛺) , (5.22)
for 1/2 < 𝑠 ≤ 𝑝 + 1.
Although the convergence rate with respect to ℎ in (5.22) is optimal, the last
term demands too much regularity. In the remainder of this example, we show how
to improve the argument using (4.35a) of Theorem 4.6. Instead of (5.19), we start
by applying (4.35a),
inf ∥ 𝑞ˆ 𝑛 − 𝑟ˆ𝑛,ℎ ∥ 𝐻 −1/2 (𝜕𝛺ℎ ) ≤ ∥ tr𝑛 (𝑞 − 𝛱 𝑅 𝑞)∥ 𝐻 −1/2 (𝜕𝛺ℎ )
𝑟𝑛,ℎ ∈ 𝑋ˆ ℎ
⟨tr𝑛 (𝑞 − 𝛱 𝑅 𝑞), 𝑦⟩ℎ
= sup . (5.23)
0≠𝑦 ∈ 𝐻 1 (𝛺ℎ ) ∥𝑦∥ 𝐻 1 (𝛺ℎ )
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
322 L. Demkowicz and J. Gopalakrishnan
Using these two estimates to bound the last term in (5.24) and returning to (5.23),
∥𝑢 − 𝑢 ℎ ∥ 𝐻 1 (𝛺) + ∥ 𝑞ˆ 𝑛 − 𝑞ˆ 𝑛,ℎ ∥ 𝐻 −1/2 (𝜕𝛺ℎ )
(
ℎ 𝑠 |𝑢| 𝐻 1+𝑠 (𝛺) + ℎ∥Δ𝑢∥ 𝛺 , 1/2 < 𝑠 < 1,
≲ (5.25)
ℎ 𝑠 |𝑢| 𝐻 1+𝑠 (𝛺) + ℎ 𝑠 |Δ𝑢| 𝐻 𝑠−1 (𝛺) , 1 ≤ 𝑠 ≤ 𝑘 + 1.
This gives optimal rates at reduced regularity requirements compared to (5.22).
For example, in the lowest-order case, if linear Lagrange elements are used for
approximating 𝑢, and if 𝑢 ∈ 𝐻 2 (𝛺), then the DPG error in 𝑢 is 𝑂(ℎ) (a convergence
rate and regularity requirement comparable to the standard finite element method),
and additionally, at the price of including a piecewise constant flux 𝑞ˆ 𝑛 , the DPG
method gives a flux error that is also 𝑂(ℎ).
Having shown a typical application of a Fortin operator to analyse a DPG method
in the above example, let us now proceed to detail the construction of the needed
grad
operator 𝛱𝑟 . Let 𝐾 be an 𝑁-simplex and let
𝑃˚𝑟 (𝐾) = {𝑢 ∈ 𝑃𝑟 (𝐾) : 𝑢| 𝜕𝐾 = 0},
𝐵𝑟0 (𝐾) = {𝑢 ∈ 𝑃𝑟 (𝐾) : 𝑢| 𝐸 = 0 for all 𝐸 ∈ △ 𝑁 −2 𝐾 }.
Let 𝜆0 , . . . , 𝜆 𝑁 denote the standard linear barycentric coordinate functions of an
𝑁-simplex 𝐾, let 𝐹𝑖 be the facet in △ 𝑁 −1 𝐾 where 𝜆𝑖 vanishes, and let
𝑁
Ö 𝑏𝐾 Ö
𝑏𝐾 = 𝜆 𝑗, 𝑏 𝐹𝑖 = = 𝜆𝑗. (5.26)
𝑗=0
𝜆𝑖 𝑗≠𝑖
and for every 𝑣 ∈ 𝐻 1 (𝐾), there is a unique 𝛱𝑟0 𝑣 ∈ 𝐵𝑟0 (𝐾) satisfying
(𝛱𝑟0 𝑣 − 𝑣, 𝑞)𝐾 = 0 for all 𝑞 ∈ 𝑃 𝑝−1 (𝐾), (5.28a)
(𝛱𝑟0 𝑣 − 𝑣, 𝜇)𝐹 = 0 for all 𝜇 ∈ 𝑃 𝑝 (𝐹), 𝐹 ∈ △ 𝑁 −1 𝐾, (5.28b)
∥𝛱𝑟0 𝑣∥ 𝐿 2 (𝐾) + ℎ 𝐾 ∥ grad 𝛱𝑟0 𝑣∥ 𝐿 2 (𝐾) ≲ ∥𝑣∥ 𝐿 2 (𝐾) + ℎ 𝐾 ∥ grad 𝑣∥ 𝐿 2 (𝐾) . (5.28c)
Proof. Using the space of polynomials of vanishing trace, Í we can count the
dimensions of 𝐵𝑟0 (𝐾). Indeed, dim 𝐵𝑟0 (𝐾) = dim 𝑃˚𝑟 (𝐾) + 𝐹 ∈ △ 𝑁 −1 𝐾 dim 𝑃˚𝑟 (𝐹).
Note that 𝑃˚𝑟 (𝐾) = 𝑏 𝐾 𝑃𝑟 − 𝑁 −1 (𝐾) and 𝑃˚𝑟 (𝐹) = 𝑏 𝐹 𝑃𝑟 − 𝑁 (𝐹) for any 𝐹 ∈ △ 𝑁 −1 𝐾.
Therefore, by choosing 𝑟 = 𝑝 + 𝑁, we have
dim 𝑃˚𝑟 (𝐾) = dim 𝑃 𝑝−1 (𝐾) and dim 𝑃˚𝑟 (𝐹) = dim 𝑃 𝑝 (𝐹),
and consequently (5.28a)–(5.28b) is a square system for 𝛱𝑟0 𝑣.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 323
Now the existence of the stated 𝛱𝑟0 𝑣 will follow from uniqueness, that is, it
suffices to prove that if 𝑣 = 0, then 𝛱𝑟0 𝑣 = 0. Since 𝛱𝑟0 𝑣 ∈ 𝐵𝑟0 (𝐾), on any
face 𝐹 ∈ △ 𝑁 −1 𝐾, we may write (𝛱𝑟0 𝑣)| 𝐹 = 𝑏 𝐹 𝑤 𝑝 for some 𝑤 𝑝 ∈ 𝑃 𝑝 (𝐹). But
then (5.28b) implies that 𝛱𝑟0 𝑣 must vanish on 𝜕𝐾, so 𝛱𝑟0 𝑣 = 𝑏 𝐾 𝑧 𝑝−1 for some
𝑧 𝑝−1 ∈ 𝑃 𝑝−1 (𝐾). Then (5.28a) implies that 𝛱𝑟0 𝑣 = 0 on 𝐾. The estimate (5.28c)
now follows by a standard scaling argument using a reference 𝑁-simplex.
grad
Proof of Theorem 5.4. Let 𝑣 ∈ 𝐻 1 (𝐾) on an 𝑁-simplex 𝐾. Define 𝛱𝑟 𝑣 =
𝛱𝑟0 (𝑣 − 𝑣) + 𝑣, where 𝑣 denotes the mean value of 𝑣 on 𝐾:
∫
1
𝑣= 𝑣.
|𝐾 | 𝐾
Equations (5.13a) and (5.13b) immediately follow from (5.28a) and (5.28b) of
Lemma 5.6 since
grad
𝛱𝑟 𝑣 − 𝑣 = (𝛱𝑟0 − 𝐼)(𝑣 − 𝑣).
To prove (5.13c), we use (5.28c) and the Poincaré inequality as follows:
grad
∥𝛱𝑟 𝑣∥ 𝐿 2 (𝐾) ≤ ∥𝑣∥ 𝐿 2 (𝐾) + ∥𝛱𝑟0 (𝑣 − 𝑣)∥ 𝐿 2 (𝐾)
≲ ∥𝑣∥ 𝐿 2 (𝐾) + ∥𝑣 − 𝑣∥ 𝐿 2 (𝐾) + ℎ 𝐾 ∥ grad(𝑣 − 𝑣)∥ 𝐿 2 (𝐾)
≲ ∥𝑣∥ 𝐿 2 (𝐾) + ℎ 𝐾 ∥ grad 𝑣∥ 𝐿 2 (𝐾)
and
grad
ℎ 𝐾 ∥ grad 𝛱𝑟 𝑣∥ 𝐿 2 (𝐾) = ℎ 𝐾 ∥ grad 𝛱𝑟0 (𝑣 − 𝑣)∥ 𝐿 2 (𝐾)
≲ ∥𝑣 − 𝑣∥ 𝐿 2 (𝐾) + ℎ 𝐾 ∥ grad(𝑣 − 𝑣)∥ 𝐿 2 (𝐾)
≲ ℎ 𝐾 ∥ grad 𝑣∥ 𝐿 2 (𝐾) .
These estimates together imply (5.13c).
When a lower-order reaction term, say (𝑢, 𝑦)𝛺 , is added to the Laplace formu-
lation (5.14), skimming through the analysis of Example 5.5, we immediately see
that we would need another Fortin operator where the moment condition (5.13a)
is strengthened to 𝑞 ∈ 𝑃 𝑝 (𝐾) in place of 𝑞 ∈ 𝑃 𝑝−1 (𝐾). To perform such modific-
ations easily, and also to better understand the structure of the Fortin operator we
have presented, it is useful to know an explicit representation of the prior Fortin
operator and its generalization, which we describe now.
Denote the set of (𝑁 + 1)-term multi-indices of length 𝑚 by
( 𝑁 +1
)
∑︁
𝑁 +1
I𝑚 = 𝛽 ≡ (𝛽1 , . . . , 𝛽 𝑁 +1 ) : 𝛽𝑖 ≥ 0 are integers and |𝛽| ≡ 𝛽𝑖 = 𝑚 .
𝑖=1
(5.29)
𝛼𝑁 +1
On an 𝑁-simplex, recall that a basis for 𝑃𝑚 (𝐾) is given by 𝜆 𝛼 = 𝜆1𝛼1 𝜆2𝛼2 · · · 𝜆 𝑁 +1
for all multi-indices 𝛼 ∈ I𝑚𝑁 +1 . Let 𝜂 𝐾
𝛼 = 𝑏 𝐾 𝜆 𝛼 and let 𝜒 𝐾 ∈ 𝑃 (𝐾) denote the
𝛽 𝑚
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
324 L. Demkowicz and J. Gopalakrishnan
𝛼, 𝛽 ∈ I𝑚𝑁 +1 ,
𝛽
𝜂𝐾 𝐾
𝛼 , 𝜒𝛽 𝐾 = 𝛿 𝛼 , (5.30)
𝛽
where 𝛿 𝛼 equals one or zero depending on whether 𝛼 equals 𝛽 or not. Let
∑︁
𝛱𝑚𝐾 𝑣 = 𝑣, 𝜒 𝛼𝐾 𝐾 𝜂 𝐾
𝛼, (5.31a)
𝑁 +1
𝛼∈I𝑚
so we obtain
𝛱𝑚𝐾 𝑣 − 𝑣, 𝑞 𝐾
=0 for all 𝑞 ∈ 𝑃𝑚 (𝐾), 𝑣 ∈ 𝐿 2 (𝐾), (5.31b)
after expanding 𝑞 in the 𝜒𝛽𝐾 -basis.
Since this construction works on a simplex of any dimension, we can repeat it
on any subsimplex of 𝐾. The barycentric coordinates of a subsimplex 𝐹 ∈ △ 𝑁 −1 𝐾
are simply the restrictions of those of 𝐾 to 𝐹, omitting the one that vanishes on 𝐹.
Using them, we repeat the construction, now with shorter multi-indices 𝛼, 𝛽 ∈ I 𝑘𝑁 .
Namely, let 𝜂 𝐹𝛼 = 𝑏 𝐹 𝜆 𝛼 and let 𝜒𝛽𝐹 ∈ 𝑃 𝑘 (𝐹) form the dual basis of 𝜆 𝛼 in the
(𝑏 𝐹 ·, ·)𝐹 inner product. Then set
∑︁
𝛱𝑘𝐹 𝑣 = 𝑣, 𝜒 𝛼𝐹 𝐹 𝜂 𝐹𝛼 . (5.32a)
𝛼∈I 𝑘𝑁
Note that its trace on a facet 𝐹 is determined solely by the 𝛱𝑘𝐹 -contribution since
˜
the traces after an application of 𝛱𝑚𝐾 or 𝛱𝑘𝐹 are zero on 𝐹 for any 𝐹˜ ≠ 𝐹. Note
that when 𝑚 = 𝑝 − 1 and 𝑘 = 𝑝, we recover the operator of Theorem 5.4.
Theorem 5.7 (A more general 𝐻 1 (𝐾) Fortin operator). The above-defined op-
erator
grad
𝛱𝑘,𝑚 : 𝐻 1 (𝐾) → 𝑃𝑟 (𝐾) for 𝑟 = max(𝑚 + 𝑁 + 1, 𝑘 + 𝑁)
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 325
Bibliographical notes. Theorem 5.4 and the construction of the Fortin operator
grad
𝛱 𝑝+3 are taken from Gopalakrishnan and Qiu (2014). Its generalization in (5.33)
is based on the recent work of Führer and Heuer (2024). They further show that
generalizing such polynomial expressions to certain exponential ones, estimates
like (5.13) but with ∥ · ∥ 𝐻 1 (𝐾) replaced by
1/2
∥𝑣∥ 𝑎 = ∥𝑣∥ 2𝐾 + 𝑎∥ grad 𝑣∥ 2𝐾
for some small parameter 𝑎, can be obtained robustly in the parameter 𝑎 as 𝑎 → 0.
