Metodi Numerici Per Le Equazioni Alle Derivate Parziali
Metodi Numerici Per Le Equazioni Alle Derivate Parziali
February 7, 2021
2
Introduction
3
4
Chapter 1
A Few Concepts
on Partial Differential Equations
(
F ind u ∈ V such that
[VP ] (1.1)
a(u, v) = F (v) ∀v ∈ V .
Lu = −∇ · (µ∇u) + β · ∇u + σu (1.4)
This is the norm of v in the Sobolev space H m (A). Note that H 0 (A) = L2 (A).
5
6 CHAPTER 1. PARTIAL DIFFERENTIAL EQUATIONS
Given two functions v1 ∈ H 1 (A1 ) and v2 ∈ H 1 (A2 ), denote by v the function defined in A by
(
v1 in A1 ,
v=
v2 in A2 .
Proof. Obviously, v ∈ L2 (A) with kvk2L2 (A) = kv1 k2L2 (A1 ) + kv2 k2L2 (A2 ) . Then, v ∈ H 1 (A) if
∂v
and only if, for 1 ≤ i ≤ d, each partial derivative ∂x i
in the sense of distributions coincides
with a function gi ∈ L (A). Since by assumption ∂xi ∈ L2 (A1 ) and ∂v
2 ∂v1 2
∂xi ∈ L (A2 ), the natural
2
Let us compute the i-th distributional derivative of v: for any test function φ ∈ D(A), one has
Z
∂v ∂φ
h , φi = − v (definition of distributional derivative)
∂xi A ∂xi
Z Z
∂φ ∂φ
= − v − v (additivity of integrals)
A1 ∂xi A2 ∂xi
Z Z
∂φ ∂φ
= − v1 − v2 (definition of v)
A1 ∂xi A2 ∂xi
Z Z Z Z
∂v1 (1) ∂v1 (2)
= φ − v1 φ ni + φ − v2 φ ni (integration by parts)
A1 ∂x i Γ A1 ∂x i Γ
Z Z
(1) (2)
= gi φ − (v1 ni + v2 ni ) φ (definition of gi ) ,
A Γ
(1)
where ni denotes the i-th component of the unit normal vector to Γ, oriented outward with
(2)
respect to A1 and ni has the similar meaning with respect to A2 . Then,
Z
∂v
h , φi = gi φ ∀φ ∈ D(A)
∂xi A
if and only if Z
(1) (2)
(v1 ni + v2 ni ) φ = 0 ∀φ ∈ D(A) .
Γ
(1) (2)
Since φ is arbitrary, this is equivalent to v1 ni + v2 ni = 0 in Γ, which in turn is equivalent to
(2) (1)
v1 − v2 = 0 on Γ since ni = −ni .
Chapter 2
Galerkin Methods
The Galerkin approach is a conceptually simple, yet powerful strategy to discretize the varia-
tional problem [VP ] introduced in (1.1), i.e., to generate a similar problem in finite dimension
(hence, computationally tractable).
The well-posedness of this problem directly stems from the assumptions made on the exact
(infinite dimensional) variational problem [VP ]. Indeed, Vδ is a closed subspace of V , being
finite dimensional; consequently, it inherits the Hilbert-space structure from V . Furthermore,
the forms a and F are continuous on Vδ since they are on the larger space V , and similarly a is
coercive on Vδ since it is on V . Thus, we can apply the Lax-Milgram theorem to [VP ]δ , getting
the following fundamental result.
Theorem 2.1.1 The discrete variational problem [VP ]δ stated in (2.1) admits one and only one
solution uδ , which satisfies
1
kuδ kV ≤ kF kV 0 . (2.2)
α
The bound (2.2) has an important interpretation: it is an instance of the property called
numerical stability. It means that a suitable norm of the discrete solution (here, the norm
of V ) can be bounded by a quantity independent of the discretization parameter δ. This is a
necessary condition to have the convergence of uδ to the exact solution u of [VP ], if we let δ
tend to 0; indeed, remember that a convergent sequence is necessarily bounded.
7
8 CHAPTER 2. GALERKIN METHODS
for suitable coefficients vj ∈ R; consequently, we can associate to vδ the vector v = (vj )1≤j≤Nδ ∈
RNδ of its coefficients with respect to the chosen basis. By definition of basis, the mapping
vδ 7→ v is an isomorphism; hence, we can identify a function in Vδ with a vector in RNδ . This
is important, because computers and programming languages are built to handle vectors and
matrices in the most efficient way. Note that to apply correctly the rules of linear algebra, v
and the subsequent vectors must be considered as column vectors.
Once a basis is selected, we obtain Nδ equations by picking each basis function as a test
function in the variational equations (2.1), i.e.,
Amazingly, this finite set of equations is equivalent to the infinitely-many equations in (2.1)
(one for each choice of vδ ∈ Vδ ); this is a consequence of the linear dependence of both sides
upon vδ . Indeed, if vδ has the representation (2.3) and the equations (2.4) hold true, then by
multiplication of both sides by vj and summation over j we get
Nδ
X Nδ
X
vj a(uδ , ϕj ) = vj F (ϕj ) ,
j=1 j=1
for some unknown coefficients uk , collected in the column vector u = (uk )1≤k≤Nδ ∈ RNδ . Sub-
stituting this expression into (2.4) and using the linearity of a with respect to the first variable,
we obtain
XNδ
uk a(ϕk , ϕj ) = F (ϕj ), 1 ≤ j ≤ Nδ . (2.6)
k=1
2.3. FEATURES OF THE STIFFNESS MATRIX 9
This suggests to introduce the square matrix A = (ajk )1≤j,k≤Nδ ∈ RNδ ×Nδ and the column
vector f = (fj )1≤j≤Nδ ∈ RNδ whose entries are
Proposition 2.2.1 Let Φδ = {ϕ1 , ϕ2 , . . . , ϕNδ } be any basis in Vδ . Then, the Galerkin problem
[VP ]δ is equivalent to the algebraic system
Au = f , (2.8)
where the entries of the matrix A and the vector f are given in (2.7).
Borrowing from a terminology used in structural mechanics, the matrix A will be called the
stiffness matrix associated with the bilinear form a and the basis Φδ .
since vδ = N
P δ
j=1 vj ϕj as well (j or k above are just dummy variables of summation). Thus,
invoking coercivity,
v T Av = a(vδ , vδ ) ≥ αkvδ k2V ≥ 0 . (2.9)
We can conclude once we have shown that v T Av = 0 implies v = 0. If v T Av = 0, then from
the previous inequality we deduce kvδ kV = 0, hence vδ = 0 ∈ Vδ . By definition of basis, this is
possible only if all its coefficients vk are zero, which is precisely equivalent to v = 0.
10 CHAPTER 2. GALERKIN METHODS
This important property has the following equivalent interpretation. If A is symmetric (which
happens if the bilinear form a(u, v) is symmetric), then positive definiteness is equivalent to the
fact that all the eigenvalues of A are strictly positive. If A is not symmetric, we can decompose
it in the sum of a symmetric matrix and an antisymmetric one,
A = As + Aas with As = 12 (A + AT ), Aas = 12 (A − AT ),
and observe that v T Aas v = 0 for any vector v, since
v T Aas v = v T (−ATas )v = −(Aas v)T v = −v T Aas v ,
which is possible only if this quantity is 0. We conclude that v T As v = v T Av > 0, i.e., As is
symmetric and positive-definite. Summarizing, we have proven the following result.
Corollary 2.3.2 All the eigenvalues of the symmetric part As of the stiffness matrix A are
strictly positive.
The previous property allows one to apply to the solution of the system (2.8) all the (direct
or iterative) techniques that exploit the positive definiteness of the matrix A in the symmetric
case (such as, e.g., the Choleski factorization or the Conjugate Gradient method), or of the
symmetric part of A in the general case (such as, e.g., the GMRES method).
Other features of the stiffness matrix are of interest. For instance, if the selected basis functions
have a localized support in the domain Ω of interest, i.e., they are different from 0 only in a
small portion of the domain, the stiffness matrix may be sparse, meaning that the percentage
of non-zero entries is “small”. Furthermore, if the basis functions are numbered convenably,
the stiffness matrix may be banded, meaning that the non-zero entries are located inside a
(often small) band around the main diagonal. All such features can be cleverly exploited to save
memory in the representation of the stiffness matrix in the computer, and to save computational
time in the solution of the linear system.
Finally, we should not forget that to every square non-singular matrix we can associate a
number ≥ 1, the condition number (with respect to a chosen norm), whose magnitude indi-
cates how “good” or “bad” is to operating with such matrix at the numerical level (in terms
of, e.g., propagation of the numerical errors in a direct method, or steps needed to match a
prescribed error tolerance in an iterative method). For a stiffness matrix, the condition number
depends on the choice of the basis used to represent functions in Vδ . There are indeed “good”
bases and “bad” bases. Many preconditioning techniques, aimed at transforming the linear
system into an equivalent one with a matrix having a better condition number, can be viewed
indeed as the application of a suitable change of basis in Vδ .
There exists two main strategies to estimate the discretization error, which are somehow
complementary to each other:
• A priori error estimation: one gives estimates of the error before even computing the
Galerkin solution uδ , by just knowing (or assuming) certain qualitative properties of the
exact solution u.
• A posteriori error estimation: one gives estimates of the error without any knowledge
of the exact solution, using only the information on the data of the problem and the
computed Galerkin solution uδ .
The former strategy is more theoretically oriented, aiming at predicting the qualitative behavior
of the discretization error, such as e.g. its asymptotic decay as the discretization parameter δ
tends to 0. The latter strategy may have a more direct and practical impact, for instance by
providing quantitative information that can be used to decide how to improve the quality of the
current approximation (adaptivity).
Proof. The exact solution of Problem [VP ] satisfies, in particular for any vδ ∈ Vδ ,
a(u, vδ ) = F (vδ ) .
a(uδ , vδ ) = F (vδ )
and using the linearity of a with respect to the first argument, we obtain the result.
This property is often known as the Galerkin orthogonality property. Indeed, when the form
a is symmetric, (2.12) can be rephrased as
(u − uδ , vδ )a = 0 , ∀vδ ∈ Vδ ,
where (·, ·)a is the inner product in V associated to a (see (1.2)). These relations show that
u − uδ is a-orthogonal to the subspace Vδ , or, equivalently, that uδ is the a-orthogonal projection
of u upon Vδ (see Fig. 2.4.1). This gives an immediate minimality property for the discretization
error: indeed, since the orthogonal projection is the closest element to u in Vδ , it holds
ku − uδ ka ≤ ku − vδ ka , ∀vδ ∈ Vδ ,
which shows that uδ is the best approximation of u in Vδ in the a-norm. In terms of the original
norm in V , using the norm equivalence (1.3), we deduce from the previous inequality
√ p
α ku − uδ kV ≤ kak ku − vδ kV , ∀vδ ∈ Vδ ,
12 CHAPTER 2. GALERKIN METHODS
whence r
kak
ku − uδ kV ≤ ku − vδ kV , ∀vδ ∈ Vδ . (2.13)
α
A similar estimate can be obtained when the form a is not symmetric, using again the Galerkin
orthogonality property. We proceed as follows:
Simplifying, we arrive at
kak
ku − uδ kV ≤ ku − vδ kV , ∀vδ ∈ Vδ . (2.14)
α
Together with (2.13), we have thereby established the following fundamental result, often referred
to as the Céa Lemma.
Theorem 2.4.2 The Galerkin discretization error satisfies the bound
ku − uδ kV ≤ Ca ku − vδ kV , ∀vδ ∈ Vδ , (2.15)
where q
kak if a is symmetric ,
α
Ca = (2.16)
kak
otherwise .
α
2.4. ERROR ANALYSIS 13
i.e., the exact solution u of our Problem [VP ] can be approximated arbitrarily well by functions
in the subspaces Vδ . Then, eq. (2.17) and the comparison theorem for limits immediately imply
that
ku − uδ kV → 0 as δ → 0 , (2.21)
which expresses the convergence of the family of Galerkin solutions to the exact solution, in
the norm of V . In short, we say that the Galerkin discretization method (2.1) is convergent (to
the solution u).
At this point, we have to check the consistency assumption for our specific choice of discretiza-
tion spaces, and also to get information about the speed of convergence of the discretization,
i.e., the way the distance distV (u, Vδ ) decays when δ tends to 0. The following result provides a
general framework, which can be applied in many different situations.
