Sobolev spaces and applications by IITK
Sobolev spaces and applications by IITK
T. Muthukumar
tmk@iitk.ac.in
Notations vii
1 Theory of Distributions 1
1.1 History . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.1 Case 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2.2 Case 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
1.3 Space of Test Functions . . . . . . . . . . . . . . . . . . . . . 4
1.3.1 Smooth Non-Analytic (Bump) Functions . . . . . . . . 6
1.3.2 Mollifiers . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.3.3 Topology on C(Ω), C0 (Ω) and Cc (Ω) . . . . . . . . . . 12
1.3.4 Algebraic and Topological Dual of Cc (Ω) . . . . . . . . 13
1.3.5 Topology on C ∞ (Ω) . . . . . . . . . . . . . . . . . . . 15
1.3.6 Inductive Limit Topology on Cc∞ (Ω) . . . . . . . . . . 17
1.3.7 Regularization and Cut-off Technique . . . . . . . . . . 19
1.4 Space of Distributions . . . . . . . . . . . . . . . . . . . . . . 25
1.4.1 Functions as Distributions . . . . . . . . . . . . . . . . 28
1.4.2 Measures as Distributions . . . . . . . . . . . . . . . . 31
1.4.3 Multipole Distributions . . . . . . . . . . . . . . . . . . 35
1.4.4 Infinite Order Distributions . . . . . . . . . . . . . . . 38
1.4.5 Topology on Distributions . . . . . . . . . . . . . . . . 38
1.4.6 Principal Value Distribution . . . . . . . . . . . . . . . 42
1.4.7 Functions, but not Distributions . . . . . . . . . . . . . 47
1.5 Operations With Distributions . . . . . . . . . . . . . . . . . . 47
1.5.1 Differentiation . . . . . . . . . . . . . . . . . . . . . . . 48
1.5.2 Product . . . . . . . . . . . . . . . . . . . . . . . . . . 59
1.5.3 Support . . . . . . . . . . . . . . . . . . . . . . . . . . 64
1.5.4 Singular Support . . . . . . . . . . . . . . . . . . . . . 67
iii
CONTENTS iv
2 Sobolev Spaces 77
2.1 Hölder Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
2.2 W k,p (Ω) Spaces . . . . . . . . . . . . . . . . . . . . . . . . . . 85
2.3 Smooth Approximations . . . . . . . . . . . . . . . . . . . . . 90
2.4 Characterisation of W0k,p (Ω) . . . . . . . . . . . . . . . . . . . 95
2.5 Extension Operators . . . . . . . . . . . . . . . . . . . . . . . 96
2.6 Topological Dual of Sobolev Spaces . . . . . . . . . . . . . . . 101
2.7 Fractional Order Sobolev Space . . . . . . . . . . . . . . . . . 103
2.8 Imbedding Results . . . . . . . . . . . . . . . . . . . . . . . . 108
2.8.1 Sobolev Inequality (1 ≤ p < n) . . . . . . . . . . . . . 109
2.8.2 Poincaré Inequality . . . . . . . . . . . . . . . . . . . . 116
2.8.3 Equality case, p = n . . . . . . . . . . . . . . . . . . . 118
2.8.4 Morrey’s Inequality (n < p < ∞) . . . . . . . . . . . . 119
2.8.5 Generalised Sobolev Imbedding . . . . . . . . . . . . . 122
2.8.6 Compact Imbedding . . . . . . . . . . . . . . . . . . . 124
2.9 Trace Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . 128
2.10 Further reading . . . . . . . . . . . . . . . . . . . . . . . . . . 142
Appendices 177
Bibliography 185
Index 187
CONTENTS vi
Notations
Symbols
Function Spaces
C k,γ (Ω) all functions in C k (Ω) whose k-th partial derivatives (k ≥ 0) are
Hölder continuous with exponent γ
C ∞ (Ω) denotes class of C ∞ (Ω) functions such that all its derivatives can be
extended continuously to Ω
Cc∞ (Ω) is the class of all infinitely differentiable functions on Ω with compact
support
∞
CK (Ω) is the class of all infinitely differentiable functions on Ω with support
contained in the compact subset K ⊂ Ω
vii
NOTATIONS viii
W k,p (Ω) is the class of all functions in Lp (Ω) whose distributional derivative
upto order k are also in Lp (Ω)
General Conventions
M(α) denotes, for α > 0, the class of all n × n matrices, A = A(x), with
L∞ (Ω) entries such that,
Theory of Distributions
1.1 History
In late 1920’s P. A. M. Dirac (cf. [Dir49]) derived the equation
d 1
ln(x) = − iπδ(x) (1.1.1)
dx x
in the study of quantum theory of collision processes, where δ is a “function”
defined and continuous in real line R satisfying the following properties:
1. δ(x) = 0 for x 6= 0.
R∞
2. −∞ δ(x) dx = 1.
R∞
3. For any continuous function defined on R, f (a) = −∞
f (x)δ(a − x) dx
for all a ∈ R.
1
CHAPTER 1. THEORY OF DISTRIBUTIONS 2
The δ became known as Dirac’s delta function. It had caused unrest among
the mathematicians of the era, because δ did not adhere to the classical
notion of function.
H. Lebesgue had laid the foundations of measure theory in his doctoral
thesis in 1902. With this knowledge, δ can be viewed as a set function on
the σ-algebra(a collection of subsets of R). δ can be viewed as a positive
measure, called Dirac measure, defined on subsets E from the σ-algebra of
R as, (
1 if 0 ∈ E
δ(E) =
0 if 0 ∈/ E.
Thus, mathematically one has given sense to properties in 1, 2 and 3. How-
ever, the differentiability of δ and H was yet to be made precise, since clas-
sically, every differentiable function is continuous.
In 1944, G. Choquet and J. Deny published a work on polyharmonic
functions in two dimensions (cf. [CD44]). Laurent Schwartz, in an attempt
to generalise the work of Choquet and Deny in higher dimensions, published
an article in 1945 (cf. [Sch45]). With this article, the theory of distributions
was discovered and also settled the issue raised in properties 4 and 5, in
addition to 1, 2 and 3.
Dirac won the physics Nobel prize (1933) and Schwartz won the Fields
medal (1950) for their respective work.