The discrete stability of the primal DPG method for the Laplace equation, discussed
in Example 5.5, was first considered in Demkowicz and Gopalakrishnan (2013).
There, and in earlier DPG analyses such as that of Demkowicz and Gopalakrishnan
(2011a), error estimates comparable to (5.22) that demand extra regularity can be
found. The discussion in Example 5.5, leading to (5.25) with better regularity
requirements, is taken from the more recent work of Führer (2018, Theorem 5).
satisfying certain moment conditions that are useful in analysis of DPG methods
where 𝐻(div, 𝛺 ℎ ) features in the test space.
We will perform our construction on the reference unit 𝑁-simplex 𝐾ˆ and map it
to a general 𝑁-simplex 𝐾. Let 𝑆 𝐾 : 𝐾ˆ → 𝐾 be a one-to-one affine map that maps
𝐾ˆ onto a general tetrahedron 𝐾, and let [𝑆 ′𝐾 ] denote the Jacobian derivative matrix
of 𝑆 𝐾 . Given vector fields 𝑞 and 𝐸 on 𝐾, we use the following pullback to map
them to 𝐾:ˆ
Ψ(𝑞) = (det[𝑆 ′𝐾 ]) [𝑆 ′𝐾 ] −1 (𝑞 ◦ 𝑆 𝐾 ). (5.35)
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
326 L. Demkowicz and J. Gopalakrishnan
where 𝑛ˆ is the unit outward normal on 𝜕 𝐾ˆ and 𝑅 𝑘+1 (𝐾) ˆ is as in (5.11). We claim
that the equations
div
𝛱ˆ 𝑘+1 ˆ 𝑞ˆ 𝐾ˆ = (𝜏,
𝜏, ˆ 𝑞)
ˆ 𝐾ˆ ˆ 𝑁,
for all 𝑞ˆ ∈ 𝑃 𝑘−1 (𝐾) (5.39a)
div
𝛱ˆ 𝑘+1 𝜏ˆ · 𝑛,
ˆ 𝑤ˆ 𝜕𝐾ˆ = ⟨𝜏ˆ · 𝑛,
ˆ 𝑤⟩ˆ 𝐻 1/2 (𝜕𝐾) for all 𝑤ˆ ∈ 𝑃 𝑘 (𝐾)
ˆ (5.39b)
div 𝜏ˆ ∈ 𝐵div (𝐾)
uniquely determine 𝛱ˆ 𝑘+1 ˆ and thus define a linear continuous operator
𝑘+1
div ˆ → 𝐵div (𝐾).
𝛱ˆ 𝑘+1 : 𝐻(div, 𝐾) 𝑘+1
ˆ
ˆ ⊂
div 𝜏ˆ is in 𝐵div (𝐾)
Indeed, if the right-hand sides of (5.39) vanish, then since 𝛱ˆ 𝑘+1 𝑘+1
ˆ we find that 𝛱ˆ div is a function in the Raviart–Thomas space all of whose
𝑅 𝑘+1 (𝐾), 𝑘+1
canonical degrees of freedom vanish (see e.g. Arnold, Falk and Winther 2006 or
Nédélec 1980, Definition 5), so
div
𝛱ˆ 𝑘+1 𝜏ˆ = 0.
div 𝜏.
Since the system (5.39) is square, we conclude that it uniquely defines 𝛱ˆ 𝑘+1 ˆ
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 327
Next, we map this operator to a general simplex 𝐾 using the Piola transform Ψ
in (5.35):
𝛱 div = Ψ −1 ◦ 𝛱ˆ div ◦ Ψ.
𝑘+1 𝑘+1
By standard mapping arguments, the stated moment conditions of 𝛱𝑘+1div now follow
from (5.39). The moment conditions also imply that for any 𝜔 ∈ 𝑃 𝑘 (𝐾),
div
div
div
div 𝛱𝑘+1 𝜏 , 𝜔 𝐾 = − 𝛱𝑘+1 𝜏, grad 𝜔 𝐾 + 𝛱𝑘+1 𝜏 · 𝑛, 𝜔 𝜕𝐾
= −(𝜏, grad 𝜔)𝐾 + (𝜔, 𝜏 · 𝑛)𝜕𝐾
= (div 𝜏, 𝜔)𝐾 ,
div is a continuous
thus proving the commutativity property (5.37). Also, since 𝛱ˆ 𝑘+1
ˆ standard scaling arguments prove
operator on 𝐻(div, 𝐾),
div div
∥𝛱𝑘+1 𝜏∥ 𝐾 + ℎ 𝐾 ∥ div 𝛱𝑘+1 𝜏∥ 𝐾 ≲ ∥𝜏∥ 𝐾 + ℎ 𝐾 ∥ div 𝜏∥ 𝐾 .
Combined with the better bound on the divergence term,
div
∥ div 𝛱𝑘+1 𝜏∥ 𝐾 ≤ ∥ div 𝜏∥ 𝐾 ,
which obviously follows from (5.37), the stated norm bound is also proved.
Bibliographical notes. Theorem 5.8 and its proof are from Gopalakrishnan and
Qiu (2014). For a construction in 𝐻(div, 𝐾) in the same spirit as Theorem 5.7, see
Führer and Heuer (2024).
they are part of a family of local commuting Fortin operators. We restrict ourselves
to the three-dimensional (3D) 𝑁 = 3 case. Let
𝑁 𝑝 (𝐷) = 𝑃 𝑝−1 (𝐷)3 + 𝑥 × 𝑃 𝑝−1 (𝐷)3
denote the Nédélec element (Nédélec 1980). Together with the Raviart–Thomas
element 𝑅 𝑝+1 (𝐾) in (5.11), it forms the following well-known (see e.g. Arnold et al.
2006) exact complex:
grad curl div
0 𝑃 𝑝+1 (𝐾)/R 𝑁 𝑝+1 (𝐾) 𝑅 𝑝+1 (𝐾) 𝑃 𝑝 (𝐾) 0. (5.40)
We will prove the following result in this subsection. There 𝛱 𝑝 denotes the 𝐿 2 -
orthogonal projection onto 𝑃 𝑝 (𝐾). In order to verify condition (5.5) in various
DPG convergence analyses, the moment conditions (5.43)–(5.48) listed below are
helpful.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
328 L. Demkowicz and J. Gopalakrishnan
such that for any 𝑣 ∈ 𝐻 1 (𝐾), 𝐸 ∈ 𝐻(curl, 𝐾) and 𝜏 ∈ 𝐻(div, 𝐾), the norm estimates
grad
∥ 𝛱 𝑝+3 𝑣∥ 𝐻 1 (𝐾) ≲ ∥𝑣∥ 𝐻 1 (𝐾) , (5.41a)
curl
∥ 𝛱 𝑝+3 𝐸 ∥ 𝐻(curl,𝐾) ≲ ∥𝐸 ∥ 𝐻(curl,𝐾) , (5.41b)
div
∥ 𝛱 𝑝+3 𝜏∥ 𝐻(div,𝐾) ≲ ∥𝜏∥ 𝐻(div,𝐾) (5.41c)
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 329
ˆ
fields 𝐸 on 𝐾, we use the following pullback to map it to 𝐾:
Φ(𝐸) = [𝑆 ′𝐾 ] 𝑡 (𝐸 ◦ 𝑆 𝐾 ). (5.49)
It is easy to see that
curl(Φ(𝐸)) = Ψ(curl 𝐸). (5.50)
We begin with a preliminary lemma whose relevance will be clear soon. Let
𝑁˚ 𝑝 (𝐾) = {𝑞 ∈ 𝑁 𝑝 (𝐾) : 𝑛 × 𝑞| 𝜕𝐾 = 0}.
Lemma 5.10. For any 𝑝 ≥ 0, if 𝐹 ∈ 𝐻(curl, 𝐾) satisfies
(curl 𝐹, 𝑤)𝐾 = 0 for all 𝑤 ∈ 𝑁˚ 𝑝+1 (𝐾), (5.51)
then there is a 𝜙 ∈ 𝑃 𝑝+3 (𝐾) such that
(𝐹 + grad 𝜙, 𝑣)𝐾 = 0 for all 𝑣 ∈ 𝑃 𝑝 (𝐾)3 . (5.52)
Proof. Proving (5.52) amounts to proving that there is a 𝜙 ∈ 𝑃 𝑝+3 (𝐾) solving
𝐴𝜙 = 𝑏, (5.53)
where 𝐴 and 𝑏 are defined using the 𝐿 2 (𝐾)3 -orthogonal projection 𝛱 𝑝 into
𝑃 𝑝 (𝐾)3 by
𝐴 = 𝛱 𝑝 grad : 𝑃 𝑝+3 (𝐾) → 𝑃 𝑝 (𝐾)3 , 𝑏 = −𝛱 𝑝 𝐹.
In other words, it suffices to show that 𝑏 ∈ range(𝐴) = ker(𝐴∗ )⊥ , where 𝐴∗ is the
𝐿 2 -adjoint of 𝐴.
Any 𝑞 ∈ 𝑃 𝑝 (𝐾)3 is in ker(𝐴∗ ) if and only if
(𝑢, 𝐴∗ 𝑞) = (𝐴𝑢, 𝑞)𝐾 = (grad 𝑢, 𝑞)𝐾
= −(𝑢, div 𝑞)𝐾 + (𝑢, 𝑞 · 𝑛)𝜕𝐾 = 0 (5.54)
for all 𝑢 ∈ 𝑃 𝑝+3 (𝐾). Recall the bubble functions 𝑏 𝐾 and 𝑏 𝐹 from (5.26). Choosing
𝑢 = 𝑏 𝐾 div 𝑞 in (5.54), we find that div 𝑞 = 0. Then, removing the term containing
div 𝑞 and choosing 𝑢 = (𝑞 · 𝑛 𝐹 )𝑏 𝐹 in (5.54), we find that (𝑞 · 𝑛 𝐹 )| 𝐹 = 0, an argument
that can repeated on every facet 𝐹. Including the obvious converse as well, we have
proved that 𝑞 ∈ ker(𝐴∗ ) if and only if div 𝑞 = 0 and 𝑞 · 𝑛| 𝜕𝐾 = 0, that is,
ker(𝐴∗ ) = curl 𝑁˚ 𝑝+1 (𝐾).
Note that for any 𝑤 ∈ 𝑁˚ 𝑝+1 (𝐾), by the given condition (5.51),
(𝑏, curl 𝑤) = −(𝐹, curl 𝑤) = −(curl 𝐹, 𝑤) = 0,
so 𝑏 ∈ ker(𝐴∗ )⊥ = range(𝐴) and (5.53) has a solution.
curl , we need an intermediate operator 𝛱
Before constructing the required 𝛱 𝑝+3 ˆ𝑐
𝑝+3
ˆ Let
on a reference unit tetrahedron 𝐾.
𝐷 𝑝+2 (𝐾) ˆ = {𝑟 ∈ 𝑅 𝑝+3 (𝐾)
ˆ = curl 𝑁 𝑝+3 (𝐾) ˆ : div 𝑟 = 0}
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
330 L. Demkowicz and J. Gopalakrishnan
and
ˆ → 𝐷 𝑝+2 (𝐾).