Theorem 2.4.3 Assume that Z ⊂ V is a normed space, for which the following consistency
condition holds: there exists a function ψZ : R+ → R+ satisfying ψZ (δ) → 0 for δ → 0, such
that
distV (w, Vδ ) ≤ ψZ (δ) kwkZ , ∀w ∈ Z . (2.22)
14 CHAPTER 2. GALERKIN METHODS
Remark 2.4.4 i) A typical example of consistency estimate (2.22) for finite element approxi-
mations, that will be proven in the forthcoming chapters, is
distH 1 (Ω) (w, Vh ) ≤ C h kwkH 2 (Ω) ∀w ∈ H 2 (Ω) ,
where (setting here δ = h) Vh is a space of continuous, piecewise polynomial functions on a
partition of the domain Ω and h is a characteristic size of the elements of the partition.
ii) Inequality (2.23) is an example of a priori error estimate for the Galerkin method.
iii) We have obtained the convergence results (2.21) or (2.23) by exploiting on the one side
the consistency condition (2.20) and on the other side the bound (2.17) in Céa’s theorem; this
bound relies on the property of coercivity of the bilinear form a, which – when applied on the
subspaces Vδ – yields the numerical stability property (2.1.1).
One can actually explicitly prove that convergence is implied by consistency and numerical sta-
bility, as an instance of a general theorem in Numerical Analysis, known as the Lax-Richmyer
Theorem.
1 1
krδ kV 0 ≤ ku − uδ kV ≤ krδ kV 0 . (2.26)
kak α
whence
rδ (v) kak ku − uδ kV kvkV
krδ kV 0 = sup ≤ sup = kak ku − uδ kV ,
v∈V \{0} kvkV v∈V \{0} kvkV
which gives the left-hand inequality in (2.26). On the other hand, from the coercivity inequality
αku − uδ k2V ≤ a(u − uδ , u − uδ ), we obtain
Remark 2.4.6 In the manipulation of the dual norm of the residual, with the aim of designing
a practical a posteriori error estimator, one can use the relation (2.25) to express the residual as
F (v − vδ ) − a(uδ , v − vδ )
krδ kV 0 = sup . (2.28)
v∈V \{0} kvkV
Choosing conveniently vδ , one can start from this expression to derive a computable error esti-
mator ηδ ' krδ kV 0 . We will give some examples in Sects. 3.4.3 and ...
16 CHAPTER 2. GALERKIN METHODS
2.5 Extensions
Hereafter, we present three variants of the classical Galerkin method discussed above. They
may lead to a better accuracy or a better computational efficiency. A combination of variants
is also possible.
The inf-sup condition is a generalization of the coercivity condition, in the sense that if Wδ = Vδ
the inf-sup condition follows from the coercivity condition: indeed, for any wδ ∈ Vδ , wδ 6= 0, one
has
a(wδ , wδ ) a(wδ , vδ )
αkwδ kV ≤ ≤ sup .
kwδ kV v ∈V kvδ kV
δ δ
vδ 6= 0
Using the inf-sup condition, we can prove the well-posedness of the variational problem (2.29).
Indeed, we have
a(uδ , vδ ) F (vδ ) F (v)
αδ kuδ kV ≤ sup = sup ≤ sup = kF kV 0 ,
vδ ∈ Vδ kvδ kV vδ ∈ Vδ kvδ kV v ∈ Vδ kvkV
vδ 6= 0 vδ 6= 0 v 6= 0
i.e.,
1
kuδ kV ≤
kF kV 0 . (2.32)
αδ
Two important consequences stem from this inequality:
2.5. EXTENSIONS 17
• if F = 0, then uδ = 0, which implies existence and uniqueness of the solution, since [VP ]δ
is a linear problem in finite dimension;
• we assess the numerical stability of the discretization, provided the inf-sup condition is
uniform in δ, i.e., there exists a constant ᾱ > 0 independent of δ such that αδ ≥ ᾱ. Indeed,
in this case we have
1
kuδ kV ≤ kF kV 0
ᾱ
and the right-hand side is independent of δ.
Finally, in order to estimate the discretization error u − uδ by exploiting the inf-sup condition,
we write u − uδ = (u − wδ ) + (wδ − uδ ) for an arbitrary wδ ∈ Wδ , and use the triangle inequality
ku − uδ kV ≤ ku − wδ kV + kwδ − uδ kV . (2.33)
The second addend on the right-hand side can be controlled by the inf-sup condition
a(wδ − uδ , vδ )
αδ kwδ − uδ kV ≤ sup ,
vδ ∈ Vδ kvδ kV
vδ 6= 0
Hence,
kak
kwδ − uδ kV ≤ ku − wδ kV .
αδ
Inserting this bound in (2.33), and taking the infimum over all wδ ∈ Wδ , we conclude that
kak
ku − uδ kV ≤ 1 + distV (u, Wδ ) . (2.34)
αδ
This result is the counterpart of the Céa Lemma for Petrov-Galerkin methods. If the inf-sup
condition is uniform, then the constant in parenthesis can be bounded independently of δ, which
means that the discretization error behaves as the best approximation error of u in Wδ . On the
contrary, if αδ becomes smaller and smaller as δ → 0, then in practice the discretization error
may be significantly larger than the best approximation error.
We see here a clear example of the Lax-Richtmyer theorem: if the discretization is numerically
stable (i.e., αδ ≥ ᾱ > 0) and consistent (i.e., distV (u, Wδ ) → 0 as δ → 0), then it is convergent
(ku − uδ kV → 0 as δ → 0).
Thus, we assume that we are given a bilinear form aδ : Vδ × Vδ → R and a linear form
Fδ : Vδ → R, and we consider the approximate discrete variational problem
(
F ind uδ ∈ Vδ such that
[VP ]δ (2.35)
aδ (uδ , vδ ) = Fδ (vδ ) ∀vδ ∈ Vδ .
This is the natural setting for a Galerkin method with numerical integration (see Sect. ... ).
To guarantee the well-posedness of this problem, we have to make a suitable assumption on
the bilinear form aδ . A sufficient assumption is the discrete coercivity condition, i.e., there exists
αδ > 0 such that
aδ (vδ , vδ ) ≥ αδ kvδ k2V ∀vδ ∈ Vδ .
In order to estimate the error u − uδ , we use again (2.33) with arbitrary wδ ∈ Vδ . By the
coercivity of aδ in Vδ , we have
aδ (wδ − uδ , vδ )
αδ kwδ − uδ kV ≤ sup ,
vδ ∈ V δ kv δ k V
vδ 6= 0
with aδ (wδ − uδ , vδ ) = aδ (wδ , vδ ) − aδ (uδ , vδ ) = aδ (wδ , vδ ) − Fδ (vδ ). Adding and subtracting the
quantity a(wδ , vδ ) − F (vδ ) on the right-hand side, we obtain
aδ (wδ − uδ , vδ ) = a(wδ , vδ ) − F (vδ ) + [aδ (wδ , vδ ) − a(wδ , vδ )] − [Fδ (vδ ) − F (vδ )]
= a(wδ − u, vδ ) + (aδ − a)(wδ , vδ ) − (Fδ − F )(vδ ) .
Thus,
a(wδ − u, vδ ) (aδ − a)(wδ , vδ ) (Fδ − F )(vδ )
αδ kwδ − uδ kV ≤ sup + sup + sup
vδ ∈ V δ kvδ kV vδ ∈ Vδ kvδ kV vδ ∈ V δ kvδ kV
vδ 6= 0 vδ 6= 0 vδ 6= 0
≤ kak ku − wδ kV + Eδ (a; wδ ) + Eδ (F ) ,
where
(a − aδ )(wδ , vδ ) (F − Fδ )(vδ )
Eδ (a; wδ ) := sup , Eδ (F ) := sup (2.36)
vδ ∈ Vδ kvδ kV vδ ∈ V δ kvδ kV
vδ 6= 0 vδ 6= 0
are discretization errors of the forms a and F , respectively. Inserting this bound in (2.33), we
get
kak
ku − uδ kV ≤ 1 + ku − wδ kV + Eδ (a; wδ ) + Eδ (F ) ;
αδ
since wδ ∈ Vδ is arbitrary, we conclude with the a-priori error bound
kak
ku − uδ kV ≤ inf 1+ ku − wδ kV + Eδ (a; wδ ) + Eδ (F ) . (2.37)
wδ ∈ Vδ αδ
wδ 6= 0
This result is known as first Strang Lemma. It implies that the convergence of the discretiza-
tion follows from the numerical stability (i.e., αδ ≥ ᾱ for some ᾱ > 0 independent of δ), and the
following consistency conditions:
- for all δ, there exists wδ ∈ Vδ such that ku − wδ kV → 0 and Eδ (a; wδ ) → 0 as δ → 0;
- Eδ (F ) → 0 as δ → 0.
2.5. EXTENSIONS 19
Throughout this chapter, I will denote a closed bounded interval of the real line; up to a
translation, we may assume that I = [0, L] for some real L > 0. By a small abuse of notation,
we will denote by L2 (I), H 1 (I), ... the corresponding spaces defined on the open interval int(I) =
(0, L).
dim Pk (E) = k + 1 .
21
22 CHAPTER 3. DISCRETIZATIONS IN ONE SPACE VARIABLE
• The basis of the monomials 1, x, x2 , . . . , xk . This is the classical basis, however for increas-
ing values of k it becomes numerically unstable, or ill-conditioned (the monomials tend to
become linearly dependent). Hence, it should never be used in computations, except for
very small values of k.
• The basis of Legendre orthogonal polynomials L0 (x), L1 (x), . . . , Lk (x), which satisfy
Z
deg Ln (x) = n and Ln (x) Lm (x) dx = δn,m , 0 ≤ n, m ≤ k .
E
Other inner products can be used in E, leading e.g. to Chebyshev or Jacobi polynomials.
All such bases are numerically stable. However, since these polynomials are non-zero at
both endpoints of the interval E, glueing them to form globally continuous functions on
the partition T leads to composite bases whose functions are supported over the whole
interval I.
• The interpolatory, or Lagrange basis Φ = {ϕ0 (x), ϕ1 (x), . . . , ϕk (x)} associated with k + 1
distinct point ξ0 < ξ1 < · · · < ξk in E. Each ϕ` ∈ Pk (E) is uniquely defined by the
conditions
ϕ` (ξi ) = δi,` , i = 0, 1, . . . , k .
Note that if E = [α, β] and we choose ξ0 = α, ξk = β, then only ϕ0 is non-zero at the left
endpoint of E, only ϕk is non-zero at the right endpoint of E, whereas all the other basis
functions ϕ` for 1 ≤ ` ≤ k − 1 vanish at both endpoints of E; a basis with such a property
is often called a boundary adapted basis. The property easily allows creating a composite
basis whose functions are supported on a single element, or at most on two contiguous
elements.
For k = 1, the basis functions are
β−x x−α
ϕ0 (x) = , ϕ1 (x) = , (3.1)
β−α β−α
Note that as soon as the polynomial degree k exceeds a few units, interpolation in E
should not be performed at equally spaced nodes (to avoid instabilities, namely intra-node
oscillations, as in the celebrated Runge phaenomenon), but rather at a family of nodes with
Gaussian-type spacing, such as e.g. the Gauss-Lobatto nodes which include the endpoints
of the element.
• The modal, or Babuška-Shen basis Ψ = {ψ0 (x), ψ1 (x), . . . , ψk (x)}, defined as follows:
ψ0 = ϕ0 and ψ1 = ϕ1 are the linear functions given in (3.1), whereas ψn for 2 ≤ n ≤ k are
primitives of Legendre polynomials, that vanish at both endpoints of E; precisely
Z β
ψn (x) = Ln−1 (s) ds = Ln (x) − Ln−2 (x) .
x
This is a subspace of L2 (I), of dimension (k + 1)N . Functions in such space need not be
continuous at the nodes of the partition (see Fig. 3.1.2, upper plot).
Let us now discuss how to create bases in these spaces, starting from bases in the polynomial
spaces on the elements.
• A basis in V(Th ; k) can be obtained by selecting a basis in each Pk (Ij ) and then extending it
by 0 outside Ij . For instance, a basis in the N -dimensional space V(Th ; 0) of the piecewise
constant functions on the partition is given by
(
1 if x ∈ Ij ,
ϕj (x) = 1≤j≤N.