1.2 Motivation
The theory of distribution is a concept that generalises the notion of function,
hence is also called generalised functions. But the beauty of the theory lies in
the fact that this new notion admits the concept of differentiation and every
distribution is infinitely differentiable. The need for such a generalisation
was felt at various instances in the history of mathematics.
1.2.1 Case 1
Recall that the wave equation utt (x, t) = c2 uxx (x, t) on R × (0, ∞), describ-
ing the vibration of an infinite string, has the general solution u(x, t) =
F (x + ct) + G(x − ct). In reality, it happens that even if F and G are not
twice differentiable, they are “solution” to the wave equation. For instance,
CHAPTER 1. THEORY OF DISTRIBUTIONS 3
1.2.2 Case 2
Consider the Burger’s equation
ut (x, t) + u(x, t)ux (x, t) = 0 x ∈ R and t ∈ (0, ∞)
u(x, 0) = φ(x) x ∈ R
Note that the characteristic curves are intersecting. This is no issue, as long
as, the solution u which is constant on these curves takes the same constant
in each characteristic curve. Unfortunately, that is not the case here:
1
x<t
x
u(x, t) = t+1 0 ≤ 1 − xt ≤ 1
0 1 ≤ x.
compact support. Suppose f was chosen from C k (a, b), then above integra-
tion by parts could be repeated k times to get
Z b Z b
(k) k
φ(x)f (x) dx = (−1) f (x)φ(k) dx.
a a
Rb Rb
Observe that the maps φ 7→ a f φ dx and φ 7→ a f φ0 dx are linear on
Cc∞ (a, b). If there exists a suitable complete topology on Cc∞ (a, b) such that
these linear maps are continuous, then f can be identified with a continuous
linear functional on Cc∞ (a, b) given by
Z b
φ 7→ f φ dx (1.3.2)
a
Proof. If φ were analytic then, for all a ∈ Ω, we have the Taylor expansion
P∞ φ(k) (a)
φ(x) = k=0 k!
(x − a)k . Analyticity of φ implies that the radius of
convergence of the Taylor series at a is infinite. For φ ∈ Cc∞ (Ω), choose
a 6∈ supp(φ). The Taylor’s expansion around a imply that φ is zero function,
which is a contradiction.
Therefore, no non-zero analytic function can sit in Cc∞ (Ω) because zero
function is the only analytic function with compact support. The function
f (x) = eax is a smooth analytic function on R for all a ∈ R. Do we have
non-zero smooth (non-analytic) functions in Cc∞ (Ω)?
CHAPTER 1. THEORY OF DISTRIBUTIONS 6
Thus,
exp(−1/h)
≤ hm (m + 1)!
h
and the larger the m, the sharper the estimate. Hence,
exp(−1/h)
f 0 (0) = lim+ = 0.
h→0 h
The same argument follows for any k + 1 derivative of f because
converges to the zero function for all x ∈ R. But for x > 0, we know that
f (x) > 0 and hence do not converge to the Taylor series at x = 0. Thus, f
is not analytic.
Example 1.4. The function f : R → R defined as
(
exp(−1/x2 ) if x 6= 0
f3 (x) =
0 if x = 0
Lemma 1.3.1 (Smooth Cut-off Function). For any interval (a, b) of R, there
is a decreasing smooth function g : R → R such that 0 < g(x) < 1 for all
x ∈ (a, b) and (
1 x≤a
g(x) =
0 x ≥ b.
h(x) = f (b − x) + f (x − a)
CHAPTER 1. THEORY OF DISTRIBUTIONS 9
on R which is (
f (b − x) x ≤ a
h(x) =
f (x − a) x ≥ b.
Since f (b − x) < f (b − x) + f (x − a), addition of positive function. Moreover,
h is smooth. Define the function
f (b − x)
g(x) =
f (b − x) + f (x − a)
and it satisfies all the desired properties.
In particular, when f = f2 , we have
e−1/(b−x)
g(x) = .
e−1/(b−x) + e−1/(x−a)
Note that the increasing function
f (x − a)
1 − g(x) =
f (b − x) + f (x − a)
is identically zero for x ≤ a and one on x ≥ b satisfying the other properties
of above lemma.
Lemma 1.3.2 (Smooth Bump Function). For any positive a, b ∈ R such
that a < b, there is a h ∈ Cc∞ (R) such that 0 ≤ h(x) ≤ 1, for all x ∈ R,
h ≡ 1 on [−a, a] and supp(h)= [−b, b].
Proof. Recall the function g obtained in above lemma. Then the function
x 7→ g(|x|) is identically one in [−a, a] and zero outside (−b, b). The desired
function h is obtained by setting h(x) = g(|x|). The function h is smooth
because it is composed with the function |x|, which is smooth except at x = 0.
But h ≡ 1 around the origin.
Lemma 1.3.3 (Smooth Bump Function on Rn ). For any 0 < a < b, there
is a h ∈ Cc∞ (Rn ) with range in [0, 1] such that 0 ≤ h(x) ≤ 1, for all x ∈ R,
h ≡ 1 on the closed ball Ba (0) and supp(h) is the closure of Bb (0).
Proof. The same function h(x) = g(|x|) works.
Theorem 1.3.4. For any compact subset K of Rn , then there exists a φ ∈
Cc∞ (Rn ) such that φ ≡ 1 on K.
CHAPTER 1. THEORY OF DISTRIBUTIONS 10
Proof. Since K is compact, there exists a a > 0 such that Ba (0) ⊃ K and
we apply previous Lemma. Alternately, an existential proof is as follows:
One can chose an open set U ⊃ K such that U is compact. Then, U and
K c form an open cover of Rn . By Theorem ??, there is a partition of unity
subordinate to the given cover of Rn . Thus, there are smooth non-negative
functions φ and ψ on Rn such that supp(φ) is in U , supp(ψ) is in K c and
φ + ψ = 1 on Rn . On K, ψ = 0 and hence φ = 1. Also, supp(φ) is a
closed subset of a compact subset U , hence φ has compact set. Thus, φ is
the desired function.