𝐶 = curl : 𝑁 𝑝+3 (𝐾) ˆ
Using 𝛱ˆ 𝑝 , the 𝐿 2 (𝐾) ˆ 3 , define 𝐵 = 𝛱ˆ 𝑝 : 𝑁 𝑝+3 →
ˆ 3 -orthogonal projection into 𝑃 𝑝 (𝐾)
𝑃 𝑝 (𝐾) . The codomains of 𝐵 and 𝐶 are endowed with the 𝐿 2 -norm, which then
ˆ 3
naturally define their 𝐿 2 -adjoints 𝐵∗ and 𝐶 ∗ . Note that one of the commutativity
properties in (5.42) and one of the moment conditions (5.45) read, respectively, as
follows:
div
𝐶𝐹 = 𝛱 𝑝+3 curl 𝐸, 𝐵𝐹 = 𝛱ˆ 𝑝 𝐸, (5.55)
curl 𝐸. Accordingly, we seek the result of the application of the Fortin
with 𝐹 = 𝛱 𝑝+3
operator in the set
𝑆(𝐸) = {𝐹 ∈ 𝑁 𝑝+3 (𝐾)
ˆ : 𝐹 satisfies (5.55)}. (5.56)
𝑐 , consider the problem of finding
For defining the intermediate operator 𝛱ˆ 𝑝+3
𝛱 𝑝+3 𝐸 ∈ 𝑁 𝑝+3 (𝐾), 𝜆 ∈ 𝑃 𝑝 (𝐾) and 𝜇 ∈ 𝐷 𝑝+2 (𝐾)
ˆ 𝑐 ˆ ˆ 3 ˆ satisfying
𝑐
𝛱ˆ 𝑝+3 𝐸 + 𝐵∗ 𝜆 + 𝐶 ∗ 𝜇 = 0, (5.57a)
𝑐
𝐵 𝛱ˆ 𝑝+3 𝐸 = 𝛱ˆ 𝑝 𝐸, (5.57b)
𝐶 𝛱ˆ 𝑐 𝐸
𝑝+3 = 𝛱ˆ div curl 𝐸,
𝑝+3 (5.57c)
div is as defined in (5.39). One may view 𝜆 and 𝜇 above as Lagrange
where 𝛱ˆ 𝑝+3
multipliers for the constrained minimization problem
𝑐
𝛱ˆ 𝑝+3 𝐸 = arg min ∥𝐹 ∥ 2𝐾ˆ , (5.58)
𝐹 ∈𝑆(𝐸)
that is, the right-hand side of (5.57) is in the range of [ 𝐶𝐵 ] (or equivalently 𝑆(𝐸)
is a non-empty feasible set of constraints). Hence, by standard arguments (see
Brezzi and Fortin 1991, Proposition 1.1, p. 38), there exists a solution to (5.57) and
𝑐 𝐸 component of the solution is unique. The linearity of 𝛱
moreover the 𝛱ˆ 𝑝+3 ˆ𝑐 𝐸
𝑝+3
with respect to the right-hand sides of (5.57) is obvious and the right-hand sides
in turn depend linearly on 𝐸. Furthermore, since the ranges of 𝐵 and 𝐶 are closed
finite-dimensional spaces, there is a 𝑐 𝐾ˆ > 0 such that (see Brezzi and Fortin 1991,
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 331
We proceed to prove that this operator has all the required properties.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
332 L. Demkowicz and J. Gopalakrishnan
Proof of Theorem 5.9. We have already shown that the 𝑁-dimensional operator
div in Theorem 5.8, restricted to 𝑁 = 3 case, satisfies all the properties stated
𝛱 𝑝+3
in the theorem. We have also shown that the operator in Theorem 5.4, restricted
grad
to 𝑁 = 3, satisfies all the stated properties of 𝛱 𝑝+3 except for its commutativity
curl
property involving 𝛱 𝑝+3 . Hence, to finish the proof, we now proceed to prove
curl 𝐸 defined in (5.65) satisfies the norm bound (5.41b), the moment
that the 𝛱 𝑝+3
conditions (5.45)–(5.46), as well as the commutativity properties
curl div
curl 𝛱 𝑝+3 𝐸 = 𝛱 𝑝+3 curl 𝐸, (5.66)
curl grad
𝛱 𝑝+3 grad 𝜙 = grad 𝛱 𝑝+3 𝜙 (5.67)
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 333
Example 5.12 (Maxwell equations). We continue Example 4.5, where the infin-
ite-dimensional spaces and forms for the primal DPG formulation of the Maxwell
cavity problem were set by (4.32), and the formulation was proved to be wellposed.
Now we focus on its discretization using subspaces 𝑋0,ℎ ⊂ 𝑋ℎ , 𝑋ˆ ℎ ⊂ 𝑋ˆ and 𝑌ℎ ⊂ 𝑌
set by
𝑋0,ℎ = {𝐸 ℎ ∈ 𝐻(curl,
˚ 𝛺) : 𝐸 ℎ | 𝐾 ∈ 𝑃 𝑝 (𝐾)3 for all 𝐾 ∈ 𝛺 ℎ }, (5.69a)
𝑋ˆ ℎ = {𝑛 × 𝐻ˆ ℎ ∈ 𝐻 −1/2 (divF , 𝜕𝛺 ℎ ) : 𝑛 × 𝐻ˆ ℎ | 𝜕𝐾 ∈ tr𝑛×
𝐾
𝑃 𝑝+1 (𝐾)3 for all 𝐾 ∈ 𝛺 ℎ },
𝑌ℎ = {𝐹ℎ ∈ 𝐻(curl, 𝛺 ℎ ) : 𝐹ℎ | 𝐾 ∈ 𝑁 𝑝+3 (𝐾) for all 𝐾 ∈ 𝛺 ℎ }. (5.69b)
To obtain error estimates, we apply Theorem 5.2, under the additional assumption
that the material coefficients 𝜇, 𝜀 are constant on each mesh element 𝐾 ∈ 𝛺 ℎ . Then
(5.41b), (5.45) and (5.46) of Theorem 5.9 verify condition (5.5) with 𝛱 = 𝛱 𝑝+3 curl .
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
334 L. Demkowicz and J. Gopalakrishnan
Bibliographical notes. The construction of the 𝐻(curl, 𝐾) Fortin operator for DPG
methods presented here is new, but is related to existing constructions. The first
𝐻(curl, 𝐾) Fortin operator was given in Carstensen et al. (2016). The construction
there uses an appropriate bubble space and is similar in spirit to our constructions
grad div in Theorem 5.8. Another natural method for
of 𝛱 𝑝+3 in Theorem 5.4 and 𝛱 𝑝+3
curl is through the constrained minimization (5.58), where the
construction of 𝛱 𝑝+3
required moment conditions are put as constraints. Such minimizations were used
to construct Fortin operators in Demkowicz (2024) and Demkowicz and Zanotti
(2020). The construction we have presented here is close but not identical to these,
because in proving Theorem 5.9 we needed to establish commutativity between
differently constructed Fortin operators. These techniques clearly show there are
multiple avenues to construct Fortin operators for DPG schemes.
which is characterized by
(𝑥 − 𝑥 ℎ , 𝑧 ℎ )𝑟 = 0 for all 𝑧 ℎ ∈ 𝑋ℎ . (6.3b)
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 335
This minimizer also minimizes a residual in a discrete dual norm and equals the
solution of the (practical) DPG method, as stated next.
Theorem 6.1 (Inexact residual minimization). Under the assumptions of The-
orem 5.2, the following are equivalent statements.
(a) 𝑥 ℎ ∈ 𝑋ℎ is the unique solution of the DPG method (5.4).
(b) 𝑥 ℎ is the unique element of 𝑋ℎ satisfying
|𝑥 − 𝑥 ℎ | 𝑟 = inf |𝑥 − 𝑧 ℎ | 𝑟 .
𝑧ℎ ∈𝑋ℎ
Proof. Follow along the lines of proof of Theorem 3.2 but using (6.1) instead of
(3.2) and noting (6.3).
Definition 6.2. Let ℓ be as in (1.1), let 𝑥˜ ℎ be any element of 𝑋ℎ , and let 𝑌 𝑟 be as
in (5.1). The element of 𝑌 𝑟 defined by
𝜀˜𝑟 = 𝑅𝑌−1𝑟 (ℓ − 𝐵𝑥˜ ℎ ) (6.4)
is called the inexact error representation of 𝑥˜ ℎ (see Definition 3.3). When 𝑥˜ ℎ is set
to the solution 𝑥 ℎ of the DPG method (5.4), then its inexact error representation is
denoted (omitting the tilde) by 𝜀𝑟 = 𝑅𝑌−1𝑟 (ℓ − 𝐵𝑥 ℎ ).
It is easy to see that 𝜀˜𝑟 in (6.4) is the unique element of 𝑌 𝑟 satisfying
(𝜀˜𝑟 , 𝑦)𝑌 = ℓ(𝑦) − 𝑏(𝑥˜ ℎ , 𝑦) for all 𝑦 ∈ 𝑌 𝑟 . (6.5)
This shows that the inexact error representation of the DPG solution, namely 𝜀𝑟 , is
𝑌 -orthogonal to the entire inexact optimal test space 𝑌ℎ𝑟 due to (5.4). Let
𝜂˜ = ∥ 𝜀˜𝑟 ∥𝑌 , 𝜂 = ∥𝜀𝑟 ∥𝑌 . (6.6)
Clearly, (6.1), (6.5) and (6.2) imply
𝜂˜ = ∥𝑅𝑌−1𝑟 𝐵(𝑥 − 𝑥˜ ℎ )∥𝑌 = ∥𝑇 𝑟 (𝑥 − 𝑥˜ ℎ )∥𝑌 .
When 𝑌 𝑟 is of the product form (5.1), the norm in (6.6) can be written in terms of
local element contributions, each of which acts as a practically computable element-
wise error indicator. It is useful to note the following analogue of Theorems 3.4.
Theorem 6.3 (Inexact error representation as a mixed solution component).
Let 𝜀𝑟 denote the inexact error representation of Definition 6.2. Then the following
are equivalent statements.
(a) 𝑥 ℎ ∈ 𝑋ℎ solves the DPG method (5.4).
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
336 L. Demkowicz and J. Gopalakrishnan
The mixed reformulation (6.7) of the DPG method in Theorem 6.3 gives further
insight into the stability of the method. In a typical two-equation mixed system,
enlarging the test space in the first equation, while often helpful to prove the inf-
sup condition by increasing the supremum, is fraught with the danger of losing
the coercivity of the first term. However, in the DPG system (6.7), the first term
(·, ·)𝑌 , being an inner product, will never lose coercivity, no matter how liberally
we enrich 𝑌 𝑟 . This explains any perceived ease in proving stability of DPG
formulations.
Theorem 6.4 (Global reliability and efficiency for any approximation). Under
the assumptions of Theorem 5.2, we have the following inequalities for the differ-
ence between the exact solution 𝑥 and any 𝑥˜ ℎ ∈ 𝑋ℎ in terms of the corresponding
computable error estimator 𝜂˜ of (6.6):
𝛾∥𝑥 − 𝑥˜ ℎ ∥ 𝑋 ≤ ∥𝛱 ∥ 𝜂˜ + osc(ℓ) (reliability), (6.8a)
𝜂˜ ≤ ∥𝑏∥ ∥𝑥 − 𝑥˜ ℎ ∥ 𝑋 (efficiency). (6.8b)
Here
osc(ℓ) = ∥ℓ ◦ (1 − 𝛱 )∥𝑌 ∗ (6.8c)
represents a term akin to data-approximation error and it admits the following
bound:
osc(ℓ) ≤ ∥𝑏∥ ∥1 − 𝛱 ∥ min ∥𝑥 − 𝑧 ℎ ∥ 𝑋 . (6.8d)
𝑧ℎ ∈𝑋ℎ
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 337
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
338 L. Demkowicz and J. Gopalakrishnan
follows from the Pythagoras theorem ∥𝜀∥𝑌2 = ∥𝜀𝑟 ∥𝑌2 + ∥𝜀 − 𝜀𝑟 ∥𝑌2 and
∥𝜀 − 𝜀𝑟 ∥𝑌2 = (𝜀, 𝜀 − 𝜀𝑟 )𝑌 by (6.12)
= ℓ(𝜀 − 𝜀𝑟 ) − 𝑏(𝑥 ℎ , 𝜀 − 𝜀𝑟 ) by (3.4a)
= ℓ(𝜀 − 𝜀𝑟 ) by (6.7b) and (3.4b).
Clearly
ℓ(𝜀 − 𝜀𝑟 ) = ℓ(𝜀 − 𝛱 𝜀) ≤ ∥ℓ ◦ (1 − 𝛱 )∥𝑌 ∗ ∥𝜀 − 𝜀𝑟 ∥𝑌 = osc(ℓ)∥𝜀 − 𝜀𝑟 ∥𝑌 ,
and the upper inequality of (6.12) follows.
Bibliographical notes. The proof given for Theorem 6.4 is a slightly simpli-
fied version of the original presented in Carstensen, Demkowicz and Gopala-
krishnan (2014a, Theorem 2.1), and produces an improved reliability constant as
in Carstensen, Gallistl, Hellwig and Weggler (2014b, Lemma 3.6). If in addition
𝛱 is idempotent, then (6.8a) can be further improved to
√︁ 2
𝛾 2 ∥𝑥 − 𝑥 ℎ ∥ 2𝑋 ≤ 𝜂2 + 𝜂 ∥𝛱 ∥ 2 − 1 + osc(ℓ) , (6.13)
as shown in Keith, Astaneh and Demkowicz (2019, Theorem 6.4). Note the
relationship between (6.12) and (6.13) when 𝛱 is an orthogonal projection. It
is easy to construct adaptive algorithms with marking strategies based on the
DPG error estimator following the usual ‘Solve → Estimate → Mark → Refine’
paradigm. In all reports of practical performance (Demkowicz et al. 2012a, Petrides
and Demkowicz 2017), such DPG algorithms work very well, but to the best of our
knowledge their convergence and optimality are yet to be rigorously proved.
7. Ultraweak formulations
A rich set of examples to apply the DPG ideas is offered by the so-called ‘ultraweak’
formulations of boundary value problems seeking 𝑢 ∈ 𝑉 satisfying 𝐴𝑢 = 𝑓 for
an 𝑓 ∈ 𝐿 2 (𝛺)𝑚 . Here 𝑉 is a space where homogeneous boundary conditions
are imposed, and 𝐴 is a general partial differential operator (specified below). In
ultraweak formulations all derivatives in 𝐴 are moved to test functions by integration
by parts, element by element. In order to use the previously developed DPG ideas, it
is important to obtain a reformulation where the trial-to-test operator 𝑇 is localized,
i.e. a formulation where the test space has the form (4.1). Such a reformulation,
set in a broken graph space, is derived and studied in this section. We prove its
wellposedness by general arguments that cover many examples at once. The first
main result (Theorem 7.6) of this section identifies conditions under which the
wellposedness of ultraweak formulations in broken graph spaces can be obtained
as soon as 𝐴 : 𝑉 → 𝐿 2 (𝛺)𝑚 is a bijection, no matter how complex the spaces of
interface variables are. Another result of this section (Theorem 7.9), which has not
appeared in previous literature, exhibits norms in which the best possible stability
of ultraweak formulations can be obtained.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 339
For any 𝑢 ∈ dom(𝐴), since 𝐴𝑢 is in 𝐿 2 (𝛺)𝑚 , the left-hand side above equals
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
340 L. Demkowicz and J. Gopalakrishnan
(𝐴𝑢, 𝜑)
˜ 𝛺 . Hence the condition in (7.5) is verified with 𝜑˜ in place of 𝑔, that is,
D(𝛺)𝑚 ⊆ dom(𝐴∗ ). (7.8)
In view of (7.4), (7.8) and (7.6), we can now identify 𝐴∗ with the formal adjoint
partial differential operator
𝑚
∑︁ ∑︁
∗
𝐴 𝜑˜ = 𝑒 𝑖 (−1) | 𝛼| 𝑎 𝑗𝑖 𝛼 𝜕 𝛼 𝜑˜ 𝑗 . (7.9)
𝑖, 𝑗=1 𝛼∈I 𝑁
𝑘
Here ∥ · ∥ 𝑆 denotes the norm of 𝐿 2 (𝑆)𝑚 ; the corresponding inner product is denoted
(·, ·)𝑆 . Our assumptions imply that these inner product spaces are complete.