0 otherwise ,
24 CHAPTER 3. DISCRETIZATIONS IN ONE SPACE VARIABLE
Figure 3.2: Examples of functions in V(Th ; 2) (upper plot), in V (Th ; 2) (middle plot), in V0 (Th ; 2)
(lower plot)
Figure 3.3: Basis functions in V (Th ; q): internal function ϕj (left plot), boundary functions ϕ0
and ϕN (center and right plot)
center and right plots). For 1 ≤ j ≤ N − 1, these functions are supported (i.e., different
from 0) only on the union Ij ∪ Ij+1 of two continguous elements, whereas ϕ0 is supported
only in I1 and ϕN is supported only in IN . The Lagrange representation of any function
v ∈ V (Th ; 1) is
XN
v(x) = v(xj )ϕj (x) , ∀x ∈ I .
j=0
Next, consider the (2N + 1)-dimensional space V (Th ; 2) of the continuous, piecewise
quadratic functions on T ; now each function v in this space is uniquely identified by
its values v(xj ) at the N + 1 nodes of the partition and v(xj−1/2 ) at the N midpoints
xj−1/2 = 21 (xj−1 + xj ) of the elements of the partition. Glueing and extending the in-
terpolatory basis functions given by (3.2) in each space P2 (Ij ), we obtain the composite
Lagrange basis Φ, formed by two families of functions: the so-called pagoda functions
(x − xj )(x − xj−1/2 )
2 if x ∈ Ij ,
h2j
ϕj (x) = (xj+1 − x)(xj+1/2 − x) 0≤j≤N (3.7)
2 2 if x ∈ Ij+1 ,
hj+1
0 otherwise ,
(with the obvious changes when j = 0 and j = N ), and the so-called bubble functions
(xj+1 − x)(x − xj ) if x ∈ Ij ,
ϕj−1/2 (x) = h2j 1≤j≤N (3.8)
0 otherwise ,
(see Fig. 3.1.2). With these basis functions, the Lagrange representation of any function
v ∈ V (Th ; 2) is
N
X N
X
v(x) = v(xj )ϕj (x) + v(xj−1/2 )ϕj−1/2 (x) , ∀x ∈ I .
j=0 j=1
26 CHAPTER 3. DISCRETIZATIONS IN ONE SPACE VARIABLE
Figure 3.4: The basis functions ϕj (upper plot) and ϕj−1/2 (lower plot) in V (Th ; 2)
• At last, a basis in V0 (Th ; k) can be obtained from the basis Φ in V (Th ; k) defined above,
simply by discarding the two basis functions ϕ0 and ϕN , which are the only ones that do
not vanish at both endpoints of I.
where I[x0 , x] ⊆ E is the interval whose endpoints are x0 and x. Using Hölder inequalities on
the right-hand side, we obtain
Z 1/2
|φ(x)| ≤ |x − x0 | max |φ0 (s)| or |φ(x)| ≤ |x − x0 |1/2 |φ0 (s)|2 ds .
s∈E E
From these inequalities, we can derive bounds of a norm of φ in terms of the length of the
interval E and a norm of φ0 . For instance,
Z 1/2
0 1/2 0 2
max |φ(x)| ≤ hE max |φ (s)| or max |φ(x)| ≤ hE |φ (s)| ds ,
x∈E s∈E x∈E E
Such inequalities extend to continuous functions with distributional first derivative in L2 (E)
or in L∞ (E), since they can be approximated arbitrarily well by smooth functions (classical
functions are dense in Sobolev spaces). In this way, one can obtain bounds as stated below.
Property 3.2.1 Let φ vanish in E and let φ0 ∈ L2 (E). Then,
1/2
kφkL2 (E) ≤ √1
2
hE kφ0 kL2 (E) and kφkL∞ (E) ≤ hE kφ0 kL2 (E) .
• Case k = 0
Let vI = IE0 v ∈ P0 (E) be the constant function that satisfies vI = v(x0 ) for some x0 ∈ E.
Then, applying Property 3.2.1 to the function φ = v − vI and noticing that vI0 = 0, we get
kv − vI kL2 (E) ≤ √12 hE kv 0 − vI0 kL2 (E) = √12 hE kv 0 kL2 (E) , which we can write as
• Case k = 1
Let vI = IE1 v ∈ P1 (E) be the linear function that interpolates v at the endpoints of E, i.e.,
vI (α) = v(α) and vI (β) = v(β) Then, assuming that v ∈ H 2 (E) and applying Property
3.2.1 to the function ϕ = v − vI , we get
kv − vI kL2 (E) ≤ √1
2
hE kv 0 − vI0 kL2 (E) . (3.9)
On the other hand, by Rolle’s theorem, the function ψ = v 0 − vI0 vanishes at some point s0
in E; applying once more Property 3.2.1 to this function, and noticing that now vI00 = 0,
we get kv 0 − vI0 kL2 (E) ≤ √12 hE kv 00 kL2 (E) . Concatenating the two previous inequalities, we
get
• Case k = 2
Let vI = IE2 v ∈ P2 (E) be the quadratic function that interpolates v at the endpoints of
E and at the baricenter γ = 12 (α + β) of E. Then, assuming that v ∈ H 3 (E), applying
Property 3.2.1 to the functions φ = v − vI , ψ = v 0 − vI0 and η = v 00 − vI00 and noticing that
now vI000 = 0, we arrive at the inequalities
• Case k > 2
Let vI = IEk v ∈ Pk (E) be the polynomial that interpolates v at k + 1 distinct points in
E. Iterating the argument above, one can prove the following error estimates: for any `, r
satisfying 0 ≤ ` < r ≤ k + 1, if v ∈ H r (E) one has
Remark 3.2.2 The order of decay of the interpolation error, i.e., the power of hE in the
previous estimates, depends not only upon the polynomial degree k, but also upon the (Sobolev)
smoothness of the function v. For instance, if v only belongs to H 1 (E) but not to H 2 (E), then
the interpolation error using a linear polynomial can only be bounded as the interpolation error
using a constant polynomial, i.e.,
for some constant c1 > 0. This bound easily follows from (3.9), observing that vI = IE1 v satisfies
vI0 = v(β)−v(α) = h1E E v 0 (s) ds, which immediately gives kvI0 kL2 (E) ≤ kv 0 kL2 (E) .
R
β−α
Similarly, for the interpolation that uses a quadratic polynomial, we have the error bounds
kv − IE2 vkL2 (E) ≤ c3 h2E |v|H 2 (E) and |v − IE2 v|H 1 (E) ≤ c4 hE |v|H 2 (E) .
• Let us start with the linear case (k = 1). For any subinterval Ij (1 ≤ j ≤ N ) of the
partition Th , let vIj = II1j v ∈ P1 (Ij ) be the linear interpolant of v at the endpoints of Ij ,
i.e., vIj (xj−1 ) = v(xj−1 ) and vIj (xj ) = v(xj ). Then, we define
which lead to the following representation in terms of the Lagrangean basis defined in (3.6)
N
X
(Ih1 v)(x) = v(xj )ϕj (x), ∀x ∈ I.
j=0
3.2. APPROXIMATION ERRORS 29
The error v − Ih1 v can be estimated using the bounds (3.10) elementwise. Precisely, as-
suming that v|Ij ∈ H 2 (Ij ) for all j, one has
N
X N
X N
X
kv − Ih1 vk2L2 (I) = k(v − Ih1 v)|Ij k2L2 (Ij ) = kv − vIj k2L2 (Ij ) ≤ 1
4 h4j |v|2H 2 (Ij ) ,
j=1 j=1 j=1
1 2
= 2 h |v|H 2 (I) if v ∈ H 2 (I) . (3.14)
= √1 h |v|H 2 (I)
2
if v ∈ H 2 (I) . (3.16)
Remark 3.2.3 On the right-hand sides of these estimates, terms like (3.13) or (3.15)
provide a more accurate information than terms like (3.14) or (3.16). Indeed, the former
terms show the interplay between the size hj of the elements and the regularity of v inside
the elements. Such expressions suggest to use smaller elements where derivative of v are
larger, thus providing a rationale for mesh adaptivity.
For the sake of simplicity, in the sequel we will mostly present error estimates with terms of
type (3.14) or (3.16) on the right-hand side, but the reader should always keep in mind that
more accurate expressions could be written instead.
Assuming that v ∈ H 3 (I) and using the bounds (3.11) elementwise, we obtain as above
the following error estimates
Note that it would be meaningless to estimate the error in the H 2 (I)-seminorm, since the
piecewise interpolant Ih2 v is continuous but in general exhibits jumps in the first derivative
at the nodes of the partition, i.e., it does not belong to H 2 (I). Also note that if v does not
belong to H 3 (I) but only to H 2 (I), then recalling Remark 3.2.2 the previous estimates
must be replaced by
kv − Ih2 vkL2 (I) ≤ c3 h2 |v|H 2 (I) , |v − Ih2 v|H 1 (I) ≤ c4 h |v|H 1 (I) .
and we assume that v ∈ H r (I) for some r satisfying 1 ≤ r ≤ k + 1, we have the following
general error estimates
kv − Ihk vkL2 (I) ≤ c0,r,k hr |v|H r (I) , |v − Ihk v|H 1 (I) ≤ c1,r,k hr−1 |v|H r (I) , (3.18)
Remark 3.2.4 The previous estimates concern a so-called h-type approximation method, mean-
ing that the quality of the approximation is dictated by the fineness of the partition Th , whereas
the polynomial degree is kept fixed. One can also consider a p-type method (sometimes called
a spectral method), in which the partition is kept fixed, whereas the polynomial degree, often
denoted by pj , is allowed to increase in each element Ij .
Combining the two stategies leads to the so-called hp-type methods, in which a reduction of
the approximation error is achieved by simultaneously refining the partition and increasing the
polynomial degrees. In this case, the typical error estimates (assuming for simplicity to have
the same polynomial degree p in each element) are
hmin(p+1,r) hmin(p+1,r)−1
kv − Ihp vkL2 (I) ≤ c0,r kvkH r (I) , |v − Ihp v|H 1 (I) ≤ c1,r kvkH r (I) ,
pr pr−1
for v ∈ H r (I) with r ≥ 1 and p ≥ 1.
Remarkably, when the solution is analytic in I, the error can be shown to decay exponentially
in p.
− (µ u0 )0 + β u0 + σu = f in (0, L) ,
(3.19)
u(0) = u(L) = 0 .
3.3. A FIRST EXAMPLE OF GALERKIN DISCRETIZATION 31
Note that L = −(µ u0 )0 + β u0 + σu is the one-dimensional version of the general linear second-
order operator introduced in (1.4). We assume that µ, β, σ ∈ L∞ (I), f ∈ L2 (I) and, in addition,
µ ≥ µ0 in I, β 0 ∈ L∞ (I) and − 21 β 0 + σ ≥ σ0 ≥ σ0 in I, for certain constants µ0 > 0 and σ0 ≥ 0.
Under these conditions, the bilinear form
Z
a(u, v) = (µ u0 v 0 + β u0 v + σ u v) (3.20)
I
To discretize it, we choose the subspace Vh = V0 (Th ; 1) of the continuous, piecewise linear
functions on a partition Th of the domain I, which vanish at the endpoints of I (see (3.5)), and
we formulate the Galerkin problem
Using the Lagrangean basis Φ = {ϕj : 1 ≤ j ≤ N − 1} in V0 (Th ; 1), we transform this problem
into the algebraic problem
Au = f ,
where, according to Sect. 2.2,
Z
A = (ajk )1≤j,k≤N −1 , ajk = (µ ϕ0k ϕ0j + β ϕ0k ϕj + σ ϕk ϕj ) ,
I
Z
f = (fj )1≤j≤N −1 , fj = f ϕj ,
I
and u = (uk )1≤k≤N −1 with uk = uh (xk ). Observe that the stiffness matrix A is the sum of three
matrices, the diffusion matrix D, the convection matrix C, and the reaction matrix R, whose
entries are Z Z Z
0 0 0
djk = µ ϕk ϕj , cjk = β ϕk ϕj , σ ϕk ϕj ,
I I I
namely,
A = D + C + R.
Recalling that supp ϕj = Ij ∪ Ij+1 for 1 ≤ j ≤ N − 1, each integral is 0 when the supports of
ϕj and ϕk are disjoint or intersect at a point only; these situations occur when |j − k| > 1. In
other words,
djk = cjk = rjk = 0 whenever k 6= j − 1, j, j + 1,
which gives the following
Property 3.3.1 The matrices D, C, R, and consequently A, are tridiagonal.
We now proceed to the computation of the entries of these matrices, as well as the right-hand
side f .
32 CHAPTER 3. DISCRETIZATIONS IN ONE SPACE VARIABLE
• Computation of D.
At first, let us compute dj,j−1 (for j > 1). The intersection of the supports of ϕj−1 and
ϕj is the interval Ij , hence Z
dj,j−1 = µ ϕ0j−1 ϕ0j .