1.3.2 Mollifiers
The functions f2 and f4 in the above examples can be tweaked to construct
a function in Cc∞ (R) such that supp(f )= [−b, b] for any positive b ∈ R. Note
that this time we do not demand that the function takes constant value 1 in
a subset of the support. Consider the transformation g2b (x) = f2 (1 − |x|/b).
Thus,
(
−b
exp b−|x| if |x| < b
g2b (x) =
0 if |x| ≥ b.
In fact, the functions g2b , g5b can be extended to n dimensions and are in
Cc∞ (Rn ) with support in B(0; b), the disk with centre at origin and radius b.
We shall now introduce an important sequence of functions in Cc∞ (Rn ),
called mollifiers. Recall the Cc∞ (Rn ) functions g2b and g5b introduced in the
CHAPTER 1. THEORY OF DISTRIBUTIONS 11
−ε2
Z Z
ε
g5 (x) dx = exp 2 dx
Rn |x|<ε ε − |x|2
−1
Z
n
= ε exp dy (by setting y = x/ε)
|y|<1 1 − |y|2
= εn c−1 ,
where
−1
Z
−1
c = exp dy.
|y|≤1 1 − |y|2
Now, set ρε (x) = cε−n g5ε (x), equivalently,
( 2
cε−n exp ε2−ε
−|x|2
if |x| < ε
ρε (x) = (1.3.3)
0 if |x| ≥ ε.
Note that ρε ≥ 0 and is in Cc∞ (Rn ) with support in B(0; ε). The sequence
{ρε } is an example of mollifiers, a particular case of the Dirac Sequence.
(iii) For every given r > 0 and ε > 0, there exists a N0 ∈ N such that
Z
ρk (x) dx < ε, ∀k > N0 .
Rn \B(0;r)
The connection between the sequence of mollifiers and Dirac delta func-
tion will become evident in the sequel. The notion of mollifiers is also an
example for the approximation of identity concept in functional analysis and
ring theory.
is complete. Thus C(Ω) is a Banach space w.r.t the uniform norm. For a non-
compact subset Ω of Rn , the space C(Ω) is not normable but are metrizable1 .
They form a locally convex complete metric space called Fréchet space. We
describe this metric briefly in the following paragraph.
For any open subset Ω of Rn , there is a sequence Kj of non-empty compact
subsets of Ω such that Ω = ∪∞ j=0 Kj and Kj ⊂ Int(Kj+1 ), for all j (exhaustion
of an open set by compact sets, cf. Lemma ??). This property is called the
σ-compactness of Ω. We define a countable family of semi-norms on C(Ω) as
pj (φ) = kφk∞,Kj .
form a local base for C(Ω). The metric induced by the family of semi-norms
on C(Ω) is
1 pj (φ − ψ)
d(φ, ψ) = max j
j∈N∪{0} 2 1 + pj (φ − ψ)
and the metric is complete and C(Ω) is a Fréchet space. This is precisely the
topology of compact convergence (uniform convergence on compact sets) or
the compact-open topology in this case. If {φm } is a Cauchy sequence w.r.t
d then pj (φm − φ` ) → 0, for all j and as m, ` tends to infinity. Thus {φm }
converges uniformly on Kj to some φ ∈ C(Ω). Then it is easy to see that
1
Any norm space is a metric space but the converse is not true always because d(x, 0)
may fail to satisfy the properties of norm. A metric d induces a norm if d(x + z, y + z) =
d(x, y) and d(αx, αy) = |α|d(x, y).
CHAPTER 1. THEORY OF DISTRIBUTIONS 13
d(φ, φm ) → 0. The metric defined above is called the Fréchet metric and is
equivalent to the metric
∞
X 1 pj (φ − ψ)
d(φ, ψ) = j 1 + p (φ − ψ)
.
j=0
2 j
The Fréchet space may be seen as a countable limit of Banach spaces. In our
case, the Banach spaces are C(Ki ) w.r.t the uniform norm, the restriction of
the semi-norm.
Exercise 3. Show that the topology given in C(Ω) is independent of the
choice the exhaustion compact sets {Kj } of Ω.
We say a function φ : Ω → R vanishes at infinity (or on the boundary),
if for every ε > 0 there exists a compact set K ⊂ Ω (depending on ε) such
that |φ(x)| < ε for all x ∈ Ω \ K. Let C0 (Ω) be the set of all continuous
functions on Ω vanishing at infinity (or on the boundary). Let Cc (Ω) be the
set of all continuous functions with compact support in Ω. Let Cb (Ω) be
the space of all bounded continuous functions on Ω. We have the inclusion
Cc (Ω) ⊂ C0 (Ω) ⊂ Cb (Ω) ⊂ C(Ω). One may assign the uniform norm on
Cb (Ω), C0 (Ω) and Cc (Ω). Under the uniform norm Cc (Ω) is dense in C0 (Ω)
which is a closed subspace of the Banach space Cb (Ω).
We shall construct a complete (non-metrizable) topology on Cc (Ω). For
every compact subset K ⊂ Ω, let CK (Ω) denote the class of all continuous
functions in Ω such that their support is in K. The space CK (Ω) is a Banach
space w.r.t the uniform norm. Note that
Cc (Ω) = ∪K⊂Ω CK (Ω)
where the union is over all compact subsets of Ω. We declare a map T on
Cc (Ω) is continuous if T restricted to CK (Ω), for each compact K ⊂ Ω, is
continuous. Such a topology is called the inductive limit topology of Cc (Ω)
with uniform norm. We say a sequence {φm } converges to φ w.r.t inductive
limit topology if there exists a compact set K such that supp(φm ) ⊂ K and
φm converges uniformly to φ. The space Cc (Ω) is complete with respect to
the inductive limit topology.
and
µ(K) = inf T (φ) ∀K ⊂ Ω. (1.3.6)
φ∈SK
Note that any linear continuous (bounded) functional Cc (Ω) is also locally
bounded and, hence, the above result is true.
For a locally compact Hausdroff space Ω, recall (cf. § 1.3.3) that the
spaces Cc (Ω) ⊂ C0 (Ω) ⊂ Cb (Ω) can be endowed with the uniform norm
where the first inclusion is dense and C0 (Ω) is closed subspace of Cb (Ω).