Lemma 7.1. The spaces 𝑊(𝑆) and 𝑊(𝑆)
˜ are Hilbert spaces.
Proof. Consider a Cauchy sequence 𝑢 𝑛 in 𝑊(𝑆). Clearly, 𝑢 𝑛 is Cauchy in 𝐿 2 (𝑆)𝑚
and 𝐴𝑢 𝑛 is Cauchy in 𝐿 2 (𝑆)𝑚 . Hence there is a 𝑢 ∈ 𝐿 2 (𝑆)𝑚 and 𝑓 ∈ 𝐿 2 (𝑆)𝑚
such that ∥𝑢 − 𝑢 𝑛 ∥ 𝑆 → 0 and ∥ 𝑓 − 𝐴𝑢 𝑛 ∥ 𝑆 → 0. To show that 𝑢 is in 𝑊(𝑆),
we use (7.7), a consequence of assumption (7.3), to get that for any 𝜑˜ ∈ D(𝑆)𝑚 ,
˜ = (𝑢 𝑛 , 𝐴∗ 𝜑)
(𝐴𝑢 𝑛 )(𝜑) ˜ 𝑆 → (𝑢, 𝐴∗ 𝜑) ˜ as 𝑛 → ∞. Since 𝐴𝑢 𝑛 → 𝑓 in
˜ 𝑆 = (𝐴𝑢)(𝜑)
𝐿 (𝛺) , this implies that the distribution 𝐴𝑢 must equal 𝑓 in 𝐿 2 (𝛺)𝑚 . This proves
2 𝑚
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 341
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
342 L. Demkowicz and J. Gopalakrishnan
(𝐴∗ 𝑤)(𝜑)
˜ = (𝐴𝜑, 𝑤) ˜ 𝛺 = (𝜑, 𝐴∗ℎ 𝑤)
˜ 𝛺 = (𝐴ℎ 𝜑, 𝑤) ˜ 𝛺 + ⟨𝐷 ℎ 𝜑, 𝑤⟩
˜ 𝑊˜ ℎ .
The last term must vanish because of the first equality of (7.19) and because
𝜑 ∈ D(𝛺)𝑚 ⊆ dom(𝐴) = 𝑉 by our assumption (7.4). Hence
|(𝐴∗ 𝑤)(𝜑)|
˜ ≤ ∥𝜑∥ 𝛺 ∥ 𝐴∗ℎ 𝑤∥
˜ 𝛺 for all 𝜑 ∈ D(𝛺)𝑚 ,
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 343
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
344 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 345
Proof. We use two functions, 𝑢˜ 𝑞ˆ and 𝑢 𝑞ˆ , both obtained from 𝑞, ˆ but by solving
two distinct boundary value problems. First, the supremum of the lemma, which
we denote by 𝑠, is attained by the function 𝑢˜ 𝑞ˆ in 𝑊˜ ℎ satisfying
(𝐴∗ℎ 𝑢˜ 𝑞ˆ , 𝐴∗ℎ 𝑤)
˜ 𝛺 + (𝑢˜ 𝑞ˆ , 𝑤)
˜ 𝛺 = −⟨𝑞,
ˆ 𝑤⟩
˜ 𝑊˜ ℎ (7.25)
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
346 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 347
and (7.31) follows with 𝑐 0 = 1/(𝑐2 + 1)1/2 . Hence, applying Theorem 4.3, the proof
is complete.
Example 7.7 (Schrödinger equation). This is an example of a second-order op-
erator for which the previous theory applies. Let 𝜕𝑥 𝑥 denote the Laplacian with
respect to a spatial variable 0 < 𝑥 < 𝐿, let 𝜕𝑡 denote the derivative 𝜕/𝜕𝑡 with
respect to 0 < 𝑡 < 𝑇 (where both 𝐿 and 𝑇 are finite), let 𝛺 = (0, 𝐿) × (0, 𝑇),
and let 𝑓 ∈ 𝐿 2 (𝛺). The classical form of the Schrödinger initial boundary value
problem is
𝚤ˆ𝜕𝑡 𝑢 − 𝜕𝑥 𝑥 𝑢 = 𝑓 , 0 < 𝑥 < 𝐿, 0 < 𝑡 < 𝑇, (7.34a)
𝑢(𝑥, 𝑡) = 0, 𝑥 = 0 or 𝑥 = 𝐿, 0 < 𝑡 < 𝑇, (7.34b)
𝑢(𝑥, 0) = 0, 0 < 𝑥 < 𝐿. (7.34c)
Here 𝑓 is any given function in 𝐿 2 (𝛺). Viewing 𝛺 as a rectangle with time as
the vertical axis, let 𝛤 denote the union of vertical boundary walls and the bottom
initial time slice, and let 𝛤˜ denote the union of vertical boundary walls and the top
final time slice. Then the initial and boundary conditions together can be written
as 𝑢| 𝛤 = 0.
To fit into the previous framework, set
𝑘 = 2, 𝑚 = 1, 𝐴 = 𝚤ˆ𝜕𝑡 − 𝜕𝑥 𝑥 .
Then the formal adjoint expression in (7.9) reads 𝐴∗ = 𝚤ˆ𝜕𝑡 − 𝜕𝑥 𝑥 = 𝐴. Hence the
graph spaces are
𝑊 = 𝑊˜ = {𝑢 ∈ 𝐿 2 (𝛺) : 𝑖𝜕𝑡 𝑢 − Δ 𝑥 𝑢 ∈ 𝐿 2 (𝛺)},
and the boundary operator 𝐷˜ = 𝐷 : 𝑊 → 𝑊 ∗ is set by ⟨𝐷𝑤, 𝑣⟩𝑊 = (𝐴𝑤, 𝑣)𝛺 −
(𝑤, 𝐴𝑣)𝛺 for all 𝑤, 𝑣 ∈ 𝑊. As usual, let D( 𝛺)
¯ denote the restrictions of functions
from D(R 𝑁 ) to 𝛺. Integration by parts shows that
∫ ∫ ∫
⟨𝐷𝜙, 𝜓⟩𝑊 = 𝚤ˆ𝑛𝑡 𝜙𝜓¯ + 𝜙𝑛 𝑥 𝜕𝑥 𝜓¯ − 𝑛 𝑥 𝜕𝑥 𝜙𝜓¯ (7.35)
𝜕𝛺 𝜕𝛺 𝜕𝛺
for all 𝜙, 𝜓 ∈ D( 𝛺),¯ where we have used the spatial and temporal components
𝑛 𝑥 , 𝑛𝑡 of the outward unit normal 𝑛 on 𝜕𝛺.
To incorporate the boundary and initial conditions into dom(𝐴), circumventing
the development of a full trace theory for the graph space, we first set
Ṽ = {𝜑 ∈ D( 𝛺)
¯ : 𝜑| 𝛤˜ = 0},
and use it to set
dom(𝐴) = {𝑢 ∈ 𝑊 : ⟨𝐷𝑣, 𝑢⟩𝑊 = 0 for all 𝑣 ∈ Ṽ }, (7.36)
or equivalently
⊥
𝑉= 𝐷 Ṽ . (7.37)
For smooth 𝑢 ∈ D( 𝛺)
¯ ∩ dom(𝐴), the integration-by-parts formula (7.35) shows that
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
348 L. Demkowicz and J. Gopalakrishnan
𝑢| 𝛤 must vanish. Note that assumptions (7.3), (7.4) and (7.10) are immediately
verified. By Lemma 7.2, the domain of the maximal adjoint is given by
⊥
𝑉˜ = 𝐷𝑉 . (7.38)
It is shown in Demkowicz et al. (2017, Theorem 3.1) that Ṽ is dense in 𝑉. ˜ Hence
⊥
(7.37) implies 𝑉 = 𝐷𝑉˜ and assumption (7.14) is also verified.
To conclude wellposedness of the ultraweak formulation (7.22) for this Schrö-
dinger problem, the only remaining assumption we need to verify is the bijectivity
stated in (7.30). Let 𝜙 𝑘 (𝑥) in 𝐻01 (0, 𝐿) and let 𝜆 𝑘 > 0 be a Laplace eigenpair
satisfying −𝜕𝑥 𝑥 𝜙 𝑘 = 𝜆 𝑘 𝜙 𝑘 normalized so that ∥𝜙 𝑘 ∥ (0,𝐿) = 1 for all natural numbers
𝑘 ≥ 1. Suppose 𝑓 ∈ 𝐿 2 (𝛺) and
∫ 𝐿 ∫ 𝑡
𝑓 𝑘 (𝑡) = 𝑓 (𝑥, 𝑡)𝜙¯𝑘 (𝑥) 𝑑𝑥, 𝑢 𝑘 (𝑡) = −ˆ𝚤 𝑒 𝚤ˆ𝜆𝑘 (𝑡 −𝑠) 𝑓 𝑘 (𝑠) 𝑑𝑠, (7.39a)
0 0
𝑀
∑︁ 𝑀
∑︁
𝐹𝑀 (𝑥, 𝑡) = 𝑓 𝑘 (𝑡)𝜙 𝑘 (𝑥), 𝑈 𝑀 (𝑥, 𝑡) = 𝑢 𝑘 (𝑡)𝜙 𝑘 (𝑥). (7.39b)
𝑘=1 𝑘=1
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 349
𝑞 + grad 𝑢 = 0 in 𝛺,
div 𝑞 = 𝑓 in 𝛺,
𝑢=0 on 𝜕𝛺.
We easily see that the adjoint operator, acting on 𝑣˜ = (𝑞, ˜ ∈ 𝐿 2 (𝛺) 𝑁 × 𝐿 2 (𝛺), is
˜ 𝑢)
Clearly assumptions (7.3), (7.4) and (7.10) hold for this example. Since 𝑉 = 𝑉˜
and 𝐷 = 𝐷,˜ the assumption (7.14) also holds since it is the same as the conclusion
(7.13) of Lemma 7.2.
Note that if (𝑞, 𝑢) and 𝐴(𝑞, 𝑢) are both in 𝐿 2 (𝛺)𝑚 , then obviously div 𝑞 ∈ 𝐿 2 (𝛺)
and grad 𝑢 ∈ 𝐿 2 (𝛺) 𝑁 , so
𝑊 = 𝑊˜ = 𝐻(div, 𝛺) × 𝐻 1 (𝛺),
𝑊ℎ = 𝑊˜ ℎ = 𝐻(div, 𝛺 ℎ ) × 𝐻 1 (𝛺 ℎ ),
using the broken Sobolev spaces defined in (4.4) and (4.33). Hence, for any
(𝑞, 𝑢) ∈ 𝑊ℎ and (𝑞,˜ 𝑢)˜ ∈ 𝑊˜ ℎ ,
∑︁
⟨𝐷 ℎ (𝑞, 𝑢), (𝑞, ˜ 𝑊˜ =
˜ 𝑢)⟩ ⟨𝑞 · 𝑛, 𝑢⟩
˜ 𝐻 1/2 (𝜕𝐾) + ⟨𝑢, 𝑞˜ · 𝑛⟩ 𝐻 −1/2 (𝜕𝐾)
𝐾 ∈𝛺ℎ
≡ ⟨𝑞 · 𝑛, 𝑢⟩
˜ ℎ + ⟨𝑢, 𝑞˜ · 𝑛⟩ℎ , (7.40)
where in the last step we have extended the previous notation of (4.8) ⟨·, ·⟩ℎ
to include sums of duality pairings in both 𝐻 1/2 (𝜕𝐾) 𝐻 −1/2 (𝜕𝐾). Since any
(𝑞, 𝑢) ∈ 𝑉 = dom(𝐴) satisfies 𝑢| 𝜕𝛺 = 0 on the global boundary, its element-by-
element trace tr(𝑢), as defined in (4.34), lies in
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
350 L. Demkowicz and J. Gopalakrishnan
The (hybrid) ultraweak formulation (7.22) now takes the form (1.1) with the fol-
lowing forms:
ˆ 𝑞ˆ 𝑛 ), (𝑟, 𝑣) ) = (𝑞, 𝑟)ℎ − (𝑢, div 𝑟)ℎ + ⟨𝑢,
𝑏( (𝑞, 𝑢, 𝑢, ˆ 𝑟 · 𝑛⟩ℎ
− (𝑞, grad 𝑣)ℎ + ⟨𝑣, 𝑞ˆ 𝑛 ⟩ℎ ,
ℓ(𝑟, 𝑣) = ( 𝑓 , 𝑣)𝛺 ,
where (𝑞, 𝑢) ∈ 𝑊ℎ , (𝑟, 𝑣) ∈ 𝑊˜ ℎ , 𝑢ˆ ∈ 𝐻˚ 1/2 (𝜕𝛺 ℎ ) and 𝑞ˆ 𝑛 ∈ 𝐻 −1/2 (𝜕𝛺 ℎ ).