Ij
On Ij the basis functions are linear, so their derivatives are constant; to be precise, by
(3.6) we have
1 1
ϕ0j−1 = − , ϕ0j = ,
hj hj
whence Z
1 1
dj,j−1 = − 2 µ = − µ̄j−1/2 ,
hj Ij hj
where µ̄j−1/2 = h1 j Ij µ dx is the mean value over Ij of the elastic coefficient µ. When this
R
number cannot be computed exactly, we can approximate it: let µj−1/2 denote the com-
puted approximation. For instance, using the mid-point quadrature rule to approximate
the integral, we obtain
µj−1/2 = µ(xj−1/2 ) where xj−1/2 = 21 (xj−1 + xj ) is the mid-point of Ij ;
using the trapezoidal rule instead, we get
µj−1/2 = 12 (µ(xj−1 ) + µ(xj )) .
In conclusion, we set
1
dj,j−1 = − µ .
hj j−1/2
Next, let us compute the diagonal entry djj . Since
Z Z
0 2
djj = µ (ϕj ) + µ (ϕ0j )2
Ij Ij+1
with
1 1
ϕ0j = on Ij , ϕ0j = − on Ij+1 ,
hj hj+1
the same approximation for the coefficient µ used above gives
1 1
djj = µj−1/2 + µ .
hj h j+1 j+1/2
Finally, by the symmetry of the matrix we get
1
dj,j+1 = dj+1,j = dj+1,(j+1)−1 = − µj+1/2 .
hj+1
Remark 3.3.2 It is worth noticing the property that the sum of the entries in each row
j, except the first and last one, is zero:
N
X −1
djk = 0 , 2 ≤ j ≤ N − 2.
k=1
Indeed,
N −1 N −1
!0
X Z xj+1 X
djk = µ ϕk ϕ0j ;
k=1 xj−1 k=1
PN −1
the linear function k=1 ϕk takes the value 1 at each internal node, hence it is identically
equal to 1 in any interval Ij , except the first and last one. Therefore, its derivative is
identically zero in the integration interval [xj−1 , xj+1 ].
Remark 3.3.3 In the relevant case in which the coefficient µ is constant and the mesh is
L
equally spaced (i.e. hj = h = N for any j), the matrix D has the simple structure
µ
D= tridiag [ −1 2 − 1] ,
h
where the symbol tridiag [a b c] denotes the tridiagonal matrix having constant value a
on the first lower diagonal, b on the main diagonal and c on the first upper diagonal.
• Computation of C.
We have Z
cj,j−1 = β ϕ0j−1 ϕj
Ij
βj−1/2
Z Z
1 1 hj
cj,j−1 = βj−1/2 ϕ0j−1 ϕj = βj−1/2 − ϕj = βj−1/2 − =− ,
Ij hj Ij hj 2 2
Finally,
βj+1/2
Z
1 hj+1
cj,j+1 = βj+1/2 ϕ0j+1 ϕj = βj+1/2 = .
Ij+1 hj+1 2 2
34 CHAPTER 3. DISCRETIZATIONS IN ONE SPACE VARIABLE
Note that, again, sum of the entries in each row j, except the first and last one, is zero.
If β is constant then C reduces to
β
C= tridiag [−1 0 1] ,
2
which is a skew-symmetric matrix (i.e., C T = −C).
• Computation of R.
We have Z
rj,j−1 = σ ϕj−1 ϕj ,
Ij
We use (3.6) for the maps ϕj−1 and ϕj on the interval Ij , and we introduce the change
the integration variable x = xj−1 + shj , 0 ≤ s ≤ 1, so that dx = hj ds. Then,
Z Z xj
1
ϕj−1 ϕj dx = (xj − x)(x − xj−1 ) dx
Ij h2j xj−1
1 3 1 1 2 1 3 1 1
Z
= h (1 − s)s ds = hj s − s = hj
h2j j 0 2 3 0 6
and so
1
rj,j−1 = σj−1/2 hj .
6
For the diagonal term Z Z
rjj = σϕ2j dx + σϕ2j dx ,
Ij Ij+1
The special case in which σ is constant on [0, L] and the subdivision is equidistant with
stepsize h yields the simplified form
1 2 1
R = σh tridiag 6 3 6 . (3.27)
• Computation of f .
We have Z Z
fj = f ϕj + f ϕj .
Ij Ij+1
Rb
A popular alternative to this formula is obtained by applying the trapezoidal rule a g '
1
2 (g(a) + g(b))(b − a) to the function g = f ϕj in each interval; in this case one sets
1 1
fj = f (xj−1 )ϕj (xj−1 ) + f (xj )ϕj (xj ) + f (xj+1 )ϕj (xj+1 ) ;
2 2
recalling the ϕj vanishes at xj−1 and xj+1 , and equals 1 at xj , we obtain the formula
hj + hj+1
fj = f (xj ) . (3.28)
2
Assuming that u ∈ H r (I) for some r satisfying 1 ≤ r ≤ k + 1, we have from the interpolation
error bounds (3.18)
Since h ≤ |I| = L, we deduce the existence of a constant Cr,k > 0 such that
Proposition 3.4.1 Under the previous assumptions, the Galerkin discretization error satisfies
the a priori bound
ku − uh kH 1 (I) ≤ C hr−1 |v|H r (I) , (3.29)
where C = Ca Cr,s and 1 ≤ r ≤ k + 1.
ku − uh kH 1 (I) ∼ K hq (3.32)
where K and q depend upon the regularity of the exact solution and the polynomial degree used
in the discretization. Indeed, this is the kind of observed behavior for a generic solution, when
the mesh-size is successively reduced.
A simple interpretation of this law is as follows. Let εh = ku − uh kH 1 (I) be the error produced
L
by a partition of mesh-size h (to fix the ideas, a uniform partition with h = N ), and let
3.4. ERROR ANALYSIS 37
h
εh/2 = ku − uh/2 kH 1 (I) be the error produced by the partition of mesh-size 2 obtained by
halving each element. Applying (formally) the law (3.32), we get
εh K hq
∼ = 2q ,
εh/2 K ( h2 )q
i.e., the effect of halving h is to divide the error roughly by 2q (e.g., by 2 in the linear case
(3.30) and by 4 in the quadratic case (3.31)). If a known solution is available, and one is
able to compute the discretization errors εh , εh/2 , εh/4 , . . . for successive refinements, one can
experimentally produce an approximation of the exponent q, and check the validity of the law
(3.32).
A more sophisticated strategy leads to identifying both K and q in (3.32). Taking the loga-
rithm of both sides, one formally obtains
which shows that in a log-log scale, h and εh are placed along a straight line with slope equal
to q. A linear regression, based on the computation of εh for several values of h, yields an
approximation of K and q.
1 F (v − vh ) − a(uh , v − vh )
ku − uh kH 1 (I) ≤ sup , ∀vh ∈ Vh . (3.33)
α v∈V \{0} kvkH 1 (I)
Let us choose vh = Ih1 v ∈ Vh , the piecewise linear interpolant of v at the nodes of the partition,
and let us set w = v − Ih1 v for convenience. Then, recalling (3.20)-(3.21), we can write
Z
F (w) − a(uh , w) = f w − (µ u0h w0 + β u0h w + σ uh w)
I
N Z !
X
= f w − (µ u0h w0 + β u0h w + σ uh w)
j=1 Ij
N Z !
X
= f w − (−(µ u0h )0 w + β u0h w + σ uh w)
j=1 Ij
N Z
X
= (f − Luh ) w
j=1 Ij
where the performed integration by parts does not produce boundary terms since w vanishes at
all nodes of the partition. Recalling (3.12), we can bound each integral as
Z
(f − Luh ) w ≤ k(f − Luh )kL2 (Ij ) kwkL2 (Ij ) ≤ c1 hj k(f − Luh )kL2 (Ij ) |v|H 1 (Ij ) ,
Ij
38 CHAPTER 3. DISCRETIZATIONS IN ONE SPACE VARIABLE
whence,
1/2 1/2
XN XN
F (w) − a(uh , w) ≤ c1 h2j k(f − Luh )k2L2 (Ij ) |v|2H 1 (Ij )
j=1 j=1
1/2
XN
= c1 h2j k(f − Luh )k2L2 (Ij ) |v|H 1 (I) .
j=1
Then, the following a posteriori upper bound for the Galerkin error holds true:
c1
ku − uh kH 1 (I) ≤ ηh . (3.35)
α
The quantity ηh is computable in practice, to the extent that the integrals expressing the L2 -
norms in (3.34) are computable. This occurs, e.g., if the forcing term f and the coefficients µ, β,
σ of the operator are piecewise polynomial on the partition; otherwise, one should approximate
them by piecewise polynomials, introducing an extra term in the bound above, called data
oscillation. For simplicity, we assume here that there is no data oscillation.
The information provided by the estimator ηh can be used to define a strategy of mesh
adaptation in which, starting from an initial partition T (0) = Th0 , one generates a sequence of
partitions T (`) = Th` , ` = 1, 2, . . . by subsequent refinements based on the estimator, until a
target accuracy is reached. Often, one step of adaptation is expressed by the chain
Indeed, suppose that, for some chosen tol > 0, we want to be guaranteed that
ii) the marked elements are Ij1 , Ij2 , . . . , IjM , where M ≤ N is the smallest integer for which
M
X N
X
2 2
ηh,j m
≥ϑ ηh,j
m=1 j=1
hT = max hE ; (4.1)
E∈T
we will set h = hT when no confusion is possible.
41
42 CHAPTER 4. GENERAL FINITE ELEMENTS
1. E is a non-empty compact and connected set in Rd , such that E = E̊ and the boundary
∂E is Lipschitz-continuous.
Note that by (4.5) ΠE v = v for all v ∈ VE , which proves that ΠE is indeed a projection upon
VE .
A finite element in which each degree of freedom is the value of v at a selected point in E is
termed a Lagrangian finite element. The projection operator ΠE is the Lagrange interpolation
operator at the set of selected nodes.
A finite element in which each degree of freedom is the value of v or of some partial derivative
of v at a selected point in E is termed a Hermitian finite element.
ii) If v̂ and v are real functions defined on subsets of Rd , the symbol v ↔ v̂ means v = v̂ ◦ F−1
or, equivalently, v̂ = v ◦ F; in other words, it holds v(x) = v̂(x̂) whenever x ↔ x̂.
iii) At last, if `ˆ and ` are linear forms defined on spaces of real functions, the symbol ` ↔ `ˆ
means `(v) = `(v ˆ ◦ F) for any v or, equivalently, `(v̂)
ˆ = `(v̂ ◦ F−1 ) for any v̂; in other
ˆ
words, it holds `(v) = `(v̂) whenever v ↔ v̂.
E = F(Ê) ,
VE = {v : v ↔ v̂ for some v̂ ∈ V̂ } , (4.7)
LE = {`j : `j ↔ `ˆj with `ˆj ∈ L̂, 1 ≤ j ≤ N̂ } ,
where N̂ = dim V̂ . It is easily checked that dim VE = dim V̂ and that LE is unisolvent for VE .
Thus, the axioms of finite element are satisfied.
Furthermore, it is immediate to check that the canonical bases in VE and V̂ are related by
ϕj ↔ ϕ̂j , 1 ≤ j ≤ N̂ , (4.8)
ΠE v ↔ Π̂ v̂ whenever v ↔ v̂ , (4.9)
If the degrees of freedom in L̂ are defined on a larger space V̂ ⊃ V̂ , then the degrees of freedom
in LE are defined on VE = {v : v ↔ v̂ for some v̂ ∈ V̂} ⊃ VE .
4.4 Examples
In this section, we will present several examples of finite elements and discuss their properties.
In preparation for that, we begin by establishing a useful result, which extends a well-known
property of polynomials in one space variable, namely: if v is a polynomial of degree k ≥ 1 in one
space variable satisfying v(x0 ) = 0 for some x0 ∈ R, then it factorizes as v(x) = (x − x0 )w(x),
where w is a polynomial of degree k − 1; if in addition v 0 (x0 ) = 0, then v factorizes as v(x) =
(x − x0 )2 u(x), where u is a polynomial of degree k − 2.
Lemma 4.4.1 Let L be a segment in the plane, not reduced to a point. If v belongs to Pk (R2 )
(for some k ≥ 1) and vanishes on L, then v factorizes as
where eL (x, y) = 0 is the equation of the line containing L. If in addition the normal derivative
∂v
to the edge ∂n vanishes on L, then v factorizes as
Proof. Assume first that the segment L lies on the y-axis. Since v(0, y) is a polynomial in y
that vanishes at the infinite points of L, it vanishes identically on the y-axis. Thus, for any fixed
y, x 7→ v(x, y) is a polynomial in one variable that vanishes at x = 0, hence, it can be factorized
as v(x, y) = x w(x, y), where w is a polynomial of degree ≤ k − 1. Since x = 0 is precisely the
equation of the y-axis, the first result is proven.