Further, one may also endow the inductive limit topology on Cc (Ω) which
completes it.
CHAPTER 1. THEORY OF DISTRIBUTIONS 15
Theorem 1.3.13. Consider Cb (Ω) endowed with the uniform topology. Then
there is an isometric isomorphism between the dual of Cb (Ω) and the space
of bounded finitely additive (not necessarily countably additive) measures,
i.e. for any continuous linear functional T : Cb (Ω) → R there is a unique
bounded, finitely additive measure µ such that
Z
T (φ) = φ(x) dµ ∀φ ∈ Cb (Ω).
Ω
Note that the class of bounded countably additive measures are a sub-
space of the class of bounded finitely additive measures.
Note that due to the closed inclusion of C0 (Ω) in Cb (Ω) the dual space
inclusion are not reversed. In fact they preserve the inclusion.
induces the same topology as the one inherited from C ∞ (Ω). However, this
norm induced topology on Cc∞ (Ω) is not complete and its completion is
C ∞ (Ω). For more details and proof of the semi-norm induced locally convex
topology refer to [Rud91].
Exercise 5. Let φ ∈ Cc∞ (R) with supp(φ)= [0, 1] and φ > 0 in (0, 1). Then
the sequence
m
X 1 1 1
ψm (x) = φ(x − i) = φ(x − 1) + φ(x − 2) + . . . + φ(x − m)
i=1
i 2 m
is Cauchy in the topology induced by the norms, but lim ψm 6∈ Cc∞ (R).
CHAPTER 1. THEORY OF DISTRIBUTIONS 17
∞
Exercise 6. CK (Ω) is a closed subspace of C ∞ (Ω) under the inherited topol-
∞
ogy of C (Ω).
Proof. For each x ∈ Ω, define the functional Tx : C ∞ (Ω) → R defined as
Tx (φ) = φ(x). For each x ∈ Ω, there is a j0 such that x ∈ Kj for all j ≥ j0 .
Then,
|Tx (φ)| = |φ(x)| ≤ pj (φ) ∀j ≥ j0 .
The functional Tx is continuous2 because uniform convergence implies point-
wise convergence. The topology on C ∞ (Ω) is uniform convergence on com-
pact subsets. Therefore the kernel of Tx ,
Cc∞ (Ω) = ∪∞ ∞
j=1 CKj (Ω)
2
Compare the functional Tx with Dirac distribution to be introduced later
CHAPTER 1. THEORY OF DISTRIBUTIONS 18
∞ ∞
and CK `
(Ω) ⊂ CK m
(Ω) for all ` < m. With the given topology on these
spaces, inherited from C ∞ (Ω), the inclusion map Ik` : CK
∞
`
∞
(Ω) → CKm
(Ω) is
∞
continuous. This is because the local base in CKm (Ω) is
∞ 1
φ ∈ CKm (Ω) | kφkm < .
m
For any such φ in the local base, we have kφk` < 1/` and is in the local base
∞
of CK `
(Ω). Thus, we endow Cc∞ with the finest3 topology that makes the
∞
inclusion maps Ij : CK j
(Ω) → Cc∞ (Ω) continuous, for all j. In other words, a
set U in Cc∞ (Ω) is said to be open if and only if Ij−1 (U ) is open in CK
∞
j
(Ω) for
all j ≥ 1. Such a topology is called the inductive limit topology with respect
∞
to CK j
(Ω) and the maps Ik` . The space Cc∞ (Ω) is complete with respect
to the inductive limit topology because any Cauchy sequence is Cauchy in
∞ ∞
CK j
(Ω), for some j. Since CK j
(Ω) is closed, the space Cc∞ (Ω) is complete
∞
w.r.t the inductive limit topology. Though each CK (Ω) is metrizable, the
∞
space Cc (Ω) is not metrizable (cf. Exercise 8).
Exercise 7. Every proper subspace of a topological vector space has empty
interior.
Proof. Let X be a topological vector space and V ( X be a vector subspace
of X. We need to show that for x ∈ V there is an open set U containing
x such that U ⊂ V . It is enough to show the claim for x = 0 (0 ∈ V )
because if U contains 0, then U + {x} ⊂ V is an open set (due to continuity
of addition) containing x. Due to vector space structure U + {x} is in V .
For every x ∈ X, we can define a function fx : R → X as fx (λ) = λx. The
function fx is continuous because the scalar multiplication map from R × X
to X is continuous. Suppose that there is an open set V containing 0 such
that V ⊂ X. Then fx−1 (V ) will be an open set containing 0 ∈ R. Thus, for
any λ ∈ fx−1 (V ), we have λx ∈ V . Thus, x ∈ V . This argument is true for
all x ∈ X, thus X ⊂ V . This implies X = V , a contradiction.
Exercise 8. The inductive limit topology on Cc∞ (Ω) is not metrizable.
Proof. Recall that Cc∞ (Ω) = ∪∞ ∞ ∞
j=1 CKj (Ω), where each closed set CKj (Ω) has
empty interior (cf. Exercise 7). Therefore, the complete space Cc∞ (Ω) is a
countable union of no-where dense sets. If Cc∞ (Ω) was metrizable then it
would contradict the Baire’s category theorem.
3
strongest or largest topology, the one with more open sets
CHAPTER 1. THEORY OF DISTRIBUTIONS 19
Definition 1.3.16. The space Cc∞ (Ω) endowed with the inductive limit topol-
ogy, and denoted as D(Ω), is called the space of test functions.
Exercise 9 (cf. Exercise 4). Show that the topology defined on D(Ω) is
independent of the choice of (exhaustion sets) the sequence of compact sets
Kj of Ω.
The sequential characterisation of the inductive limit topology on D(Ω)
is given below. We define only for the zero converging sequence due to
continuity of the addition operation.
Theorem 1.3.18 (cf. Theorem ??). For 1 ≤ p < ∞, the class of all simple4
functions S(Ω) is dense in Lp (Ω).
Further, a step ahead, we show that the space of test functions, Cc∞ (Ω) is
densely contained in Lp (Ω). This result is established by using the technique
of regularization by convolution introduced by Leray and Friedrichs.