By Theorem 7.6, this ultraweak formulation is wellposed if 𝐴 : 𝑉 → 𝐿 2 (𝛺) 𝑁 +1
is a bijection, that is, if there is a unique 𝑞 ∈ 𝐻(div, 𝛺) and 𝑢 ∈ 𝐻˚ 1 (𝛺) satisfying
𝑞 + grad 𝑢 = 𝐺 on 𝛺, (7.41a)
div 𝑞 = 𝐹 on 𝛺 (7.41b)
for any given 𝐹 ∈ 𝐿 2 (𝛺) and 𝐺 ∈ 𝐿 2 (𝛺) 𝑁 . To verify this condition, it is sufficient
to note that 𝑞 and 𝑢 satisfy (7.41) if and only if they form the unique solution
of the well-known mixed weak problem (see e.g. Brezzi and Fortin 1991, Ch. II,
Prop. 1.3) to find 𝑞 in 𝐻(div, 𝛺) and 𝑢 in 𝐿 2 (𝛺) such that
(𝑞, 𝑟)𝛺 − (𝑢, div 𝑟)𝛺 = (𝐺, 𝑟)𝛺 for all 𝑟 ∈ 𝐻(div, 𝛺), (7.42a)
2
(div 𝑞, 𝑤)𝛺 = (𝐹, 𝑤)𝛺 for all 𝑤 ∈ 𝐿 (𝛺). (7.42b)
It is easy to see that (7.42a) also implies that 𝑢 ∈ 𝐻˚ 1 (𝛺) and (7.41a) holds. Hence
the unique solution (𝑞, 𝑢) of (7.42) is in 𝑉 and solves 𝐴(𝑞, 𝑢) = (𝐺, 𝐹), thus
verifying assumption (7.30).
Bibliographical notes. Theorem 7.6 and the treatment of the Schrödinger equation
(Example 7.7) by DPG methods appeared first in Demkowicz et al. (2017). There
it is also pointed out why it is not advisable to split the Schrödinger equation
into a first-order system. This is the reason for staying with the original second-
order form of the Schrödinger equation while deriving the ultraweak formulation in
Example 7.7. The wellposedness result in Example 7.8 was first proved in Demko-
wicz and Gopalakrishnan (2011a), but using different techniques. An application
of Theorem 7.6 to the spacetime wave equation can be found in Gopalakrishnan
and Sepúlveda (2019). That paper also notes how the spacetime wave operator
produces a 𝐷 ℎ operator with a non-trivial null space (even though 𝑞ˆ = 𝐷 ℎ 𝑢 can
be uniquely determined) and how one overcomes the consequent difficulties in
practically solving an ultraweak discretization.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 351
Proof. We need only prove the second equality in (7.43). The first equality of
(7.43) then follows from the second by Proposition 3.6, and moreover, (7.44) then
follows from Theorem 3.2(b).
Let 𝑤˜ ∈ 𝑊˜ ℎ . We will produce a (𝑤, 𝑟)
ˆ ∈ 𝑉 × 𝑄 satisfying
ˆ 𝑤)
𝑏((𝑤, 𝑟), ˜ 𝑌2 ,𝑠 ,
˜ = ∥ 𝑤∥ ∥(𝑤, 𝑟)∥
ˆ 𝑋,𝑠 ≤ ∥ 𝑤∥
˜ 𝑌 ,𝑠 . (7.45)
By virtue of (7.30), there is a 𝑧 ∈ 𝑉 such that
𝐴𝑧 = 𝑤.
˜ (7.46)
Then
˜ 𝑌2 ,𝑠 = (𝑠 −2 𝑤, ˜ 𝛺 + 𝐴∗ℎ 𝑤,
˜ 𝐴∗ℎ 𝑤˜
∥ 𝑤∥ ˜ 𝑤) 𝛺
= (𝐴(𝑠 −2 𝑧), 𝑤)
˜ 𝛺 + 𝐴∗ℎ 𝑤,
˜ 𝐴∗ℎ 𝑤˜ 𝛺
= 𝑠 −2 𝑧, 𝐴∗ℎ 𝑤˜ 𝛺 + ⟨𝐷 ℎ (𝑠 −2 𝑧), 𝑤⟩
˜ 𝑊˜ ℎ + 𝐴∗ℎ 𝑤,
˜ 𝐴∗ℎ 𝑤˜ 𝛺
= 𝑏((𝑤, 𝑟),
ˆ 𝑤)˜
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
352 L. Demkowicz and J. Gopalakrishnan
with
𝑤 = 𝑠 −2 𝑧 + 𝐴∗ℎ 𝑤˜ ∈ 𝐿 2 (𝛺)𝑚 , 𝑟ˆ = 𝐷 ℎ (𝑠2 𝑧) ∈ 𝑄. (7.47)
Moreover,
ˆ 2𝑋,𝑠 =
∥(𝑤, 𝑟)∥ inf ∥𝑤 − 𝑣∥ 2𝛺 + 𝑠2 ∥ 𝐴𝑣∥ 2𝛺
−1 { 𝑞}
𝑣 ∈𝐷ℎ,𝑉 ˆ
where we have used the formulas for 𝑤 and 𝑧, from (7.47) and (7.46) respectively,
in the last step. This proves (7.45), from which it readily follows that
|𝑏((𝑣, 𝑣ˆ ), 𝑤)|
˜ |𝑏((𝑤, 𝑟),
ˆ 𝑤)|
˜
sup ≥ ≥ ∥ 𝑤∥
˜ 𝑌 ,𝑠 . (7.48)
0≠(𝑣, 𝑣)∈𝑋
ˆ ∥(𝑣, 𝑣ˆ )∥ 𝑋,𝑠 ∥(𝑤, 𝑟)∥
ˆ 𝑋,𝑠
In fact the supremum equals ∥ 𝑤∥˜ 𝑌 ,𝑠 because the reverse inequality also holds,
as we now show. Letting (𝑣, 𝑣ˆ ) be any element in 𝑋 and choosing any 𝑧 ∈ 𝑉 such
that 𝑣ˆ = 𝐷 ℎ 𝑧,
˜ = (𝑣, 𝐴∗ℎ 𝑤)
𝑏((𝑣, 𝑣ˆ ), 𝑤) ˜ 𝛺 + ⟨𝐷 ℎ 𝑧, 𝑤⟩
˜ 𝑊˜ ℎ
= (𝑣, 𝐴∗ℎ 𝑤) ˜ 𝛺 − (𝑧, 𝐴∗ℎ 𝑤)
˜ 𝛺 + (𝐴𝑧, 𝑤) ˜ 𝛺
∗
= (𝑣 − 𝑧, 𝐴ℎ 𝑤)
˜ 𝛺 + (𝐴𝑧, 𝑤)
˜ 𝛺
1/2
≤ ∥𝑣 − 𝑧∥ 2𝛺 + 𝑠2 ∥ 𝐴𝑧∥ 2𝛺 ∥ 𝑤∥
˜ 𝑌 ,𝑠 .
−1 { 𝑣ˆ }, we obtain
Taking the infimum over all 𝑧 ∈ 𝐷 ℎ,𝑉
˜ ≤ ∥(𝑣, 𝑣ˆ )∥ 𝑋,𝑠 ∥ 𝑤∥
𝑏((𝑣, 𝑣ˆ ), 𝑤) ˜ 𝑌 ,𝑠 ,
The trial norm ∥ · ∥ 𝑋,𝑠 of Theorem 7.9 is related to Peetre’s 𝐾-functional (Bergh
and Löfström 1976). To see this, suppose there is a 𝐶𝑉 > 0 such that
Assumption (7.30) certainly implies the existence of such a 𝐶𝑉 (by the Closed
Range Theorem). Let 𝑉0 = {𝑤 ℎ ∈ 𝑊ℎ : 𝐷 ℎ 𝑤 ℎ = 0} be the kernel of 𝐷 ℎ , which is
a closed subspace by the continuity of 𝐷 ℎ . By Lemma 7.4, 𝑉0 is a subspace of 𝑉.
Hence, for any 𝑢ˆ ∈ 𝑄, the set 𝐷 ℎ,𝑉−1 { 𝑢}
ˆ equals the affine translate 𝑣 𝑢ˆ + 𝑉0 for any
𝑣 𝑢ˆ ∈ 𝑉 with the property 𝐷 ℎ 𝑣 𝑢ˆ = 𝑢.
ˆ Minimization over this closed coset gives a
minimal extension 𝐸 𝑢ˆ of 𝑢ˆ defined by
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 353
−1 { 𝑢}
Note that by (7.49), ∥ 𝐴𝑣∥ 𝛺 is a norm on 𝐷 ℎ,𝑉 ˆ ⊂ 𝑉. Since 𝐸 is defined through
minimization over a translate of 𝑉0 , we see that
ˆ 𝐴𝑣 0 )𝛺 = 0
(𝐴𝐸 𝑢, for all 𝑣 0 ∈ 𝑉0 . (7.50)
Define the 𝐾-functional for the scale of spaces between 𝑉0 and 𝐿 2 (𝛺)𝑚 by
𝐾(𝑠, 𝑤) = inf ∥𝑤 − 𝑣 0 ∥ 2𝛺 + 𝑠2 ∥ 𝐴𝑣 0 ∥ 2𝛺 (7.51)
𝑣0 ∈𝑉0
for any 𝑤 ∈ 𝐿 2 (𝛺)𝑚 . The next two results help us better understand the norm
∥ · ∥ 𝑋,𝑠 in Theorem 7.9.
ˆ ∈ 𝑋,
Proposition 7.10. Suppose (7.49) holds. Then, for any (𝑢, 𝑢)
ˆ 2𝑋,𝑠 = 𝑠2 ∥ 𝐴𝐸 𝑢∥
∥(𝑢, 𝑢)∥ ˆ 2𝛺 + 𝐾(𝑠, 𝑢 − 𝐸 𝑢).
ˆ
−1 { 𝑢}
Proof. Writing any 𝑣 ∈ 𝐷 ℎ,𝑉 ˆ as 𝑣 = 𝐸 𝑢ˆ + 𝑣 0 for a 𝑣 0 ∈ 𝑉0 ,
∥𝑢 − 𝑣∥ 2𝛺 + 𝑠2 ∥ 𝐴𝑣∥ 2𝛺 = ∥𝑢 − 𝐸 𝑢ˆ − 𝑣 0 ∥ 2𝛺 + 𝑠2 ∥ 𝐴(𝐸 𝑢ˆ + 𝑣 0 )∥ 2𝛺
= ∥𝑢 − 𝐸 𝑢ˆ − 𝑣 0 ∥ 2𝛺 + 𝑠2 ∥ 𝐴𝐸 𝑢∥
ˆ 2𝛺 + 𝑠2 ∥ 𝐴𝑣 0 ∥ 2𝛺 ,
where the last equality is due to (7.50). Hence the result follows by minimizing
over 𝑣 0 ∈ 𝑉0 .
√︁
Proposition 7.11. Let 𝐶𝑉 be as in (7.49), 𝑐 𝑠 = 𝐶𝑉2 /𝑠2 , and 𝑘 𝑠 = 12 (𝑐 𝑠 + 𝑐2𝑠 + 4𝑐 𝑠 ).
ˆ ∈ 𝑋 and 𝑠 > 0, we have these two-sided bounds:
Then, for all (𝑢, 𝑢)
(1 + 𝑘 𝑠 )−1 ∥(𝑢, 𝑢)∥
ˆ 2𝑋,𝑠 ≤ ∥𝑢∥ 2𝛺 + 𝑠2 ∥ 𝐴𝐸 𝑢∥
ˆ 2𝛺 ≤ (1 + 𝑘 𝑠 ) ∥(𝑢, 𝑢)∥
ˆ 2𝑋,𝑠 . (7.52)
Proof. By the triangle inequality,
∥𝑢∥ 2𝛺 ≤ (∥𝑢 − 𝐸 𝑢ˆ − 𝑣 0 ∥ 𝛺 + ∥𝐸 𝑢ˆ + 𝑣 0 ∥ 𝛺 )2
≤ (1 + 𝛼 −2 )∥𝑢 − 𝐸 𝑢ˆ − 𝑣 0 ∥ 2𝛺 + (1 + 𝛼2 )𝐶𝑉2 ∥ 𝐴(𝐸 𝑢ˆ + 𝑣 0 )∥ 2𝛺 ,
where we have used (7.49) and the inequality (𝑎 + 𝑏)2 ≤ (1 + 𝛼 −2 )𝑎 2 + (1 + 𝛼2 )𝑏 2
for numbers 𝑎, 𝑏 and 𝛼 > 0. Using (7.50),
ˆ 2𝛺 ≤ (1 + 𝛼 −2 )∥𝑢 − 𝐸 𝑢ˆ − 𝑣 0 ∥ 2𝛺
∥𝑢∥ 2𝛺 + 𝑠2 ∥ 𝐴𝐸 𝑢∥
+ [(1 + 𝛼2 )𝑐 𝑠 + 1] 𝑠2 ∥ 𝐴𝐸 𝑢∥
ˆ 2𝛺 + 𝑠2 ∥ 𝐴𝑣 0 ∥ 2𝛺
√︁
with 𝑐 𝑠 = 𝐶𝑉2 /𝑠2 . Now set 𝛼2 = 12 (−𝑐 𝑠 + 𝑐2𝑠 + 4𝑐 𝑠 )/𝑐 𝑠 so that (1 + 𝛼2 )𝑐 𝑠 = 𝛼 −2 .