If L is arbitrarily placed in the plane, let us fix an orthogonal coordinate system ξ-η such
that L lies on the η-axis, and let (ξ, η) = Φ(x, y) be the affine change of variable between the
two systems of coordinates. Then, ṽ(ξ, η) = v(Φ−1 (ξ, η)) is a polynomial of degree ≤ k in the
variables ξ, η that vanishes on a segment on the axis ξ = 0, hence it factorizes as ṽ(ξ, η) =
ξ w̃(ξ, η) with w̃ polynomial of degree ≤ k − 1. Going back to the original coordinates, we get
v(x, y) = ξ(x, y) w(x, y), where ξ(x, y) = 0 is precisely the equation of the η-axis expressed in
the x, y-coordinates.
The proof of the second result can be done by similar arguments. The extension to the
three-dimensional case is trivial.
xB = 13 (a1 + a2 + a3 )
T̂ = {(x, y) ∈ R2 : 0 ≤ x, y, x + y ≤ 1} .
`(v) = v(xB )
(2)
• The set LT of the values at the midpoints of the edges:
(3)
• The set LT of the integral averages on the edges:
Z Z Z
1 1 1
v, v, v.
|L1 | L1 |L2 | L2 |L3 | L3
(4)
• The set LT of the moments of order ≤ 1 on the triangle:
Z Z Z
1 1 1
v, vx , vy .
|T | T |T | T |T | T
With any of these choices for LT , the triple (T, P1 (T ), LT ) satisfies the axioms of finite element.
(1)
Let us check that the set LT of degrees of freedom is unisolvent for P1 (T ). There are
many ways to do this; here, let us apply Lemma 4.4.1. Suppose that v ∈ P1 (T ) satisfies
v(a1 ) = v(a2 ) = v(a3 ) = 0. Then, v vanishes identically on each edge of the triangle. Indeed,
let us consider for instance the edge L1 = [a1 , a2 ] and let
be the affine parametrization of this edge. The function η(t) = v(x(t)) is a polynomial of degree
≤ 1 in the variable t, and satisfies η(0) = v(a1 ) = 0 and η(1) = v(a2 ) = 0; hence, it vanishes
identically, proving that v|L1 vanishes identically. By Lemma 4.4.1, we can write v = eL1 w with
w a polynomial of degree 0, i.e., a constant. If we compute v at a3 , we have
and eL1 (a3 ) 6= 0 since a3 does not lie on the line containing L1 . Hence, necessarily w = 0, which
proves that v is identically zero.
(1)
The finite element (T, P1 (T ), LT ) is called the Courant element.
(2)
A similar argument proves that the set LT is unisolvent for P1 (T ). In general, so is the set
of degrees of freedom given by the values of v at any three points in T which are not aligned.
The corresponding finite elements are of Lagrangian type. Such degrees of freedom are defined
on VT = C 0 (T ).
(3) (4)
Example 4.4.2 Prove that the sets LT and LT are unisolvent for P1 (T ).
4.4. EXAMPLES 47
Figure 4.1: Various examples of finite elements with their degrees of freedom
We choose the space VE = VT = P3 (T ), whose dimension is 10. We propose two sets of degrees
(1)
of freedom. The first one, LT , is given by the values
v(aij ) , 0 ≤ i, j, i + j ≤ 3 ,
at the ten nodes in T defined as follows: divide each edge into three segments of equal length,
then draw the three family of lines, each one parallel to an edge, passing through the subdivision
points (including the vertices) (see Figure...). The desired nodes are the points in T where three
lines intersect.
Extending the arguments given above for the quadratic case to the cubic case, one can show
(1) (1)
that LT is unisolvent for P3 (T ), hence (T, P3 (T ), LT ) is a Lagrangian finite element.
Remark 4.4.3 More generally, the previous construction of nodes can be done starting from
the partition of each edge into k > 3 segments of equal length. One obtains a total of (k+1)(k+2)
2
nodes aij with 0 ≤ i, j and i + j ≤ k. The values of v at these nodes form a unisolvent set of
Lagrangian degrees of freedom in Pk (T ).
i j
In the reference element Ê = T̂ , the coordinates of these nodes are âij = k, k .
The use of such finite elements is limited to small values of k. Indeed, since the nodes are
uniformly spaced in the element, the corresponding interpolation operator suffers from numerical
instability for increasing values of k.
(2)
The second set, LT , of degrees of freedom for P3 (T ) leads to our first example of a Hermitian
finite element. The ten degrees of freedom are the values of v and its gradient at the three
vertices, plus the value of v at the barycenter, i.e.,
Let us check that this set is unisolvent. Suppose that v ∈ P3 (T ) has all such degrees of freedom
equal to 0. Consider again the parametrization (4.11) of the edge L1 = [a1 , a2 ]. In this case,
the function η(t) = v(x(t)) is a polynomial of degree ≤ 3 in the variable t; its derivative with
respect to t is given by η 0 (t) = ∇v(x(t)) · (a2 − a1 ). One has
which implies that η factorizes as η(t) = t2 (1 − t)2 u(t); but since the degree of η is ≤ 3,
necessarily u, hence η, is identically 0. This proves that v vanishes on L1 , and similarly on L2
and L3 . Applying Lemma 4.4.1 repeatedly, we obtain the factorization v = eL1 eL2 eL3 w, where
w is a constant. Finally, we use the last condition,
to deduce that necessarily w = 0 since xB does not lie on any edge; hence, v is identically 0.
(2)
For the Hermitian element (T, P3 (T ), LT ), it is natural to choose VE = VT = C 1 (T ).
4.4. EXAMPLES 49
VT = {v ∈ C 1 (T ) : v|Ti ∈ P3 (Ti ), i = 1, 2, 3} .
Obviously, P3 (T ) ⊂ VT . But the latter space contains functions that are not global polynomials,
since
dim VT = 12 .
Indeed, to build a function in this space, one may start from three independent polynomials
vi ∈ P3 (Ti ), i = 1, 2, 3, which globally depend on 30 coefficients, and enforce the continuity
of these functions and their gradients across the interfaces between triangles; one can show
that such a ‘glueing’ process amounts to enforcing 18 independent linear conditions, whence the
dimension 12 = 30 − 18 of VT .
As a set LT of degrees of freedom, one chooses the values of the function v and its gradient
at the vertices of the triangle, plus the values of the normal derivative at the midpoints of the
edges, i.e.,
∂v
v(aj ) , ∇v(aj ) , (aj,j+1 ) for 1 ≤ j ≤ 3 .
∂n
One can prove that this set of degrees of freedom in unisolvent for VT , which makes the triple
(T, VT , LT ) a Hermitian finite element. It is called the Hsieh-Clough-Tocher element, or
simply the HCT element. We will come back to this element later on.
∂v
v(aj ) , ∇v(aj ) , Hv(aj ) , (aj,j+1 ) for 1 ≤ j ≤ 3 .
∂n
n 2 o
Note that the Hessian matrix Hv = ∂x∂p ∂x v
q
is symmetric by Schwarz’ theorem, thus prescrib-
ing it at a point amounts to prescribe 3 linear conditions. It follows that LT contains exactly
21 degrees of freedom.
In order to prove that it is unisolvent, consider a function v ∈ P5 (T ) that has all such degrees
of freedom equal to 0. Let us use again the parametrization (4.11) of the edge L1 = [a1 , a2 ]. In
this case, the function η(t) = v(x(t)) is a polynomial of degree ≤ 5 in the variable t; its first
derivative with respect to t is given by η 0 (t) = ∇v(x(t)) · (a2 − a1 ), whereas its second derivative
is given by η 00 (t) = (a2 − a1 ) · Hv(x(t))(a2 − a1 ). Hence, one has
which implies that η factorizes as η(t) = t3 (1 − t)3 %(t) for some %; but since the degree of η is
≤ 5, necessarily %, hence η, is identically 0. This proves that v vanishes on L1 .
50 CHAPTER 4. GENERAL FINITE ELEMENTS
xB = 41 (a1 + a2 + a3 + a4 )
R̂ = {(x, y) ∈ R2 : 0 ≤ x, y ≤ 1} .
where ∂i2 v(ai ) denotes the second partial derivative of v taken along the two directions of the
edges that meet at ai .
4.4.3 3D elements
... ...
In many cases, we will need subspaces of this space, formed by functions with some kind of
global continuity (in the classical sense), or regularity (in the modern functional sense). So,
subspaces of interest could be for instance
Assuming that functions in each VE have the requested continuity or regularity, we are faced with
the problem of ‘glueing together’ functions across the interfaces between contiguous elements.
To be precise, suppose that E1 and E2 are two elements of the partition, which share a full
edge (in 2D) or face (in 3D); let L = ∂E1 ∩ ∂E2 denote this common part. Given two functions
v1 ∈ VE1 and v2 ∈ VE2 , let us denote by v the function defined in E1 ∪ E2 as
(
v1 on E1 ,
v= (4.13)
v2 on E2 .
• In particular, can we identify degrees of freedom that are local on each edge/face, so that
if we give them the same value on both sides of an interface between two elements, we
enforce automatically the desired continuity?
(3) (4)
Similar considerations apply to the other sets of degrees of freedom LT and LT discussed in
(i)
Case B of the previous section. Thus, none of the finite elements (T, P1 (T ), LT ) with i = 2, 3, 4
is of class C 0 .
We are ready for the following definition.
Definition 4.5.1 Suppose that VE ⊂ C 0 (E). We say that the finite element (E, VE , LE ) is of
class C 0 if for any edge/face L of E there exists a subset LE,L ⊂ LE of degrees of freedom with
the following properties:
i) each form ` ∈ LE,L depends only on the values taken on L by the argument v;
where λ is a linear form on the space VE,L = {v|L : v ∈ VE } of the restrictions to L of the
functions in VE .
Item ii) means that the set ΛE,L collecting all such forms λ is unisolvent for VE,L .
Examples 4.5.2 Each of the finite elements presented in Cases from C to K of the previous
section is of class C 0 .
Indeed, for the quadratic element presented in Case C, the degrees of freedom relative to the
edge L are
v(a1 ), v(a2 ), v(a1,2 ) ,
which form a unisolvent set for the space P2 (L) = (P2 (T ))|L . Similar considerations apply to
the cubic Lagrangian element presented in Case D, and to their higher-order extensions.
For the cubic Hermitian element presented in Case D and the composite Hsieh-Clough-Tocher
element presented in Case E, the degrees of freedom relative to the edge L provide the values
of the forms
∂v ∂v
v(a1 ), v(a2 ), (a1 ), (a2 ) ,
∂τ ∂τ
which form a unisolvent set for the space P3 (L) = (P3 (T ))|L (for the HCT element, remember
that functions in VT are polynomials of degree ≤ 3 near any edge of T ). Note that the value of
∂v
the tangential derivative ∂τ can be obtained from the assigned value of the gradient ∇v by the
∂v
formula ∂τ = ∇v · τ .
For the Argyris element presented in Case F, the degrees of freedom relative to the edge L
provide the values of the forms
∂v ∂v ∂2v ∂2v
v(a1 ), v(a2 ), (a1 ), (a2 ) (a1 ), (a2 ) ,
∂τ ∂τ ∂τ 2 ∂τ 2
which form a unisolvent set for the space P5 (L) = (P5 (T ))|L . Here, the second tangential
∂v 2
derivative is given by ∂τ 2
= τ · (Hv)τ .
54 CHAPTER 4. GENERAL FINITE ELEMENTS
With such definition at hand, let us go back to the two elements E1 and E2 of the partition
of Ω, which share a common edge/face L. Assume that the finite elements (E1 , VE1 , LE1 ) and
(E2 , VE2 , LE2 ) are of class C 0 . Furthermore, assume that the space of the restrictions to L of
the functions in VE1 and in VE2 coincide, i.e.,
(remember (4.14) after Definition 4.5.3). Given two functions v1 ∈ VE1 and v2 ∈ VE2 , let us
equate all the degrees of freedom on L, namely let us enforce
λ(v1 |L − v2 |L ) = 0 ∀λ ∈ ΛL ,
v1 |L − v2 |L = 0 , i.e., v1 |L = v2 |L .
We conclude that the matched function v defined in (4.13) is continuous across L. This argument
shows that the matching (4.15) of the local degrees of freedom on L guarantees C 0 -continuity in
E1 ∪ E2 .
Definition 4.5.3 Suppose that VE ⊂ C 1 (E). We say that the finite element (E, VE , LE ) is of
class C 1 if it is of class C 0 and if for any edge/face L of E there exists a subset L0E,L ⊂ LE of
degrees of freedom with the following properties:
i) each form ` ∈ L0E,L depends only on the values taken on L by the normal derivative ∂v
∂n of
the argument v;
Examples 4.5.4 Obviously, a Lagrangian finite element cannot be of class C 1 , since one cannot
infer the values of the normal derivative on an edge by the assigned degrees of freedom.