4
By our definition, simple function is non-zero on a finite measure. A simple function
φ is a non-zero function on Rn having the (canonical) form
k
X
φ(x) = ai 1Ei
i=1
with disjoint measurable subsets Ei ⊂ Rn with µ(Ei ) < +∞ and ai 6= 0, for all i, and
ai 6= aj for i 6= j.
CHAPTER 1. THEORY OF DISTRIBUTIONS 20
kf ∗ gkp ≤ kf k1 kgkp .
5
Proposition 1.3.23. Let f ∈ L1 (Rn ) and g ∈ Lp (Rn ), for 1 ≤ p ≤ ∞.
Then
supp(f ∗ g) ⊂ supp(f ) + supp(g)
∂ρε (x − y)
Z
= f (y) dy
Bε (x) ∂xi
(interchange of limits is due to the uniform convergence)
∂ρε (x − y)
Z
∂ρε
= f (y) dy = ∗ f.
Ω ∂xi ∂xi
Similarly, one can show that, for any tuple α, Dα fε (x) = (Dα ρε ∗ f )(x).
Thus, uε ∈ C ∞ (Ωε ).
Proposition 1.3.25. 6 Let f ∈ Cck (Rn ) (k ≥ 1) and let g ∈ L1loc (Rn ). Then
f ∗ g ∈ C k (Rn ) and for all |α| ≤ k
Dα (f ∗ g) = Dα f ∗ g = f ∗ Dα g.
Theorem 1.3.26 (Regularization technique). C ∞ (Rn ) is dense in C(Rn )
under the uniform convergence on compact sets topology.
Proof. Let g ∈ C(Rn ) and K ⊂ Rn be a compact subset. Note that g is
uniformly continuous on K. Hence, for every η > 0, there exist a δ > 0
(independent of x and dependent on K and η) such that |g(x − y) − g(x)| <
η whenever |y| < δ for all x ∈ K. For each m ∈ N, set ρm := ρ1/m ,
the sequence of mollifiers. Define gm := ρm ∗ g. Note that gm ∈ C ∞ (Rn )
(Dα gm = Dα ρm ∗ g). Now, for all x ∈ Rn ,
Z Z
|gm (x) − g(x)| = g(x − y)ρm (y) dy − g(x) ρm (y) dy
|y|≤1/m |y|≤1/m
Z
≤ |g(x − y) − g(x)|ρm (y) dy
|y|≤1/m
6
Refer Brezis for proof
CHAPTER 1. THEORY OF DISTRIBUTIONS 22
Since the δ is independent of x ∈ K, we have kgm − gk∞ < η for all m > 1/δ.
Hence, gm → g uniformly on K.
Theorem 1.3.27. For any Ω ⊆ Rn , Cc∞ (Ω) is dense in Cc (Ω) under the
uniform topology.
Proof. Let g ∈ Cc (Ω) and K := supp(g). One can view Cc (Ω) as a subset of
Cc (Rn ) under the following identification: Each g ∈ Cc (Ω) is extended to Rn
as g̃ (
g(x) x ∈ K
g̃(x) =
0 x ∈ Rn \ K.
Corollary 1.3.28. For any Ω ⊆ Rn , Cc∞ (Ω) is dense in C(Ω) under the
uniform convergence on compact sets topology.
The first term converges to zero by Theorem 1.3.29 and the second term
converges to zero by Dominated convergence theorem.
The case p = ∞ is ignored in the above results, because the L∞ -limit of
ρm ∗ f is continuous and we do have discontinuous functions in L∞ (Ω).
Theorem 1.3.31 (Kolmogorov Compactness Criteria). Let p ∈ [1, ∞) and
let A be a subset of Lp (Rn ). Then A is relatively compact in Lp (Rn ) iff the
following conditions are satisfied:
7
The type of functions, φk , are called cut-off functions
CHAPTER 1. THEORY OF DISTRIBUTIONS 24
Let (ρn )n∈N be a mollifier. It follows from Theorem 1.3.30 that, for all n ≥ 1
and f ∈ Lp (Rn )
Z
p
kf − f ∗ ρn kp ≤ ρn (y)kf − τy f kpp dy.
Rn
Hence
kf − f ∗ ρn kp ≤ sup kf − τy f kp .
1
|y|≤ n
kf − f ∗ ρN (ε) kp < ε.
The last inequality follows from the invariance property of the Lebesgue
measure. Moreover,
|(f ∗ ρn )(x)| ≤ kf kp kρn kq .
Let us consider the family A = {f ∗ ρN (ε) : Br (0) → R | f ∈ A}. By
using (i) and (iii), and Ascoli-Arzela result, we observe that A is relatively
CHAPTER 1. THEORY OF DISTRIBUTIONS 25
compact w.r.t the uniform topology on Br (0). Hence, there exists a finite set
{f1 , . . . , fk } ⊂ A such that
Thus, for all f ∈ A, there exists some j ∈ {1, 2, . . . , k} such that, for all
x ∈ Br (0)
|f ∗ ρN (ε) (x) − fj ∗ ρN (ε) (x)| ≤ ε|Br (0)|−1/p .
Hence,
Z 1/p Z 1/p
p p
kf − fj kp ≤ |f | dx + |fj | dx
|x|>r |x|>r
+kf − f ∗ ρN (ε) kp + kfj − fj ∗ ρN (ε) kp
+kf ∗ ρN (ε) − fj ∗ ρN (ε) kp,Br (0) .
Finally,
kf − fj kp ≤ 5ε
and, hence, A is precompact in Lp (Rn ).
Hint. For each compact set K there is an positive integer i0 such that K ⊂
Ki , for all i ≥ i0 .
It is enough to define for zero convergent sequences because addition op-
eration is continuous. For a first countable space the notion of continuity and
sequential continuity are equivalent. A Hausdorff topological vector space is
metrizable iff it is first countable. We know that D(Ω) is not metrizable
(cf. Exercise 8) and hence cannot be first countable.
Theorem 1.4.2. Let T : D(Ω) → R be a linear map. Then the following are
equivalent:
(ii) For every compact subset K ⊂ Ω, there exists a constant CK > 0 and
an integer NK ≥ 0 (both depending on K) such that
∞
|T (φ)| ≤ CK kφkNK , ∀φ ∈ CK (Ω).