Then 1 + 𝛼 −2 = 1 + (1 + 𝛼2 )𝑐 𝑠 = 1 + 𝑘 𝑠 and the last inequality of (7.52) follows
after taking the infimum over all 𝑣 0 ∈ 𝑉0 and applying Proposition 7.10.
For the first inequality of (7.52), we begin by noting that the choice of 𝑣 0 = 0 in
(7.51) gives 𝐾(𝑠, 𝑤) ≤ ∥𝑤∥ 2𝛺 . Together with Proposition 7.10, we then have
ˆ 𝑋,𝑠 ≤ 𝑠2 ∥ 𝐴𝐸 𝑢∥
∥(𝑢, 𝑢)∥ ˆ 2𝛺 + ∥𝑢 − 𝐸 𝑢∥
ˆ 2𝛺 .
By the triangle inequality and (7.49),
ˆ 𝑋,𝑠 ≤ (1 + 𝛼 −2 )∥𝑢∥ 2𝛺 + [(1 + 𝛼 −2 )𝑐 𝑠 + 1]𝑠2 ∥ 𝐴𝐸 𝑢∥
∥(𝑢, 𝑢)∥ ˆ 2𝛺 .
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
354 L. Demkowicz and J. Gopalakrishnan
where the constant 𝑘 𝑠 is as in Proposition 7.11. At the price of increasing the quasi-
optimality constant from the optimal one, this estimate gives a simpler implication
of (7.44) in easier norms.
Example 7.12 (Helmholtz equation for time-harmonic waves). The Helmholtz
equation arises in varied applications, including electromagnetics and acoustics.
For example, in the latter, the physics of acoustical disturbances (Courant and
Friedrichs 1948) show that by linearizing the isentropic Euler equations around a
hydrostatic solution and assuming harmonic time variations, we obtain
𝚤ˆ𝜔𝑣 + grad 𝜙 = 𝐺 in 𝛺, (7.54a)
𝚤ˆ𝜔𝜙 + div 𝑣 = 𝐹 in 𝛺, (7.54b)
for some given 𝜔 > 0, 𝐹 ∈ 𝐿 2 (𝛺) and 𝐺 ∈ 𝐿 2 (𝛺) 𝑁 . Here 𝑣 : 𝛺 → C 𝑁 and
𝜙 : 𝛺 → C are velocity and pressure variables, respectively, associated to the
acoustic perturbations from equilibrium, complexified under the standard time-
harmonic assumption. These equations must be supplemented by a boundary
condition. Let us consider the impedance boundary condition
𝑣·𝑛−𝜙=0 on 𝜕𝛺. (7.54c)
Other Dirichlet, Neumann or mixed-type boundary conditions can equally well be
considered. Note that taking the divergence of (7.54a) and substituting the value
of div 𝑣 from (7.54b), we recover the popular second-order form of the Helmholtz
equation for 𝜙 (which we shall not use here).
The first-order system (7.54) can be written as 𝐴𝑢 = 𝑓 using the group variable
𝑢 = (𝑣, 𝜙) ∈ 𝐿 2 (𝛺) 𝑁 × 𝐿 2 (𝛺), the unbounded operator
𝐴𝑢 = (ˆ𝚤 𝜔𝑣 + grad 𝜙, 𝚤ˆ𝜔𝜙 + div 𝑣)
and 𝑓 = (𝐺, 𝐹) ∈ 𝐿 2 (𝛺) 𝑁 × 𝐿 2 (𝛺). Clearly (7.54) is in the setting of (7.2) with
𝑚 = 𝑁 + 1 and dom 𝐴 equal to
𝑉 = {(𝑧, 𝜇) ∈ 𝐻(div, 𝛺) × 𝐻 1 (𝛺) : 𝑧 · 𝑛 = 𝜇 on 𝜕𝛺}.
Its adjoint is
𝐴∗ 𝑢˜ = (−ˆ𝚤 𝜔𝑣˜ − grad 𝜙,
˜ −ˆ𝚤 𝜔 𝜙˜ − div 𝑣˜ )
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 355
This estimate was arrived at by other means (without using Theorem 7.9) in Gopala-
krishnan, Muga and Olivares (2014). There it was offered as a justification for the
practically visible marked improvement in DPG Helmholtz solutions as 𝛿 is made
smaller. The improvement in solutions was further justified there via a dispersion
analysis of the DPG method for the Helmholtz equation on an infinite uniform
stencil. Since the test norm
∥ 𝑤∥ ˜ 2𝛺 + ∥ 𝐴∗ℎ 𝑤∥
˜ 𝑌2 ,1/ 𝛿 = 𝛿2 ∥ 𝑤∥ ˜ 2𝛺
becomes smaller as 𝛿 is made smaller, a takeaway from such observations is that it
pays to use a weaker norm on the test space when computing wave solutions.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
356 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 357
We then set the norm on 𝑌 = 𝐻 1 (𝛺)𝑚 to be the optimal test norm, that is,
∥𝑦∥𝑌 = ||||𝑦||||𝑌 . (8.4b)
Clearly, for any 𝑣, 𝑦 ∈ 𝑌 ,
(𝑣, 𝑦)𝑌 = (𝐴∗ 𝑣, 𝐴∗ 𝑦)𝛺 + 𝑣(1) · 𝑦(1)
= (−𝑣¤ + 𝐾 ∗ 𝑣, − 𝑦¤ + 𝐾 ∗ 𝑦)𝛺 + 𝑣(1) · 𝑦(1) (8.4c)
is the inner product that generates the optimal test norm above. Here 𝑣¤ = 𝑑𝑣/𝑑𝑡.
This is one of the rare cases where the optimal test norm is so readily calculable.
The ideal Petrov Galerkin method (i.e. the IPG method of Definition 2.3) uses the
optimal test space, which we now examine. Using the inner product in (8.4c), the
ˆ ∈𝑋
variational problem for the optimal test function 𝑣 corresponding to any (𝑤, 𝑤)
reads as follows for any 𝑦 ∈ 𝐻 1 (𝛺)𝑚 :
(−𝑣¤ + 𝐾 ∗ 𝑣, − 𝑦¤ + 𝐾 ∗ 𝑦)𝛺 + 𝑣(1) · 𝑦(1) = (𝑤, − 𝑦¤ + 𝐾 ∗ 𝑦)𝛺 + 𝑤ˆ 𝑦(1). (8.5)
For any 𝑔 ∈ 𝐿 2 (𝛺), since the initial value problem 𝑦¤ − 𝐾 ∗ 𝑦 = −𝑔 with initial
condition 𝑦(1) = 0 is solvable, we find that (8.5) implies (−𝑣¤ + 𝐾 ∗ 𝑣, 𝑔)𝛺 = (𝑤, 𝑔)𝛺
for all 𝑔 in 𝐿 2 (𝛺), i.e. −𝑣¤ + 𝐾 ∗ 𝑣 = 𝑤. Using this in (8.5), we then conclude that
𝑣(1) = 𝑤.ˆ Thus the optimal test function 𝑣 of any (𝑤, 𝑤) ˆ ∈ 𝑋 is the solution of
𝑑𝑣
− + 𝐾𝑣 = 𝑤 in 𝛺, (8.6a)
𝑑𝑡
𝑣(1) = 𝑤.
ˆ (8.6b)
Recall the matrix exponential, defined by
∞
∑︁ 1 𝑘
𝑒𝐴 = 𝐴 . (8.7)
𝑘=0
𝑘!
Using it, the solution of (8.6) can be written down in closed form by the variation of
constants method. Namely, the optimal test function 𝑣 and the trial-to-test operator
𝑇 are given by
∫ 𝑡
𝐾 ∗ (𝑡 −1) 𝐾 ∗𝑡 ∗
𝑣(𝑡) = 𝑇(𝑤, 𝑤)
ˆ =𝑒 𝑤ˆ + 𝑒 𝑒 −𝐾 𝜏 𝑤(𝜏) 𝑑𝜏. (8.8)
1
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
358 L. Demkowicz and J. Gopalakrishnan
It is easy to see that the infimum equals ∥𝑢 − 𝛱𝑈 𝑢∥ 𝛺 . Hence the identities of (8.10)
follow.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 359
Then let
𝑣 𝑟 (𝑀, 𝑡) = 𝑀 −𝑟 −1 (𝑅𝑟 (𝑀, 𝑡) − 𝑅𝑟 (𝑀, 1) 𝑣ˆ (𝑀, 𝑡)). (8.11)
Using this notation, we can express the previously given optimal test functions as
𝑣 𝑟 ,𝑖 = 𝑣 𝑟 (𝐾 ∗ , 𝑡)𝑒 𝑖 for all 𝑟 = 0, 1, . . . , 𝑝.
When these optimal test function expressions are substituted into (8.9), we obtain
a system for the discrete solution
𝑝
∑︁
𝑢ℎ = 𝑢 ℎ, 𝑗 𝑡 𝑗 , 𝑢 ℎ, 𝑗 ∈ C𝑚 , (8.12a)
𝑗=0
where
∫ 1
𝑎𝑟 𝑗 ≔ 𝑡 𝑗+𝑟 𝑑𝑡.
0
This is a system of 𝑝 + 1 equations for the (vector-valued) unknowns 𝑢 ℎ, 𝑗 , 𝑗 =
0, . . . , 𝑝. The endpoint trace, which equals the exact solution by Proposition 8.1,
is given by
∫ 1
𝑢ˆ ℎ = 𝑣ˆ (𝐾, 0) 𝑢 0 + 𝑣ˆ (𝐾, 𝑡) 𝑓 (𝜏) 𝑑𝜏. (8.12c)
0
Thus, to compute 𝑢ˆ ℎ and 𝑢 ℎ, 𝑗 , we need techniques to compute the integrals in-
volving matrix exponentials.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
360 L. Demkowicz and J. Gopalakrishnan
and 𝑙˜𝑖 (𝜏) are the corresponding mapped Lagrange polynomials on interval [𝑡 𝑘−1 , 𝑡 𝑘 ].
Substituting the approximation for 𝑓 (𝜏) into the formula (8.13) from variation of
constants, we obtain a time-marching scheme,
𝑠
∑︁
𝑢 𝑘 = 𝑒 −ℎ𝑘 𝐾 𝑢 𝑘−1 + ℎ 𝑘 𝑏 𝑖 (−ℎ 𝑘 𝐾) 𝑓𝑖 ,
𝑖=1
It is standard to compute the weights using the so-called ‘𝜙-functions’ (see e.g.
Al-Mohy and Higham 2011 or Niesen and Wright 2012), defined as follows:
𝜙0 (𝑧) ≔ 𝑒 𝑧 ,
∫ 1
𝜃 𝑝−1
𝜙 𝑝 (𝑧) ≔ 𝑒 (1− 𝜃)𝑧 𝑑𝜃
0 (𝑝 − 1)!
1 1
= 𝜙 𝑝−1 (𝑧) − , 𝑝 = 1, 2, . . . .
𝑧 (𝑝 − 1)!
The two simple examples below show how they are used.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 361
Similarly,
1 𝑐1
𝑏 2 (𝑧) = 𝜙2 (𝑧) − 𝜙1 (𝑧).
𝑐2 − 𝑐1 𝑐2 − 𝑐1
Thus we obtain
To connect these existing results to the IPG scheme, first note that (8.12c) is
exactly the variation of constants formula (8.13) (which is also as expected from
the endpoint exactness result of Proposition 8.1). Hence the above-described
standard exponential integrator formulas can be used to compute the IPG fluxes
𝑢ˆ ℎ at the time steps 𝑡 𝑘 . It remains to discuss how to compute the solution 𝑢 ℎ in
between.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
362 L. Demkowicz and J. Gopalakrishnan
Proposition 8.4. The following relations between optimal test functions (8.11)
and the 𝜙 functions hold:
𝑟
∑︁ 𝑟!
𝑣 𝑟 (𝑀, 0) = (−1)𝑟 − 𝑗 𝜙𝑟 − 𝑗+1 (−𝑀),
𝑗=0
𝑗!
∫ 1 𝑟 (8.14)
∑︁ 𝑟!
𝑣 𝑟 (𝑀, 𝑡) 𝑡 𝑞 𝑑𝑡 = 𝑞! (−1)𝑟 − 𝑗 𝜙𝑟 − 𝑗+𝑞+2 (−𝑀)
0 𝑗=0
𝑗!
Example 8.5 (A one-point integrator for the interior IPG solution). Utilizing
Proposition 8.4, the system (8.12b) for 𝑝 = 0 reduces to the following scheme:
𝑘
𝑢 ℎ,0 = 𝜙1 (−ℎ 𝑘 𝐾)𝑢ˆ ℎ𝑘−1 + ℎ 𝑘 𝜙2 (−ℎ 𝑘 𝐾) 𝑓1 .
It computes a constant interior solution given from 𝑢ˆ ℎ𝑘−1 and 𝑓1 that is guaranteed
to equal the mean of the exact solution.