Consider at first the cubic Hermitian element presented in Case D. The degrees of freedom
relative to the edge L provide the values of the two forms
∂v ∂v
(a1 ) and (a2 ) ;
∂n ∂n
4.5. FINITE ELEMENTS OF CLASS C 0 AND C 1 55
however, if v ∈ P3 (T ), its normal derivative on L belongs to P2 (L), which has dimension 3. The
∂v
two forms do not allow us to uniquely identify ∂n on L. Consequently, the cubic Hermitian
element presented in Case D is not of class C . 1
It is precisely this pitfall that suggested to enrich the space P3 (T ), leading to the introduc-
tion of the composite Hsieh-Clough-Tocher element presented in Case E. Indeed, its degrees of
freedom relative to the edge L provide the values of the forms
∂v ∂v ∂v
(a1 ), (a1,2 ), (a2 ) ,
∂n ∂n ∂n
which form a unisolvent set for P2 (L). Hence, the Hsieh-Clough-Tocher element if of class C 1 .
At last, consider the quintic Argyris element presented in Case F, and note that if v ∈ P5 (T ),
its normal derivative on L belongs to P4 (L). The degrees of freedom relative to the edge L
provide the values of the forms
∂v ∂v ∂v ∂2v ∂2v
(a1 ), (a1,2 ), (a2 ) , (a1 ), (a2 ) ,
∂n ∂n ∂n ∂τ ∂n ∂τ ∂n
which form a unisolvent set for P4 (L). Hence, the Argyris element if of class C 1 .
The argument given above about the C 0 -matching of two functions v1 ∈ VE1 and v2 ∈ VE2
across a common interface L can be easily extended to the case of C 1 -matching. Assuming that
the same finite element of class C 1 is used in E1 and E2 , it is enough to equate all the local
degrees of freedom on L to satisfy the matching conditions
∂v1 ∂v2
v1 |L = v2 |L and = .
∂n |L ∂n |L
`j (ϕj ) = 1 , `k (ϕj ) = 0 ∀k 6= j .
The construction of the basis functions can be done locally, element by element. Local basis
functions on contiguous elements, associated with the same degree of freedom, are automatically
matched across the common interface.
For instance, if one uses the Courant element in the definition of the space V (T ) := V(T ) ∩
C 0 (Ω) = V(T ) ∩ H 1 (Ω), the global pyramidal basis function ϕj associated with a vertex aj of
56 CHAPTER 4. GENERAL FINITE ELEMENTS
the triangulation is obtained by glueing together the local linear basis functions associated with
aj in each triangle having aj as a vertex.
The global basis functions obtained in this way have local support: if the function ϕj is
associated as above with the degree of freedom `j , it will be identically 0 in all elements, except
in those which include `j in their set L of degrees of freedom.
With similar arguments, one can see that a global canonical projection operator ΠT can be
defined by glueing together the local projections operators in each element; precisely, ΠT v is
such that
(ΠT v)|E = ΠE v|E ∀E ∈ T .
Interelement continuity is automatically guaranteed by the assumption that the used finite
elements are of class C 0 or C 1 .
For instance, if the Courant element is used, ΠT is the global interpolation operator by
piecewise affine functions: in each T ∈ T , (ΠT v)|T is the affine function which interpolates v at
the three vertices of the triangle.
Chapter 5
Given an admissible triangulation T in the domain Ω, let (E, VE , LE ) be a finite element asso-
ciated with any E ∈ T , and let ΠE the canonical projection upon VE . As already mentioned in
Sect. 4.5.3, we can glue together the local projections to define a global projection ΠT over the
space VT defined in (4.12), by setting
Assuming that both v and ΠT v belong to the space H ` (Ω) for some ` ≥ 0, we want to produce
a measure of the approximation error v − ΠT v, for instance by estimating the semi-norm
|v − ΠT v|`,Ω .
using the additivity property of the integral with respect to the domain of integration. We will
accomplish this task under the (common) assumption that each finite element is the affine image
of a reference element (Ê, V̂ , L̂), as defined in Sect. 4.3.1. Then, our strategy will consist of
three steps:
1. transform the error on the generic element E into an error on the reference element Ê;
3. transform back the result from the reference element to the generic element E.
We start by performing Step 2; next, we discuss how semi-norms transform when going from
a generic element to the reference element and back (useful in Steps 1 and 3); then, we combine
the ingredients by getting estimates on the local errors, and finally on the global error.
57
58 CHAPTER 5. FINITE ELEMENT APPROXIMATION
Assumption 5.1.1 The space V̂ contains a subspace of polynomial functions. Precisely, there
exists an integer k ≥ 0 such that
Pk (Ê) ⊆ V̂ .
Assumption 5.1.2 Functions in V̂ have a certain Sobolev regularity. Precisely, there exists an
integer m ≥ 0 such that
V̂ ⊂ H m (Ê) .
Assumption 5.1.3 The degrees of freedom in L̂ are defined and continuous in a Sobolev space.
Precisely, there exists an integer q ≥ 0 such that
H q (Ê) ⊆ V̂
and there exists a constant Ĉ1 > 0 such that for all `ˆ ∈ L̂ it holds
ˆ
|`(v̂)| ≤ Ĉ1 kv̂kq,Ê ∀v̂ ∈ H q (Ê) .
Examples 5.1.4 For the Courant element, one has k = 1, m is arbitrary since any polynomial
is a C ∞ function, whereas q = 2 since the degrees of freedom are well defined for functions in
C 0 (Ê) and in dimension d = 2 it holds H 2 (Ê) ⊂ C 0 (Ê) with continuous injection.
For the Hsieh-Clough-Tocher element, one has k = 3, m = 2 since functions in V̂ are made of
three pieces of polynomials which are glued in a C 1 manner, whereas q = 3 since the degrees of
freedom are well defined for functions in C 1 (Ê) and in dimension d = 2 it holds H 3 (Ê) ⊂ C 1 (Ê)
with continuous injection.
For the Argyris element, one has k = 5, m is arbitrary since any polynomial is a C ∞ function,
whereas q = 4 since the degrees of freedom are well defined for functions in C 2 (Ê) and in
dimension d = 2 it holds H 4 (Ê) ⊂ C 2 (Ê) with continuous injection.
Under the previous assumptions, let us estimate the projection error v̂ − Π̂v̂. To this end,
recalling that Π̂v̂ = N̂ ˆ
P
j=1 `j (v̂)ϕ̂j and using Assumption 5.1.3, we have
N̂
X XN̂
kΠ̂v̂km,Ê ≤ |`ˆj (v̂)| kϕ̂j km,Ê ≤ Ĉ1 kϕ̂j k m,Ê
kv̂k
q,Ê =: Ĉ2 kv̂kq,Ê ∀v̂ ∈ H q (Ê) .
j=1 j=1
This shows that Π̂ is a linear continuous operator between the spaces H q (Ê) and H m (Ê), i.e.,
Π̂ ∈ L(H q (Ê), H m (Ê)).
On the other hand, the identity mapping I satisfies I ∈ L(H q (Ê), H m (Ê)) if and only if
kv̂km,Ê ≤ Ckv̂kq,Ê for all v̂ ∈ H q (Ê), which is possible if and only if m ≤ q. This motivates the
following assumption.
m ≤ q.
5.1. ERROR BOUNDS ON THE REFERENCE ELEMENT 59
Note that for the Courant element or the Argyris element, and in general for all finite elements
in which V̂ is a space of polynomials, the fulfillment of this assumption suggests the definition
of m, i.e., one chooses m := q (hence, m = 2 for the Courant element or m = 4 for the Argyris
element). On the other hand, for the Hsieh-Clough-Tocher element, we have already observed
that m = 2 while q = 3, which shows that the assumption is fulfilled in this case.
Under the previous assumption, we have I − Π̂ ∈ L(H q (Ê), H m (Ê)), which means that there
exists a constant Ĉ3 > 0 such that
q ≤ k + 1.
Note that this assumption is fulfilled for the three elements considered in the Examples above.
Under the previous assumption, recalling inequality (5.2) and the definition (1.6) of Sobolev
norm which implies the property kv̂kq,Ê ≤ kv̂kk+1,Ê , one has for 0 ≤ ` ≤ m
At this point, we recall the property that Π̂ is a projection operator, i.e., one has Π̂v̂ = v̂
for all v̂ ∈ V̂ . In particular, recalling Assumption 5.1.1, one has Π̂p̂ = p̂ for all polynomials
p̂ ∈ Pk (Ê). This property can be exploited to write the identity
and to apply the bound (5.3) to the function v̂ − p̂, getting the bound
or, equivalently,
|v̂ − Π̂v̂|`,Ê ≤ Ĉ3 min kv̂ − p̂kk+1,Ê . (5.4)
p̂∈Pk (Ê)
Note that the quantity σk+1 (v̂) := minp̂∈Pk (Ê) kv̂ − p̂kk+1,Ê is equal to 0 if and only if v̂ = p̂
for some p̂ ∈ Pk (Ê), i.e., if and only if v̂ is a polynomial of degree ≤ k. There is another
quantity that has the same property, namely the semi-norm |v̂|k+1,Ê : indeed, it is easily seen
(by considering e.g. the Taylor expansion of v̂) that all the derivatives of v̂ of order k + 1 are
identically 0 if and only if v̂ is a polynomial of degree ≤ k. Thus, it is natural to ask ourselves
whether there is some relation between the two quantities. The following result, known as
Denis-Lions Lemma, answers this question showing that the two quantities are equivalent.
Lemma 5.1.7 There exists a constant ĈDL > 0 (depending only on k) such that
On the contrary, the second inequality is a deep result, which can be proven by subtle arguments
in Functional Analysis.
Using this inequality in (5.3), we obtain the existence of a constant Ĉ (= Ĉ3 ĈDL ) such that
Finally, we extend this inequality by observing that if we just assume v̂ ∈ H r (Ê) for some
integer r satisfying q ≤ r ≤ k + 1, then one has the inclusion
hence, all the deductions above can be repeated with k replaced by r − 1, yielding the inequality
Proposition 5.1.8 Under Assumptions (5.1.1)-(5.1.3) and (5.1.5)-(5.1.6) (which imply the in-
equalities m ≤ q ≤ k + 1), let the integers ` and r satisfy the inequalities
0≤`≤m and q ≤ r ≤ k + 1.
Then, there exists a constant Ĉ > 0 (depending on m and k) such that the following inequality
holds:
|v̂ − Π̂v̂|`,Ê ≤ Ĉ|v̂|r,Ê ∀v̂ ∈ H r (Ê) . (5.6)
An equivalent condition for polygonal/polyhedral elements is that the minimum angle of each
element does not degenerate to zero, when varying the element in any trangulation of the family.
We make the following assumption.
Assumption 5.2.2 All the triangulations T considered in the sequel belong to a family F of
regular triangulations of Ω.
We are going to use this assumption in the next properties. Herafter, we denote by c1 , c2 , . . .
positive constants independent of the element E.
Property 5.2.3 The determinants of the Jacobian matrices JF = B and JF−1 = B−1 satisfy
the inequalities
c1 hdE ≤ |det B| ≤ c2 hdE ∀E ∈ T , (5.9)
and
c3 h−d −1
E ≤ |det B | = |det B|
−1
≤ c4 h−d
E ∀E ∈ T . (5.10)
Property 5.2.4 The Euclidean norms of the Jacobian matrices JF = B and JF−1 = B−1
satisfy the inequalities
kBk ≤ c5 hE ∀E ∈ T , (5.12)
and
kB−1 k ≤ c6 h−1
E ∀E ∈ T . (5.13)
If b̂ = bÊ denotes the diameter of the largest ball contained in the reference element Ê, for any z
with kzk = 1 there exist two points x̂1 and x̂2 on the boundary of this ball, hence in Ê, such that
62 CHAPTER 5. FINITE ELEMENT APPROXIMATION
x̂1 − x̂2 = b̂ z. Their images x1 = F(x̂1 ) and x2 = F(x̂2 ) belong to E, hence, kx1 − x2 k ≤ hE .
On the other hand,
hE
showing that kBzk ≤ . Since z is arbitrary, we arrive at (5.12). Inequality (5.13) can be
b̂
obtained similarly, by reversing the roles of E and Ê and applying again Assumption 5.2.2.