Proof. ((i) =⇒ (ii)): Let T be continuous on D(Ω). Then, for any compact
∞
subset K ⊂ Ω, the restriction of T to CK (Ω) is continuous (cf. Exercise 10).
∞
The inverse image of (−c, c) under T is open set in CK (Ω) containing origin.
∞
Since CK (Ω) is first countable (normed space), there is a local base at 0.
∞
Thus, there is a NK and for all φ ∈ CK (Ω) such that kφkNK ≤ 1/NK , we
∞
have |T (φ)| < c. Thus, for φ ∈ CK (Ω),
φ
T <c
NK kφkNK
if f ∈ Lploc (Ω) for all 1 < p < ∞, then f ∈ L1loc (Ω). Because the Hölder’s
inequality implies that for any compact subset K of Ω,
Z Z 1/p
p
|f (x)| dx ≤ |f | dx (µ(K))1/q < +∞.
K K
Thus, f ∈ L1loc (Ω). Since any Lp (Ω), for 1 ≤ p < ∞, is in Lploc (Ω) and hence
are in L1loc (Ω).
Exercise 13. If f is continuous on Ω, then f ∈ L1loc (Ω).
We shall now observe that to every locally integrable function one can
associate a distribution. For any f ∈ L1loc (Ω), we define the functional Tf on
D(Ω) defined as, Z
Tf (φ) = f (x)φ(x) dx.
Ω
Thus, f ∈ L1loc (Ω). The function 1K denotes the characteristic function which
takes 1 on K and zero on the K c .
CHAPTER 1. THEORY OF DISTRIBUTIONS 30
then f ≡ 0 in Ω.
Proof. Let us assume f is continuous. If f 6= 0 in Ω, then there is a x0 ∈
Ω such that f (x0 ) = λ 6= 0. Define the function R g = f /λ which is also
continuous on Ω such that g(x0 ) = 1. Note the gφ = 0 for all φ ∈ D(Ω).
By the continuity of g, there is an r > 0 such that, for all x ∈ Br (x0 ) ⊂ Ω,
g(x) R> 1/2. Choose a test function ψ ∈ D(Ω) such that supp(ψ)⊂ Br (x0 )
and Ω ψ(x) dx = 1. For instance, choose ψ to be the mollifier function
ψ(x) = ρr (x − x0 ). Then, in particular,
Z Z Z
1 1
0= gψ dx = gψ dx > ψ dx = > 0,
Ω Br (x0 ) 2 Br (x0 ) 2
then f = 0 a.e. in Ω.
Proof. Note that it is enough to prove that f = 0 a.e. on all compact subsets
K of Ω. For any compact subset K of Ω, choose φK ∈ D(Ω) such that
φK ≡ 1. Define, for each x ∈ K,
Z
fε (x) := f φK ∗ ρε (x) = f (y)φK (y)ρε (x − y) dy.
K
where |µ| denotes the total variation of the measure. The distribution Tµ is
of zero order.
CHAPTER 1. THEORY OF DISTRIBUTIONS 32
Proof. We shall give two proofs: The first proof is the usual measure theory
technique, If g = 1E , the characteristic function of E, then
Z Z
1E dδa = dδa = δa (E) = 1E (a).
Ω E
Pk
For any simple function g = i=1 αi 1Ei , then
Z k
X
g(x) dδa = αi δa (Ei ) = g(a).
Ω i=1
8
Yes! we mean “any” function, because the class of Dirac measurable sets is the power
set and hence any function is Dirac measurable. Though, in the above context we have
restricted the Dirac measure to the class of Borel σ-algebra
CHAPTER 1. THEORY OF DISTRIBUTIONS 33
E := {x ∈ Ω/g(x) 6= h(x)}.
Hence, a ∈
/ E and therefore δa (E) = 0. This implies that g = h almost
everywhere w.r.t δa and, hence, their integrals coincide
Z Z
g(x) dδa = h(x)dδa .
Ω Ω
But Z
h(x)dδa = g(a)δa (Ω) = g(a).
Ω
With the above observations, we note that the Dirac distribution is just
R
Therefore, 1 ≤ B(a;ε) |f | dx. Since f ∈ L1loc (Ω), the quantity on RHS is finite.
Hence, as ε → 0, we get a contradiction 1 ≤ 0.
There are many more measures that induce singular distribution. In fact,
one can identify the class of measures which induce regular distribution.
They are, precisely, the absolutely continuous9 measures w.r.t the Lebesgue
measure. We shall not dwell on this topic, but a quick summary is as follows:
Note that for each f ∈ L1loc (Ω), one can define a signed measure
Z
µf (E) := f dx (1.4.2)
E
Note that so far we have only seen examples of distributions which are of
zero order. This is because of the following result:
0 ≤ T (λψ(x) − φ(x))
T (φ) ≤ λT (ψ) = T (ψ)kφk0 .
The inequality |µ|(B(0; 1)) ≥ m, for all m ∈ N, implies that |µ|(B(0; 1))
is infinite which contradicts the fact that the Radon measure µ is finite on
compact subsets of R.
CHAPTER 1. THEORY OF DISTRIBUTIONS 38
We shall see later that the multipole distribution δaα is, in fact, the α-th
distributional derivative of the Dirac measure δa up to a sign change. This
motivation was behind the choice of notation of multipole distribution. In
this section, we have noted that the space of Radon measures is properly
contained in the space of distributions, i.e., R(R) ( D0 (R).
Exercise 20. Do we have singular distributions which are neither induced by
Radon measures nor distributional derivatives of Radon measures? (Hyper-
functions?)
The larger the compact set becomes, the higher the derivatives of φ needs to
be taken in. Thus, there is no fixed m for all compact subsets of R.
Example 1.15. Define the functional T ∈ D0 (0, ∞) as
∞
X
T (φ) = φ(k) (1/k).
k=1
For every φ ∈ D(Ω), one can define the linear functional Λφ : D0 (Ω) → R as
follows, Λφ (T ) = T (φ). The linear functionals Λφ are included in the second
(algebraic) dual of D(Ω). We consider the coarsest10 topology on D0 (Ω) such
that all the linear maps Λφ : D0 (Ω) → R, corresponding to each φ ∈ D(Ω),
are continuous. This is called the weak-* topology on D0 (Ω). For every open
subset of V ⊂ R, consider the collection of subsets {Λ−1 0
φ (V )} of D (Ω), for all
φ ∈ D(Ω). The weak-* topology is the topology generated by this collection
of subsets in D0 (Ω). The space D0 (Ω) is sequentially weak-* complete but
not weak-* complete.