Example 8.6 (A two-point integrator for the interior IPG solution). For 𝑝 =
2, the system (8.12b) for the two coefficients of the interior solution reduces to the
following system of two equations after applying Proposition 8.4:
𝑘 1 𝑘
𝑢 ℎ,0 + 𝑢 ℎ,1 = 𝑔1 𝑢ˆ ℎ𝑘−1 , 𝑓1 , 𝑓2 , (8.15a)
2
1 𝑘 1 𝑘
𝑢 ℎ,0 + 𝑢 ℎ,1 = 𝑔2 𝑢ˆ ℎ𝑘−1 , 𝑓1 , 𝑓2 , (8.15b)
2 3
where
𝑔1 𝑢ˆ ℎ𝑘−1 , 𝑓1 , 𝑓2 = 𝜙1 (−ℎ 𝑘 𝐾)𝑢ˆ ℎ𝑘−1
1 𝑐2
+ ℎ𝑘 𝜙3 (−ℎ 𝑘 𝐾) − 𝜙2 (−ℎ 𝑘 𝐾) 𝑓1
𝑐1 − 𝑐2 𝑐1 − 𝑐2
1 𝑐1
+ ℎ𝑘 𝜙3 (−ℎ 𝑘 𝐾) − 𝜙2 (−ℎ 𝑘 𝐾) 𝑓2 ,
𝑐2 − 𝑐1 𝑐2 − 𝑐1
and
𝑔2 𝑢ˆ ℎ𝑘−1 , 𝑓1 , 𝑓2 = 𝜙1 (−ℎ 𝑘 𝐾)𝑢ˆ ℎ𝑘−1 − 𝜙2 (−ℎ 𝑘 𝐾)𝑢ˆ ℎ𝑘−1
1 𝑐2
+ ℎ𝑘 (𝜙3 (−ℎ 𝑘 𝐾) − 𝜙4 (−ℎ 𝑘 𝐾)) − (𝜙2 (−ℎ 𝑘 𝐾) − 𝜙3 (−ℎ 𝑘 𝐾)) 𝑓1
𝑐1 − 𝑐2 𝑐1 − 𝑐2
1 𝑐1
+ ℎ𝑘 (𝜙3 (−ℎ 𝑘 𝐾) − 𝜙4 (−ℎ 𝑘 𝐾)) − (𝜙2 (−ℎ 𝑘 𝐾) − 𝜙3 (−ℎ 𝑘 𝐾)) 𝑓2 .
𝑐2 − 𝑐1 𝑐2 − 𝑐1
The equations of (8.15) give the best 𝐿 2 -approximation of the interior solution in
the space of linear functions.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 363
for all 𝑣 ∈ 𝑌 . For this example, it is possible to derive an explicit formula for
function 𝜀, as shown in Muñoz-Matute, Demkowicz and Pardo (2022). However,
applying the formula requires coming up with special quadrature rules for matrix-
valued functions and it is cumbersome to use. Instead, it is recommended to
compute an inexact error representation 𝜀𝑟 using (6.7a), namely
(𝜀𝑟 , 𝑣)𝑌 = ℓ(𝑣) − 𝑏((𝑢 ℎ , 𝑢ˆ ℎ ), 𝑣) for all 𝑣 ∈ 𝑌 𝑟 , (8.16)
opt
with a 𝑌𝑟obtained by enlarging 𝑌ℎ by at least one linearly independent function.
opt
(One may, for example, set 𝑌 𝑟 = 𝑇(𝑃 𝑝+1 (𝛺) × C𝑚 ) ⊃ 𝑌ℎ .) Solving for 𝜀𝑟 from
(8.16) then only involves solving a small linear system after the computation of
𝑢 ℎ and 𝑢ˆ ℎ . Moreover, 𝜀𝑟 is almost as good an error estimator as 𝜀 because of the
following result.
Proposition 8.7 (Error estimator for time integrator). Set 𝑏 and ℓ as in (8.3),
opt
𝑋ℎ = 𝑃 𝑝 (𝛺) × C𝑚 , 𝛺 = (0, 1), and using any 𝑌 𝑟 ⊃ 𝑌ℎ , solve the practical DPG
method (6.7) for 𝑥 ℎ ∈ 𝑋ℎ and 𝜀 ∈ 𝑌 . Then 𝑥 ℎ coincides with the solution 𝑢 ℎ , 𝑢ˆ ℎ
𝑟 𝑟
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
364 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 365
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
366 L. Demkowicz and J. Gopalakrishnan
This has implications for both the DPG method and a dual DPG* method defined
shortly. First, the DPG method, in the form of the mixed system in Theorem 6.3(b),
is a discretization of (9.2), or its equivalent form (9.8), with
𝐹(𝑧, 𝑦) = ℓ(𝑦). (9.12)
Hence, once the inf-sup conditions (1.2a) and (9.9) are verified, the DPG method in
the mixed form (6.7) can be used to regularize and solve overdetermined systems,
even when it is not possible to verify the uniqueness assumption (1.2b) or the
compatibility condition (9.5). Moreover, if 𝐵 is a continuous bijection (so that the
system is no longer overdetermined) and 𝐹 is as in (9.12), then it is easy to see that
𝜁 = 0, and that 𝜁 𝑟 = 𝜀𝑟 together with 𝑥 ℎ solves the DPG method (6.7). Then (9.11)
reduces to
∥𝑥 − 𝑥 ℎ ∥ 𝑋 + ∥𝜀𝑟 ∥𝑌 ≲ inf 𝑟 𝑟 ∥𝑥 − 𝑧 ℎ ∥ 𝑋 , (9.13)
𝑧ℎ ∈𝑋ℎ ,𝑦 ∈𝑌
an error estimate we can also conclude from the theory in prior sections.
Next, consider dual formulations of (9.8). Since the operator generated by the
form 𝑎(·, ·) is self-adjoint, ‘dual problems’ of (9.8) take the same form as (9.8). By
a DPG* method we mean the method (9.10) for the case
𝐹(𝑧, 𝑦) = 𝑔(𝑧),
where 𝑔 ∈ 𝑋 ∗ . To distinguish from the previous case, let us now rename 𝜁 𝑟 as
𝜉 𝑟 and 𝑥 ℎ as 𝜆 ℎ . We can then rewrite (9.10) to express the DPG* method as the
method that finds 𝜉 𝑟 ∈ 𝑌 𝑟 and 𝜆 ℎ ∈ 𝑋ℎ satisfying
(𝜉 𝑟 , 𝑦)𝑌 + 𝑏(𝜆 ℎ , 𝑦) = 0 for all 𝑦 ∈ 𝑌 𝑟 , (9.14a)
𝑏(𝑧, 𝜉 𝑟 ) = 𝑔(𝑧) for all 𝑧 ∈ 𝑋ℎ . (9.14b)
Now it is evident that this is a discretization of (9.3) with the roles of 𝑋 and 𝑌
reversed, 𝐵∗ in place of 𝐵, 𝜉 in place of 𝑥, and 𝑔 in place of ℓ, that is,
(𝜉, 𝑦)𝑌 + 𝑏(𝜆, 𝑦) = 0 for all 𝑦 ∈ 𝑌 , (9.15a)
𝑏(𝑧, 𝜉) = 𝑔(𝑧) for all 𝑧 ∈ 𝑋, (9.15b)
thus revealing the connection with underdetermined systems. By verifying the
exact same inf-sup conditions as for the DPG method, namely (1.2a) and (9.9),
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 367
the estimate (9.11) then gives that the DPG* method is uniquely solvable and the
solution satisfies
∥𝜆 − 𝜆 ℎ ∥ 𝑋 + ∥𝜉 − 𝜉 𝑟 ∥𝑌 ≲ inf 𝑟 𝑟 ∥𝜆 − 𝑧 ℎ ∥ 𝑋 + ∥𝜉 − 𝑦 𝑟 ∥𝑌 . (9.16)
𝑧ℎ ∈𝑋ℎ ,𝑦 ∈𝑌
An important difference between the DPG* estimate (9.16) and the DPG estimate
(9.13) is that convergence in (9.16) depends on the regularity of an extraneous
Lagrange multiplier 𝜆.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
368 L. Demkowicz and J. Gopalakrishnan
In examples, one would want to leverage regularity of the dual solution, if available,
to verify (9.23c) and obtain some 𝑐(ℎ) that goes to zero as ℎ decreases.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 369
Theorem 9.2 (Duality argument for DPG formulations). Assume the setting
of (9.23) and (4.12). Then
ˆ 0) − (𝑥 ℎ , 𝑥ˆ ℎ , 𝜀𝑟 )∥ 𝑋0 × 𝑋×𝑌
∥𝑥 − 𝑥 ℎ ∥ 𝐿 ≤ 𝑐(ℎ) ∥𝑎∥ ∥(𝑥, 𝑥, ˆ . (9.24)
Then (9.23a) is obvious. The dual problem (9.23b) for 𝜉𝑔 = (𝑥 𝑔 , 𝑥ˆ𝑔 , 𝜀 𝑔 ) ∈ 𝐻˚ 1 (𝛺) ×
𝐻 −1/2 (𝜕𝛺 ℎ ) × 𝐻 1 (𝛺 ℎ ), after complex conjugations as needed, reads as follows:
(𝜀 𝑔 , 𝑦)𝑌 + (grad 𝑥 𝑔 , grad 𝑦)ℎ − ⟨𝑥ˆ𝑔 , 𝑦⟩ℎ = 0, (9.26a)
(grad 𝜀 𝑔 , grad 𝑧)ℎ = (𝑔, 𝑧)𝛺 , (9.26b)
⟨𝑧ˆ, 𝜀 𝑔 ⟩ℎ = 0 (9.26c)
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
370 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 371
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
372 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 373
namely
(𝑀 ∗ 𝑦)(𝑧) = (𝑀 𝑧)(𝑦) (10.6)
for any 𝑦 ∈ 𝑌 and 𝑥 ∈ 𝑋 (with no conjugation since the spaces are now over R).
Proposition 10.1. In the above setting, for any 𝑥 ∈ 𝑋,
(∇𝐽)(𝑥) = −𝑅 𝑋−1 (𝑑𝐵 𝑥 )∗ 𝑅𝑌−1 (ℓ − 𝐵(𝑥)).
Proof. For any 𝑥, 𝑧 ∈ 𝑋, since 𝑑𝐶 𝑥 𝑧 = (𝑑𝑅𝑌−1 (ℓ − 𝐵(𝑥)))(𝑧) = −𝑅𝑌−1 𝑑𝐵 𝑥 𝑧 and
𝐽(𝑥) = 12 (𝐶(𝑥), 𝐶(𝑥))𝑌 ,
𝑑𝐽 𝑥 𝑧 = (𝑑𝐶 𝑥 𝑧, 𝐶(𝑥))𝑌 = (−𝑅𝑌−1 𝑑𝐵 𝑥 𝑧, 𝐶(𝑥))𝑌
= −(𝑑𝐵 𝑥 𝑧)(𝐶(𝑥)) = −((𝑑𝐵 𝑥 )∗𝐶(𝑥))(𝑧)
by (10.6). Now the result follows from (10.4).
By Proposition 10.1, the steepest descent iteration becomes
𝑥 𝑛+1 = 𝑥 𝑛 + 𝛼𝑅 𝑋−1 (𝑑𝐵 𝑥𝑛 )∗ 𝑅𝑌−1 (ℓ − 𝐵(𝑥 𝑛 )), 𝑛 = 0, 1, . . . . (10.7)
It is applicable for DPG formulations when the inverse of the Gram matrices of
both the 𝑋 and the 𝑌 inner products can be efficiently applied; see the discussion
in Section 10.4.
Example 10.2 (Specialization to the linear case). Suppose the operator 𝐵 in
(10.1) is a linear continuous bijection. Then 𝑑𝐵 𝑥 = 𝐵 is independent of 𝑥. By
Proposition 10.1, the steepest descent iteration (10.5) then becomes
𝑥 𝑛+1 = 𝑥 𝑛 + 𝛼𝑅 𝑋−1 𝐵∗ 𝑅𝑌−1 (ℓ − 𝐵𝑥 𝑛 ).
If 𝑥 is the exact solution of (1.1), then this can be rewritten as
𝑥 − 𝑥 𝑛+1 = 𝐼 − 𝛼𝑅 𝑋−1 𝐵∗ 𝑅𝑌−1 𝐵 (𝑥 − 𝑥 𝑛 ).
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
374 L. Demkowicz and J. Gopalakrishnan
Using ∥𝑏∥ and the inf-sup constant 𝛾 introduced in Section 1, we know that for any
𝑤 ∈ 𝑋,
𝛾∥𝑤∥ 𝑋 ≤ ∥𝐵𝑤∥𝑌 ∗ ≤ ∥𝑏∥ ∥𝑤∥ 𝑋 ,
which implies
(𝛼𝛾 2 − 1)∥𝑤∥ 2 ≤ 𝛼∥𝐵𝑤∥𝑌2 ∗ − ∥𝑤∥ 2𝑋 ≤ (𝛼∥𝑏∥ 2 − 1)∥𝑤∥ 2𝑋 .
This shows that the sufficient condition (10.8) for convergence can be met if
−𝑞 ≤ 𝛼𝛾 2 − 1 and 𝛼∥𝑏∥ 2 − 1 ≤ 𝑞,
or equivalently
1−𝑞 1+𝑞
≤𝛼≤ .