We are ready to establish the main result of this section, namely we are going to learn how to
bound a semi-norm of a function by means of the corresponding semi-norm of the transformed
function under the mapping v ↔ v̂.
d/2−`
|v|`,E ≤ C[`] hE |v̂|`,Ê ∀v ∈ H ` (E) , (5.14)
and
`−d/2
|v̂|`,Ê ≤ C [`] hE |v|`,E ∀v̂ ∈ H ` (Ê) . (5.15)
Proof. Let us begin by proving (5.15) for ` = 0; recalling (5.11) and (5.10), we have
Z Z
2
|v̂|0,Ê = 2
v̂ = |det B|−1
v 2 ≤ c4 h−d 2
E |v|0,E ,
Ê E
whence
ˆ ≤ kBT k k∇vk = kBk k∇vk .
k∇v̂k
Thus, again by (5.11), (5.10) and now (5.12), we get
Z Z
ˆ −1
2
|v̂|1,Ê = 2
k∇v̂k ≤ |det B| kBk 2
k∇vk2 ≤ c4 c25 h2−d 2
E |v|1,E ,
Ê E
whence
1/2
X ∂v 2
∂v̂ −1
X ∂v X X
|bji |2 )1/2 |bji |2 )1/2 |v|`,E .
∂ x̂i ◦ F ≤ |bji | ≤(
=(
`−1,E ∂xj
`−1,E
∂xj
`−1,E
j j j j
Thus, using the property |bji | ≤ kBk for all i, j, and (5.2.4), we obtain
X ∂v̂ 2 X
−1
|bji |2 |v|2`,E ≤ d2 kBk|v|2`,E ≤ c5 d2 h2E |v|2`,E .
∂ x̂i ◦ F ≤
i `−1,E i,j
This together with (5.18) yields the desired result. The proof of (5.14) follows the same path,
with the roles of E and Ê exchanged.
Proposition 5.3.1 Under Assumptions 5.1.1–5.1.3, 5.1.5–5.1.6 and 5.2.2, let the integers `
and r satisfy the inequalities
0 ≤ ` ≤ m ≤ q ≤ r ≤ k + 1.
Then, there exists a constant C`,r,k > 0 independent of the element E ∈ T such that the following
inequality holds:
|v − ΠE v|`,E ≤ C`,r,k hr−`
E |v|r,E ∀v ∈ H r (E) . (5.19)
Proof. Recalling the identity (4.10), one has (v − ΠE v)ˆ = v̂ − Π̂ v̂. Then, concatenating
(5.14), (5.6) and (5.15), one obtains
d/2−`
|v − ΠE v|`,E ≤ C[`] hE |v̂ − Π̂v̂|`,Ê
d/2−`
≤ Ĉ C[`] hE |v̂|r,Ê
d/2−` r−d/2
≤ C [r] Ĉ C[`] hE hE |v|r,E ,
Examples 5.3.2 For the Courant element (T, P1 (T ), LT ), one has k = 1 and m = q = k+1 = 2.
Thus, only the choice r = 2 is possible, which gives the estimates
Note that the choice ` = 2 yields a trivial inequality, since the second derivatives of ΠT v are
identically zero.
64 CHAPTER 5. FINITE ELEMENT APPROXIMATION
kv − ΠT vk0,T ≤ Ch3T |v|3,T , |v − ΠT v|1,T ≤ Ch2T |v|3,T , |v − ΠT v|2,T ≤ ChT |v|3,T , (5.21)
kv−ΠT vk0,T ≤ Ch4T |v|4,T , |v−ΠT v|1,T ≤ Ch3T |v|4,T , |v−ΠT v|2,T ≤ Ch2T |v|4,T . (5.22)
Finally, for the Argyris element, one has k = 5, m = q = 4 and k = 5. In the case of maximal
regularity r = 6, one has the estimates, valid for all v ∈ H 6 (T ),
kv−ΠT vk0,T ≤ Ch6T |v|6,T , |v−ΠT v|1,T ≤ Ch5T |v|6,T , |v−ΠT v|2,T ≤ Ch4T |v|6,T . (5.23)
ΠT v ∈ H m̄ (Ω) ∀v ∈ H q (Ω) .
Proposition 5.4.1 Under the same assumptions as in Proposition 5.3.1, the following inequal-
ities hold for any ` and r satisfying 0 ≤ ` ≤ m̄ ≤ q ≤ r ≤ k + 1:
where h is the meshsize defined in (4.1), and the constants C`,r,k are the same as those in (5.19).
Proof. Using (5.19) to estimate the local errors, and recalling that h = maxE∈T hE , we have
X
|v − ΠT v|2`,Ω = |(v − ΠT v)|E |2`,E
E∈T
X
= |v|E − ΠE v|E |2`,E
E∈T
2(r−`)
X
2
≤ C`,r,k hE |v|E |2r,E
E∈T
X
2
≤ C`,r,k h2(r−`) |v|E |2r,E
E∈T
2 2(r−`)
≤ C`,r,k h |v|2r,Ω ,
which highlights that the squared global error is a sum of contributions coming from each
element, where the local measure of smoothness of v is multiplied by a power of the local
meshsize. This ideally suggests that, in order to have balanced contributions to the error, one
should take smaller meshsizes where the derivatives of v are larger, and allow larger meshsizes
where the derivatives of v are smaller.
Since h ≤ diam Ω for P` any 2possible triangulation of Ω, the quantity in parenthesis can be
2
bounded by C̄`,r,k =: j=0 Cj,r,k (diam Ω)2(`−j) , whence the result.
Examples 5.4.4 For the Courant element, one has m̄ = 1, and the global estimates
kv − ΠT vk0,Ω ≤ Ch2 |v|2,Ω , kv − ΠT vk1,Ω ≤ Ch|v|2,Ω , ∀v ∈ H 2 (Ω) . (5.26)
For the quadratic Lagrangian element, one has again m̄ = 1 and, for the maximal regularity,
the global estimates
kv − ΠT vk0,Ω ≤ Ch3 |v|3,Ω , kv − ΠT vk1,Ω ≤ Ch2 |v|3,Ω , ∀v ∈ H 3 (Ω) . (5.27)
For the Hsieh-Clough-Tocher element, one has m̄ = 1 and, for the maximal regularity, the
global estimates
kv − ΠT vk0,Ω ≤ Ch4 |v|4,Ω , kv − ΠT vk2,Ω ≤ Ch2 |v|4,Ω , ∀v ∈ H 4 (Ω) . (5.28)
Finally, for the Argyris element, one has again m̄ = 1 and, for the maximal regularity, the
global estimates
kv − ΠT vk0,Ω ≤ Ch6 |v|6,Ω , kv − ΠT vk2,Ω ≤ Ch4 |v|6,Ω , ∀v ∈ H 6 (Ω) . (5.29)
66 CHAPTER 5. FINITE ELEMENT APPROXIMATION
Chapter 6
In this chapter, we apply the approximation results established in Chapter 5 in order to estimate
the discretization errors produced by a Galerkin method.
We will also provide inverse estimates, which prove useful in several applications.
where V(T ) is defined in (4.12); note that functions in Vh are globally continuous and inherit
the boundary conditions possibly prescribed by the definition of V . Once Vh is defined, we can
introduce the Galerkin solution of the discrete variational problem [VP ]δ given in (2.1) for δ = h,
namely,
uh ∈ Vh : a(uh , vh ) = F (vh ) ∀vh ∈ Vh . (6.3)
The error ku − uh k1,Ω can be estimated by Céa’s Lemma (Theorem 2.4.2). To this end, we
assume that the chosen finite element satisfies Assumptions 5.1.1–5.1.3 and 5.1.5–5.1.6, and we
consider the projection operator Πh := ΠT defined in (5.1), for which the approximation error
is estimated in (5.25). Note that
Πh v ∈ V(T ) ∩ H 1 (Ω)
67
68 CHAPTER 6. FEM DISCRETIZATIONS OF ELLIPTIC PROBLEMS
if v ∈ H r (Ω) with r > 1, since our reference finite element is of class C 0 . We further assume
that Πh v satisfies the same boundary conditions as v, namely it vanishes where v vanishes; this
condition, always satisfied in practice, means that
Πh : H r (Ω) ∩ V → Vh .
Hence, if u ∈ H r (Ω) we can pick vh = Πh u in (2.15) and, using the bound (5.25) with ` = 1, we
get
ku − uh k1,Ω ≤ Ca ku − Πh uk1,Ω ≤ Ca C̄1,r,k hr−1 |u|r,Ω .
Summarizing, we have obtained the following important result.
Proposition 6.1.1 Let u be the solution of (6.1) above, with V ⊆ H 1 (Ω). Let uh ∈ Vh be the
solution of the Galerkin problem (6.3). If the chosen finite element is of class C 0 and satisfies
the other assumptions stated above, there exists a constant C > 0 such that if u ∈ H r (Ω) for
some r satisfying q ≤ r ≤ k + 1, one has the error bound
ku − uh k1,Ω ≤ C hr−1 |u|r,Ω . (6.4)
In particular, in the case of maximal regularity u ∈ H k+1 (Ω), one has
ku − uh k1,Ω ≤ C hk |u|k+1,Ω . (6.5)
For instance, if we use the Courant element (linear Lagrangian element), we have a linear
decay of the error:
ku − uh k1,Ω ≤ C h|u|2,Ω if u ∈ H 2 (Ω) , (6.6)
whereas if we use the quadratic Lagrangian element (T, P2 (T ), LT ), we have a quadratic decay:
ku − uh k1,Ω ≤ C h2 |u|3,Ω if u ∈ H 3 (Ω) . (6.7)
Remark 6.1.2 Going back to the general Theorem 2.4.3, its assumptions are satisfied with the
choice Z = H r (Ω) ∩ V , which is dense in V since H r (Ω) is dense in H 1 (Ω). The function ψZ in
the theorem is ψZ (h) = C̄1,r,k hr−1 from (5.25), and the abstract bound (2.23) is precisely (6.4).
Furthermore, the theorem assures that the Galerkin solution uh converges to u for any solution
u ∈ V of the variational problem [VP ].
Let us now suppose that the variational problem (6.1) corresponds to a fourth-order elliptic
problem, as the Timoshenko plate bending problem considered in Sect. ....; in this case, V ⊆
H 2 (Ω) with kvkV = kvk2,Ω . If we use a reference element of class C 1 , then the operator Πh = ΠT
satisfies
Πh ∈ V(T ) ∩ H 2 (Ω)
if v ∈ H r (Ω) with r > 2. Therefore, the previous considerations apply with the obvious changes,
and one gets the error bound
ku − uh k2,Ω ≤ C hr−2 |u|r,Ω , (6.8)
which holds for any u ∈ H r (Ω) with 2 < q ≤ r ≤ k + 1. In particular, in the case of maximal
regularity u ∈ H k+1 (Ω), one gets
ku − uh k2,Ω ≤ C h2 |u|4,Ω
for the Hsieh-Clough-Tocher element, and
ku − uh k2,Ω ≤ C h4 |u|6,Ω
for the Argyris element.
6.2. ERROR ESTIMATES IN THE L2 -NORM 69
where the Cauchy-Schwarz inequality has been used. Identity (6.9) means that the L2 -norm
of the function v coincides with the norm of the linear continuous form g 7→ (v, g)Ω on L2 (Ω),
i.e., the norm of this form in the dual space L2 (Ω)0 ; this is coherent with the identification
L2 (Ω)0 = L2 (Ω) via the Riesz Representation Theorem in Hilbert spaces.
We apply (6.9) to v = u − uh , getting
(u − uh , g)Ω
ku − uh k0,Ω = sup . (6.10)
g∈L2 (Ω), g6=0 kgk0,Ω
For any g ∈ L2 (Ω), we solve the auxiliary problem: find w ∈ V such that
This problem is uniquely solvable, since the adjoint bilinear form aT , defined by aT (w, v) =
a(v, w) for all w, v ∈ V , fulfills the same assumptions of the Lax-Milgram theorem as the form
a does. Let us denote by wg the solution of problem (6.11). Then, we have
where we have used the Galerkin orthogonality property (2.12). Substituting in (6.10), we obtain
distH 1 (Ω) (wg , Vh )
ku − uh k0,Ω ≤ kak ku − uh k1,Ω sup . (6.12)
g∈L2 (Ω), g6=0 kgk0,Ω
We now give conditions which guarantee that the sup quantity is O(h). To this end, let us
assume that
g ∈ L2 (Ω) =⇒ wg ∈ H 2 (Ω) , with kwg k2,Ω ≤ Creg kgk0,Ω (6.13)
70 CHAPTER 6. FEM DISCRETIZATIONS OF ELLIPTIC PROBLEMS
for some Creg > 0. This is a regularity property of the adjoint problem (6.11), which holds, e.g,
if the coefficients of the elliptic operator are smooth and the boundary ∂Ω is smooth (or the
domain Ω is convex). Furthermore, let us assume that
distH 1 (Ω) (wg , Vh ) ≤ kwg − Πh wg k1,Ω ≤ C̄1,2,k h|v|2,Ω ≤ C̄1,2,k Creg h kgk0,Ω ,
Proposition 6.2.1 Under the assumptions of Proposition 6.1.1 and the assumptions (6.13)–
(6.14) above, there exists a constant C > 0 such that if u ∈ H r (Ω) with 2 ≤ r ≤ k + 1, one has
the error bound
ku − uh k0,Ω ≤ C hr |u|r,Ω . (6.15)
The result we have obtained means that the Galerkin discretization error u − uh behaves
asymptotically (for h → 0) as the projection error u − Πh u not only in the energy norm, but
also in the norm of L2 (Ω).