∞
and hence T is continuous on CK (Ω) for all compact subsets K of Ω. Thus,
T is continuous on D(Ω) and hence T ∈ D0 (Ω).
10
weakest or smallest topology, one with fewer open sets
CHAPTER 1. THEORY OF DISTRIBUTIONS 40
Note that they converge point-wise to zero for all x ∈ R. Let Tm denote the
distribution corresponding to fm and hence
Z 1/m Z 2/m
2 2 2
Tm (φ) = m xφ(x) dx + m − x φ(x) dx.
0 1/m n
Both the integral above converges to φ(0)/2 and hence the sequence of dis-
tributions converges to δ0 . Let us give the proof for the first integral. Since
Z 1/m
2 1
m x dx =
0 2
we consider
Z 1/m Z 1/m
2 1 2
m xφ(x) dx − φ(0) = m x[φ(x) − φ(0)] dx
0 2 0
Z 1/m
2
≤ m |x||φ(x) − φ(0)| dx.
0
Hence,
Z 1/m
2 1
lim m xφ(x) dx = φ(0).
m→∞ 0 2
Similarly,
R one can show that the second integral is also (1/2)φ(0) and hence
fm (x)φ(x) dx = φ(0).
The following theorem gives the condition that is violated by the above
example for the point-wise limit to coincide with the distributional limit.
T (φ) = ki=1 Ti (φ). The topology on D0 (Ω) can be used to give the notion
P
of series of distributions.
Definition 1.4.16.
P∞ For any countable collection of distributions {Ti }∞
1 ⊂
0 0
D (Ω) the series i=1 T
Pm i is said to converge to S ∈ D (Ω) if the sequence of
0
partial sums Tm := i=1 Ti converges to S in D (Ω).
exists. We say that the principal value of the integral exists if the limit (with
b = a)
Z b
lim f (x) dx
a→+∞ −a
exists.
exists. We say that the principal value of the integral exists if the limit (with
b = a) Z
lim+ f (x) dx
a→0 |x|>a
exists.
We have already seen that every locally integrable function can be identi-
/ L1loc (R).
fied with a distribution. We have already noted, in §1.4.1, that 1/x ∈
Consider the function fε : R → C, for each ε > 0, defined as
1
fε (x) = .
x + iε
R1
Thus, we expect −1 (1/x) dx = −iπ. This cannot be made sense classically11 .
Using the theory of distributions one can give a meaning to this observation.
We do know that 1/x ∈ L1loc (R \ {0}) and hence induces a distribution
in D(R \ {0}). One can extend this distribution to yield a distribution cor-
1
responding 1/x on R. We define the linear functional PV x on D(R) as
Z
1 1
PV (φ) = lim φ(x) dx. (1.4.3)
x ε→0 |x|≥ε x
As defined before, the limit on RHS is called the Cauchy’s Principal Value
and hence the use of ‘PV’ in the notation. Note that P V (1/x) is defined as
a distributional limit of the sequence of distributions corresponding to the
locally integrable functions
(
1
|x| > ε,
fε (x) := x
0 |x| ≤ ε,
1
in D0 (Ω).
i.e., fε * P V x
Lemma 1.4.19. The principal value of 1/x (limit on the RHS) exists.
Proof. Let φ ∈ D(R). Without loss of generality, let supp(φ)⊂ [−a, a] for
some real a > 0. For ε > 0 small enough, consider
Z Z −ε Z a
φ(x) φ(x) φ(x)
dx = dx + dx
|x|≥ε x −a x ε x
Z a Z a Z a
φ(x) φ(−x) φ(x) − φ(−x)
= dx − dx = dx.
ε x ε x ε x
11
recall (1.1.1)
CHAPTER 1. THEORY OF DISTRIBUTIONS 45
R1
Set ψ(x) = −1
φ0 (xt) dt, then
Z a Z a Z a
1 1
PV (φ) = xψ(x) dx = ψ(x) dx ≤ |ψ(x)| dx.
x 0 x 0 0
The principal value distribution is not the only choice for the function
1/x. Consider the linear functional on D(R) corresponding to 1/x,
Z −ε Z ∞
φ(x) φ(x)
T (φ) = lim dx + dx.
ε→0 −∞ x 2ε x
Exercise 22. Show that the T defined above is a distribution. Also, compare
T and PV(1/x).
Proof. Let φ ∈ D(R). Without loss of generality, let supp(φ)⊂ [−a, a] for
some real a > 0. For ε > 0 small enough, consider
Z −ε Z ∞ Z −2ε Z −ε
φ(x) φ(x) φ(x) φ(x)
dx + dx = dx + dx
−∞ x 2ε x −a x −2ε x
Z a
φ(x)
+ dx
2ε x
Z a Z −ε
φ(x) − φ(−x) φ(x)
= dx + dx
2ε x −2ε x
Z a Z 2
φ(−εx)
= ψ(x) dx − dx.
2ε 1 x
Therefore,
Z a Z 2 Z a
φ(0)
T (φ) = ψ(x) dx − dx = ψ(x) dx − φ(0) ln 2.
0 1 x 0
Thus, |T (φ)| ≤ 2akφ0 k0 + kφk0 ln 2 ≤ Ckφk1 where C = max(2a, ln 2). T is
again a distribution of order one. Now, consider
Z 2ε Z 2
φ(x) φ(εx)
(PV(1/x) − T )(φ) = lim dx = lim dx = φ(0) ln 2.
ε→0 ε x ε→0 1 x
Now, choose m > N and ψ such that supp(ψ)⊂ (1/(m+1), 1/(m−1)). Then
T (ψ) = ψ (m) (1/m). Since T and S coincide for ψ ∈ D(0, ∞), we have
(m)
|δ1/m (ψ)| = |ψ (m) (1/m)| = |T (ψ)| = |S(ψ)| ≤ CkψkN .