𝛾 2 ∥𝑏∥ 2
The smallest possible contraction constant
∥𝑏∥ 2 − 𝛾 2 𝐶 2 − 1 ∥𝑏∥
𝑞= = , where 𝐶 = ,
∥𝑏∥ 2 + 𝛾 2 𝐶 2 + 1 𝛾
is achieved when the lower and upper bounds for 𝛼 coincide. To summarize,
selecting any 𝑞 such that (𝐶 2 − 1)/(𝐶 2 + 1) ≤ 𝑞 < 1 and setting any 𝛼 satisfying
(1 − 𝑞)/𝛾 2 ≤ 𝛼 ≤ (1 + 𝑞)/∥𝑏∥ 2 , the steepest descent iterations with such an 𝛼
converge, and the error-reducing operator is a contractive map with a contraction
constant 𝑞 or higher.
If 𝑋 and 𝑌 are endowed with optimal norms that make 𝑏(·, ·) into a generalized
duality pairing as in Definition 3.5, then 𝛾 = ∥𝑏∥ = 𝐶 = 1. Hence 𝑞 = 0 and the
steepest descent method with 𝛼 = 1 delivers the DPG solution in just one step,
independently of the initial iterate.
As we have shown previously (see e.g. (9.6)), an equivalent way to compute this
minimizer is by solving
(𝑑𝐵 𝑥𝑛 )∗ 𝑅𝑌−1 𝑑𝐵 𝑥𝑛 Δ𝑥 = (𝑑𝐵 𝑥𝑛 )∗ 𝑅𝑌−1 (ℓ − 𝐵(𝑥 𝑛 )). (10.10)
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 375
This is solvable when 𝑑𝐵 𝑥𝑛 satisfies the inf-sup condition. Comparing (10.10) with
the increment Δ𝑥f of the steepest descent step iteration (10.7) with unit step size
𝛼 = 1, which solves
f = (𝑑𝐵 𝑥𝑛 )∗ 𝑅 −1 (ℓ − 𝐵(𝑥 𝑛 )),
𝑅 𝑋 Δ𝑥 𝑌
we see that the two iterations coincide with each other provided
𝑅 𝑋 = (𝑑𝐵 𝑥𝑛 )∗ 𝑅𝑌−1 𝑑𝐵 𝑥𝑛 ,
that is, provided we use the step-dependent energy norm
|||Δ𝑢||| 𝑋 = ∥𝑑𝐵 𝑥𝑛 Δ𝑢∥𝑌 ∗
for the space 𝑋 (which, of course, is a norm when the linearized operator 𝑑𝐵 𝑥𝑛
satisfies the inf-sup condition).
10.4. Trade-offs
The steepest descent iteration (10.7) requires the application of the inverse of the
Gram matrices of both the 𝑋 and 𝑌 inner products due to the presence of 𝑅 𝑋−1 and
𝑅𝑌−1 there. In Section 4 we discussed at length the DPG localization techniques to
make 𝑅𝑌−1 easy. However, we also need 𝑅 𝑋−1 to implement (10.5). This is a drawback
of the descent approach. Nonetheless, 𝑅 𝑋 is a linear Hermitian positive definite
operator. Moreover, the component spaces of 𝑋 have norms that are either standard
norms or quotient trace norms, which may be treated as norms arising from Schur
complements of Gram matrices of standard norms. As shown in Barker, Dobrev,
Gopalakrishnan and Kolev (2018), such Schur complements and the corresponding
𝑅 𝑋 can be efficiently preconditioned using off-the-shelf preconditioners.
Consider the alternative of the Newton–Raphson iterations which, in the context
of the minimum residual methodology, translates into the Gauss–Newton method.
Here we require the linearization 𝑑𝐵 𝑥 to satisfy the inf-sup condition. This require-
ment is absent for the steepest descent methodology and is another reason to opt
for it. For example, in nonlinear elasticity, the linearized problem may be singular
(such as in buckling) which, in numerics, manifests as bad conditioning. In the
steepest descent approach, we need only invert well-conditioned Riesz operators.
In both methodologies we need 𝑑𝐵 𝑥 , but in steepest descent we use it only to com-
pute the load, whereas in the Gauss–Newton approach we also use it to compute
and invert the stiffness matrix. Steepest descent naturally provides the possibility
of incorporating additional constraints by solving the minimum residual problem
with additional constraints. Incorporating the constraints through a penalty method
results in a minimal modification of the algorithm and, in particular, allows the
use of a penalty term that is only once differentiable; see Bristeau et al. (1985).
This is a common situation for inequality constraints. Thus the steepest descent
methodology appears to be much more robust than Gauss–Newton.
On the negative side, the steepest descent iterations deliver only linear conver-
gence compared with the quadratic convergence of Newton methods. It may be
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
376 L. Demkowicz and J. Gopalakrishnan
advantageous to combine the two methodologies into a single algorithm. For ex-
ample, one may start with the more robust steepest descent iterations and finish
with Gauss–Newton iterations once the increments become small enough.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 377
and shells include those of Calo, Collier and Niemi (2014), Führer, Heuer and
Niemi (2022, 2023) and Führer, Heuer and Sayas (2020).
DPG ideas are playing an important role in development of parameter-robust
methods. The work of Broersen, Dahmen and Stevenson (2018) gave stability
estimates for the DPG method applied to the linear transport equation that are
uniform in the relative orientation of the local mesh and the advection direction.
For singularly perturbed problems, specifically for advection-dominated diffusion,
parameter-robust stability was confirmed in Demkowicz and Heuer (2013) and
Chan, Heuer, Bui-Thanh and Demkowicz (2014b). The case of reaction-dominated
diffusion was studied in Heuer and Karkulik (2017b).
One of the attractive features of the DPG method is that it only requires the solu-
tion of a symmetric positive definite system, even when the original boundary value
problem is non-self-adjoint. This can be leveraged in the design of iterative solvers
and high performance computing. In Barker et al. (2018), one can find specifics
on how to combine off-the-shelf algebraic preconditioners effectively to develop
highly scalable DPG solvers. A DPG solver for harmonic wave propagation, in-
tegrated within an adaptive procedure, through a two-grid-like preconditioner for
the conjugate gradient method, was developed in Petrides and Demkowicz (2017)
as well as in Badger, Henneking, Petrides and Demkowicz (2023); it exhibits
excellent practical efficiency. Connecting DPG with other similar saddle-point
least-squares systems, certain solvers are suggested in Bacuta, Hayes and Jacavage
(2021b). Parameter robustness in DPG solvers is still highly sought after in specific
applications and remains an active area of research.
The design of stable spacetime formulations by DPG techniques is another topic
we have not detailed in this review, except for the early work in Demkowicz et al.
(2017), which is applicable beyond spacetime problems. Since then, spacetime
DPG formulations for transient waves have been studied in Gopalakrishnan and
Sepúlveda (2019), Sepúlveda (2018) and Ernesti and Wieners (2019), and a space-
time DPG method for the heat equation was developed in Diening and Storn (2022).
The approach of Demkowicz et al. (2017) to prove wellposedness in graph spaces
(along the lines of the theory of Friedrichs systems) was found to be difficult for
various spacetime problems. A new approach has been proposed in the recent
work of Führer, González and Karkulik (2024) using Bochner spaces. This shows
promise for reducing the technicalities in proving convergence of DPG and residual
minimization methods for spacetime methods.
An exciting new frontier is the use of DPG ideas for variationally correct machine
learning approaches. The very recent work of Rojas et al. (2024) defines a quadratic
loss functional, motivated by DPG-type formulations, within a physics-informed
neural network to solve a boundary value problem (and earlier developments in
physics-informed neural networks can be found in Kharazmi, Zhang and Karniada-
kis 2021). The recent work of Bachmayr, Dahmen and Oster (2024) centres around
learning the parameter-to-solution map for systems of partial differential equa-
tions that depend on a potentially large number of parameters. These works show
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
378 L. Demkowicz and J. Gopalakrishnan
Acknowledgements
The ideas behind DPG methods were developed over the years, utilizing support
from the Air Force Office of Scientific Research (AFOSR) and the National Science
Foundation (NSF). Both authors are immensely grateful for this support. The
preparation of this review was supported in part by AFOSR grant FA9550-23-1-
0103 and NSF grant 2245077. This work also benefited from activities organized
under the auspices of NSF RTG grant 2136228 as well as a decadal series of
workshops on DPG and residual minimization methods held in Austin (USA,
2013), Delft (The Netherlands, 2015), Portland (USA, 2017), Berlin (Germany,
2019), Santiago (Chile, 2022) and Bilbao (Spain, 2024). The authors gratefully
acknowledge feedback from several participants of these workshops which shaped
this review.
References
A. H. Al-Mohy and N. J. Higham (2011), Computing the action of the matrix exponential,
with an application to exponential integrators, SIAM J. Sci. Comput. 33, 488–511.
D. N. Arnold, R. S. Falk and R. Winther (2006), Finite element exterior calculus, homolo-
gical techniques, and applications, Acta Numer. 15, 1–155.
I. Babuška (1971), Error-bounds for finite element method, Numer. Math. 16, 322–333.
I. Babuška, A. K. Aziz, G. Fix and R. B. Kellogg (1972), Survey lectures on the mathe-
matical foundations of the finite element method, in The Mathematical Foundations of
the Finite Element Method with Applications to Partial Differential Equations (A. K.
Aziz, ed.), Academic Press, pp. 1–359.
M. Bachmayr, W. Dahmen and M. Oster (2024), Variationally correct neural residual
regression for parametric PDEs: On the viability of controlled accuracy. Available at
arXiv:2405.20065.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 379
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
380 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 381
L. Demkowicz and J. Gopalakrishnan (2011a), Analysis of the DPG method for the Poisson
equation, SIAM J. Numer. Anal. 49, 1788–1809.
L. Demkowicz and J. Gopalakrishnan (2011b), A class of discontinuous Petrov–Galerkin
methods, Part II: Optimal test functions, Numer. Methods Partial Differential Equations
27, 70–105.
L. Demkowicz and J. Gopalakrishnan (2013), A primal DPG method without a first-order
reformulation, Comput. Math. Appl. 66, 1058–1064.
L. Demkowicz and J. Gopalakrishnan (2017), Discontinuous Petrov Galerkin (DPG)
method, in Encyclopedia of Computational Mechanics, second edition, Wiley Com-
putational Mechanics Online.
L. Demkowicz and N. Heuer (2013), Robust DPG method for convection-dominated dif-
fusion problems, SIAM J. Numer. Anal. 51, 2514–2537.
L. Demkowicz and J. T. Oden (1986a), An adaptive characteristic Petrov–Galerkin finite
element method for convection-dominated linear and nonlinear parabolic problems in
one space variable, J. Comput. Phys. 67, 188–213.
L. Demkowicz and J. T. Oden (1986b), An adaptive characteristic Petrov–Galerkin finite
element method for convection-dominated linear and nonlinear parabolic problems in
two space variables, Comput. Methods Appl. Mech. Engrg 55, 63–87.
L. Demkowicz and P. Zanotti (2020), Construction of DPG Fortin operators revisited,
Comput. Math. Appl. 80, 2261–2271.
L. Demkowicz, J. Gopalakrishnan and B. Keith (2020), The DPG-star method, Comput.
Math. Appl. 79, 3092–3116.
L. Demkowicz, J. Gopalakrishnan and A. Niemi (2012a), A class of discontinuous Petrov–
Galerkin methods, Part III: Adaptivity, Appl. Numer. Math. 62, 396–427.
L. Demkowicz, J. Gopalakrishnan, I. Muga and J. Zitelli (2012b), Wavenumber explicit
analysis for a DPG method for the multidimensional Helmholtz equation, Comput.
Methods Appl. Mech. Engrg 213/216, 126–138.
L. Demkowicz, J. Gopalakrishnan, S. Nagaraj and P. Sepúlveda (2017), A spacetime DPG
method for the Schrödinger equation, SIAM J. Numer. Anal. 55, 1740–1759.
L. Diening and J. Storn (2022), A space-time DPG method for the heat equation, Comput.
Math. Appl. 105, 41–53.
I. Ekeland and R. Témam (1999), Convex Analysis and Variational Problems, Vol. 28 of
Classics in Applied Mathematics, SIAM.
T. Ellis, J. Chan and L. Demkowicz (2016), Robust DPG methods for transient convection–
diffusion, in Building Bridges: Connections and Challenges in Modern Approaches to
Numerical Partial Differential Equations (G. R. Barrenechea et al., eds), Vol. 114 of
Lecture Notes in Computational Science and Engineering, Springer.
T. Ellis, L. Demkowicz and J. Chan (2014), Locally conservative discontinuous Petrov–
Galerkin finite elements for fluid problems, Comput. Math. Appl. 68, 1530–1549.
A. Ern and J.-L. Guermond (2021), Finite Elements II, Springer.
A. Ern, J.-L. Guermond and G. Caplain (2007), An intrinsic criterion for the bijectivity of
Hilbert operators related to Friedrichs’ systems, Commun. Partial Differential Equations
32, 317–341.
J. Ernesti and C. Wieners (2019), A space-time discontinuous Petrov–Galerkin method for
acoustic waves, in Space-Time Methods: Applications to Partial Differential Equations
(U. Langer and O. Steinbach, eds), De Gruyter, pp. 89–116.
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
382 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
The discontinuous Petrov–Galerkin method 383
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102
384 L. Demkowicz and J. Gopalakrishnan
Downloaded from https://www.cambridge.org/core. IP address: 84.33.184.126, on 22 Jul 2025 at 09:22:39, subject to the Cambridge Core terms of use, available at
https://www.cambridge.org/core/terms. https://doi.org/10.1017/S0962492924000102