For instance, with the Courant element we have a quadratic decay of the error:
To this end, we require regularity of the exact solution in Sobolev spaces based on L∞ (Ω),
namely for r ≥ 0 we consider the spaces
Denoting as usual by k the polynomial degree of the chosen finite element, one can prove the
following error estimate:
h2 |u|2,∞,Ω if k = 1, d = 1 ,
ku − uh kL∞ (Ω) ≤ C h2 | log h| |u|2,∞,Ω if k = 1, d ≥ 2 , (6.18)
k+1
h |u|k+1,∞,Ω if k ≥ 2, d ≥ 1 ,
kykX ≤ C kykY ∀y ∈ Y .
i.e., any norm can be bounded by a constant times any other norm. This statement (which is
not true in infinite dimension!) is a simple consequence of Weierstrass’ theorem for continuous
functions in Euclidean spaces. Thus, there exists a constant CZ > 0 depending on Z such that
This is an instance of inverse inequality, which expresses the possibility of controlling a stronger
norm (the Y -norm) by a weaker norm (the X-norm), in finite dimension. In general, the constant
CZ depends on the dimension of Z, and blows-up when this dimension tends to infinity. In
applications, one is interested in estimating the size of the constant CZ , highlighting such a
dependence.
Let us discuss the following important example. Let X = L2 (Ω) and Y = H 1 (Ω). Consider
the finite-element subspace
Vh = V(T ) ∩ H 1 (Ω)
introduced in (6.2), under the same assumptions. We want to find a constant CVh satisfying
The idea is to go from a generic element to the reference element and back, using the equivalence
of norms on the reference element as well as the semi-norm transformations presented in Sect.
5.2. Hereafter, we will set vE = vh|E and we denote by v̂E ∈ V̂ the function on the reference
element associated to vE by the transformation ↔ introduced in Sect. 4.3.1. Using (5.14), one
gets X X X
|vh |21,Ω = |vE |21,E ≤ C[1]
2
hd−2
E |v̂ E |2
1,Ê
≤ C 2
[1] hd−2 2
E kv̂E k1,Ê .
E∈T E∈T E∈T
Since all norms are equivalent on the finite dimensional space V̂ , there exists a constant Ĉ such
that kv̂k1,Ê ≤ Ĉkv̂k0,Ê for all v̂ ∈ V̂ . Hence, using now (5.15), we have
X X
−d
|vh |21,Ω ≤ C[1]
2
Ĉ 2 hd−2 2 2 2 [0] 2
E kv̂E k0,Ê ≤ C[1] Ĉ (C ) hd−2 2
E hE kvE k0,E .
E∈T E∈T
Now, h−1 −1
E can be bounded by hmin , where hmin := minE∈T hE (recall that h = hT = maxE∈T hE ).
Hence,
X
|vh |21,Ω ≤ C[1]
2
Ĉ 2 (C [0] )2 h−2
min kvE k20,E = C[1]
2
Ĉ 2 (C [0] )2 h−2 2
min kvh k0,Ω .
E∈T
Adding kvh k20,Ω to both sides, we obtain the following inverse inequality.
Proposition 6.4.1 There exists a constant C > 0 independent of the triangulation such that
We can make the meshsize h, rather than hmin , appear in the inequality, if we assume that
hmin is comparable to h. This is expressed by the following definition.
Definition 6.4.2 A family F = {T } of admissible triangulations of Ω is said to be a family of
quasi-uniform triangulations if there exists a constant τ ≥ 1 such that
h
≤τ ∀T ∈ F . (6.21)
hmin
1
In a quasi-uniform triangulation, one has hmin ≥ τ h, which rules out the possibility of having
elements in T exceedingly smaller than others.
Corollary 6.4.3 If T belongs to a quasi-uniform family of triangulations, one can find a con-
stant C > 0 independent of the triangulation such that
The inverse inequalities we have just established are sharp, in the sense that they would not
hold with an exponent of h larger than −1. Indeed, consider a uniform mesh on which the
Courant elements are used. If ϕj denotes as usual the Lagrange basis function associated with
a node xj , it easily checked that
Finally, we mention that inverse inequalities in other norms can be obtained with the same
technique described above. Here is an example, which extends (6.22).
6.4. INVERSE INEQUALITIES 73
Proposition 6.4.4 Assume that the chosen reference finite element is of class C 1 , and let Vh =
V(T ) ∩ H 2 (Ω). Then, there exists a constant C > 0 independent of the triangulation such that
We close this section by presenting two applications of the inverse inequalities, where we
provide estimates of the eigenvalue of certain matrices.
introduced in (2.7), appears in the linear system that expresses a Galerkin method in algebraic
terms. We aim at deriving an upper bound for the condition number cond2 (A) of A in the
Euclidean norm.
To this end, we need an auxiliary result. Recall that each function vh ∈ Vh is uniquely
identified by its degrees of freedom, which are the values vj = vh (xj ) at the nodes xj , since
vh = N Nh and let kvk denote its Euclidean norm.
P h
j=1 vj ϕj ; let v = (vj ) ∈ R
Property 6.4.5 There exist constants c1 , c2 > 0 independent of the triangulation T , such that
with
(C [0] )−2 hdE kv̂E k20,E ≤ kvE k20,E ≤ C[0]
2
hdE kv̂E k20,E
by Proposition 5.2.5 with ` = 0. Since the triangulation is quasi-uniform by assumption, we
have τ −1 h ≤ hE ≤ h for any E ∈ T (recall (6.21)). Hence,
Now observe that v̂E = vE , since by definition of image of the reference finite element, the
values of the degrees of freedom of a function v ∈ VE are the same as the values of the degrees
of freedom of the image v̂ ∈ V̂ (recall (4.7)). Furthermore, observe that kv̂k is a norm in V̂ ,
74 CHAPTER 6. FEM DISCRETIZATIONS OF ELLIPTIC PROBLEMS
We conclude by observing that by the regularity of the triangulation, there exists a constant
C? > 0 independent of the triangulation, such that
X
kvk2 ≤ kvE k2 ≤ C? kvk2 ;
E∈T
indeed, each entry of v (say, vh (xj ) for some j) is also an entry of one or more vectors vE (those
corresponding to the elements E containing xj ), but the number of such vectors is uniformly
bounded, since the number of elements which have non-empty intersections with a given element
is bounded by the constant γ appearing in (5.8).
Aw = µ w .
Multiplying by wT , we obtain
wT Aw = µ wT w .
PNh
We have already observed that wT Aw = a(wh , wh ), where wh = j=1 wj ϕj (see (2.9)). On the
other hand, wT w = kwk2 ; hence, µ satisfies
a(wh , wh ) = µ kwk2 .
Let us denote by α and kak the coercivity and continuity constants of the form a. By Property
6.4.5, we have
simplifying the common factor kwh k20,Ω at the endpoints, we obtain the lower bound
α c21 hd ≤ µ . (6.26)
µ kwk2 = a(wh , wh ) ≤ kak kwh k21,Ω ≤ kak C 2 h−2 kwh k20,Ω ≤ kak C 2 h−2 c22 hd kwk2 ,
6.4. INVERSE INEQUALITIES 75
Proposition 6.4.6 Under the previous assumptions, the condition number of the stiffness ma-
trix A satisfies the bound
cond2 (A) ≤ C h−2 , (6.28)
Using additional arguments, it is possible to prove that the bound is sharp, namely, cond2 (A)
does grow proportionally to h−2 as h → 0. This shows that the stiffness matrix associated with
a Lagrangian basis becomes more and more ill-conditioned as the mesh is refined.
Remark 6.4.7 If the stiffness matrix A corresponds to the discretization of a fourth order
problem, i.e., if the bilinear form a is coercive and continuous with respect to the H 2 (Ω) norm
and the Hsieh-Clough-Tocher element or the Argyris element is used to generate the subspace
Vh , then (6.28) is replaced by
cond2 (A) ≤ C h−4 . (6.29)
Let H and V separable Hilbert spaces, such that V ⊂ H with compact injection (i.e., if {vi }
is a sequence bounded in V , one can extract a subsequence {vik } which converges to some v in
H). This occurs, e.g., if H = L2 (Ω) and V = H01 (Ω).
Let a : V × V → R be a symmetric, continuous and coercive bilinear form in V ; similarly, let
b : H × H → R be a symmetric, continuous and coercive bilinear form in H.
Consider the spectral problem:
(
F ind λ ∈ R and w ∈ V, w 6= 0, such that
(6.30)
a(w, v) = λ b(w, v) ∀v ∈ V .
One can prove by the Spectral Theorem for compact self-adjoint operators, that the problem
admits a non-decreasing sequence of strictly positive eigenvalues λn and corresponding eigen-
functions wn , satisfying
• a(wn , wm ) = b(wn , wm ) = 0 if n 6= m.
76 CHAPTER 6. FEM DISCRETIZATIONS OF ELLIPTIC PROBLEMS
Example 6.4.8 An important example is the spectral problem for the Laplace operator with
Dirichlet boundary conditions: (
−∆w = λw in Ω ,
(6.31)
w=0 on ∂Ω ,
which describes the free vibrations of a homogeneous, elastic thin membrane. The problem can
be formulated variationally as
Problem (6.30) can be discretized by the Galerkin method. Let Vh a finite dimensional
subspace of V , whose dimension is Nh . Consider the problem
(
F ind λh ∈ R and wh ∈ Vh , wh 6= 0, such that
(6.32)
a(wh , vh ) = λh b(wh , vh ) ∀vh ∈ Vh .
Introducing a basis in Vh and denoting by A the stiffness matrix associated with the bilinear
form a, and by B the stiffness matrix associated with the bilinear form b, the discrete spectral
problem can be formulated at the algebraic level as
Aw = λh Bw , (6.33)
where w is the vector collecting the coefficients of the representation of wh in the chosen basis.
This is a generalized eigenvalue problem, which in Matlab can be solved by the command
eig(A, B).
The discrete eigenvalues and eigenfunctions satisfy properties similar to those of the exct
problem, namely:
We are interested in deriving lower and upper bounds for the discrete eigenvalues.To this end,
let us assume that the following inverse inequality holds:
This is precisely (6.22) if H is endowed by the L2 (Ω)-norm and V is endowed by the H 1 (Ω)-
norm. Let α, kak denote the coercivity and continuity constants of the form a with respect to
the norm in V ; similarly, let β, kbk denote the coercivity and continuity constants of the form b
with respect to the norm in H. Furthermore, let c > 0 the constant of the continuity inequality
kvkH ≤ c kvkV for all v ∈ V .
Picking vh = wh in (6.32), we get
a(wh , wh ) = λh b(wh , wh ) .
which – after cancellation of the term kwh k2H 6= 0 at the endpoints, yield
α
≤ λh .
c2 kbk
λh βkwh k2H ≤ λh b(wh , wh ) = a(wh , wh ) ≤ kak kwh k2V ≤ kak C h−2 kwh k2H ,
Proposition 6.4.9 If the inverse inequality (6.34) holds, then any eigenvalue of the discrete
spectral problem (6.32) satisfies the bounds
α C kak −2
≤ λh ≤ h . (6.35)
c2 kbk β
For this application, too, a more sophisticated analysis shows that the two bounds are indeed
sharp. Precisely, there exists constants c? , C? > 0 independent of h such that
Remark 6.4.10 One can prove that each eigenvalue λh,n converges to the corresponding exact
eigenvalue λn , at a rate proportional to the square of the distance of the eigenfunction wn from
the subspace Vh , i.e.,
|λn − λh,n | ≤ C dist2V (wn , Vh ) .
For instance, if the Courant element is used and the eigenfunctions belong to H 2 (Ω) (which
occurs e.g. if Ω is convex), then the convergence of the eigenvalues is quadratic in h. However,
the constant depends on n, and gets larger and larger as n increases; in other words, on the
same mesh smaller eigenvalues are approximated better than larger eigenvalues.
78 CHAPTER 6. FEM DISCRETIZATIONS OF ELLIPTIC PROBLEMS
Chapter 7
Discretizations
of Parabolic Problems
7.1
79