(m)
Thus, we have that the order of the multipole distribution δ1/m is at most
N , a quantity smaller than m. This is a contradiction. Hence, we can have
no extension S of T to R.
1.5.1 Differentiation
Recall the discussion leading to (1.3.1). For any f ∈ C ∞ (R), φ ∈ D(R) and
all k ∈ N, using integration by parts we have
Z Z
(k) k
f φ dx = (−1) f φ(k) dx.
|(Dα T )(φ)| = |(−1)|α| T (Dα φ)| = |T (Dα φ)| ≤ CK kDα φkN ≤ CK kφkN +|α| .
(ii) Consider
where (
1 if 0 < x ≤ 1
v(x) =
0 if 1 < x < 2.
Thus, Du = v.
Proof. Consider
Z 2 Z 1 Z 2
0 0
Du(φ) = − uφ dx = − xφ dx − 2 φ0 dx
0 0 1
Z 1
= φ dx + φ(1).
0
0
Then, Du = v + δ1 in D (Ω) where
(
1 if 0 < x ≤ 1
v=
0 if 1 < x < 2.
Observe above that the function Ha had a jump discontinuity and Dirac
measure appeared as its derivative at the point of “jump”. This is the feature
of Dirac measure. It appears as a derivative at points of jump. However,
note that in Example 1.23 and 1.24, the jump at every point did not give
rise to a Dirac measure because outside the set of jump points (in that case
Q) the function was continuous.
Example 1.26. Consider the locally integrable function f : R → R defined as
(
|x| |x| < a
f (x) =
0 |x| > a.
Therefore, DTf = aδ−a −aδa +Tf 0 . The function f is not weakly differentiable
(compare this with Example 1.22).
Example 1.27. Consider the discontinuous function
(
x2 + x x < 1
f (x) =
e−5x x > 1.
CHAPTER 1. THEORY OF DISTRIBUTIONS 53
(
2x + 1 x<1
f 0 (x) =
−5e−5x x > 1.
Z 1 Z ∞
Tf 0 (φ) = (2x + 1)φ(x) dx − 5 e−5x φ(x) dx.
−∞ 1
Z 1 Z ∞
0 0
DTf (φ) = −Tf (φ ) = − 2
(x + x)φ (x) dx − e−5x φ0 (x) dx
−∞ 1
Z 1 Z ∞
2 1
= −(x + x)φ(x) |−∞ + (2x + 1)φ(x) dx − 5 e−5x φ(x) dx
−∞ 1
− e−5x φ(x) |∞
1
−5
= (e − 2)φ(1) + Tf 0 (φ).
Example 1.28. Consider the Cantor function fC : [0, 1] → [0, 1] (cf. Ap-
pendix A) extended continuously to R by setting
(
0 x≤0
fC (x) =
1 x ≥ 1.
X∞ Z bk
= φ(1) − ck φ0 dx
k=1 ak
Exercise 26. For any β ∈ R, we have already noted that the function f (x) =
|x|−β is in L1loc (Rn ) for β < n. However, the function is weakly differentiable
for β + 1 < n and its weak partial derivative is fxi (x) = − |x|ββ+1 |x|
xi
.
R
Exercise 27. Let φ ∈ D(R). Show that R φ(x) dx = 0 iff there is a ψ ∈ D(R)
such that φ(x) = ψ 0 (x).
Proof. Let there exist a ψ ∈ D(R) such that φ(x) = ψ 0 (x). Then,
Z Z a
φ(x) dx = lim ψ 0 (x) dx = lim [ψ(a) − ψ(−a)] = 0.
R a→∞ −a a→∞
Rx
Conversely, let R φ(x) dx = 0. Then, set ψ(x) := −∞ φ(t) dt and ψ 0 = φ.
R
0 = DT (φ) = −T (φ0 )
WeR have T (ψ) = 0 for all ψ ∈ E. Using Exercise 27,R we know that ψ ∈ E
iff R ψ(x) dx = 0. Choose any χ ∈ D(R)R such that R χ = 1 and, for each
φ ∈ D(R), set ψ := φ − βχ where β = R φ. Then ψ ∈ E and T (ψ) = 0.
Therefore, Z
T (φ) = βT (χ) = T (χ)φ(x) dx.
R
Thus, T = Tλ where λ = T (χ), a constant.
An alternate proof of above exercise is given in Exercise 45. By “regular
distribution generated by a constant function” we mean the function is con-
stant except on a set of measure zero. (cf. Example 1.23). The same is not
true with weak derivative. If the weak derivative is zero, then the function
is constant on each connected component of Ω.
CHAPTER 1. THEORY OF DISTRIBUTIONS 56
Thus, the two notions of derivatives coincide for all functions which respect
integration by parts.
Then,
Z b
[F (x)g(x) + G(x)f (x)] dx = F (b)G(b) − F (a)G(a).
a
Consider the square [a, b] × [a, b] in R2 and its bisection by the line y = x in
R2 . Then the first integral is evaluated in the region above the bisecting line
and the second integral is evaluated below the bisecting line. Thus, combined
together the region of computation is the square [a, b] × [a, b]. Hence,
Z b Z bZ b
[F g + Gf ] dt = f (x)g(y) dx dy + F (a)G(b) + G(a)F (b)
a a a
− 2F (a)G(a)
Z b Z b
= f (x) dx g(y) dy + F (a)G(b) + G(a)F (b)
a a
− 2F (a)G(a)
= [F (b) − F (a)][G(b) − G(a)] + F (a)G(b) + G(a)F (b)
− 2F (a)G(a)
= F (b)G(b) − F (a)G(a).
The above proposition shows that, in one dimensional case, the inte-
gration by parts formula can be extended, precisely, to the class of abso-
lutely continuous functions. Every absolutely continuous function on R is of
CHAPTER 1. THEORY OF DISTRIBUTIONS 58
This shows that |u| has a weak gradient and the formula is obtained.
Example 1.31. The above theorem is very special to the space of distributions
D0 (Ω). For ∞
√ instance, the result is not true in C (Ω). Consider the sequence
fm = (1/ m) sin mx that converges uniformly to 0 and hence converges√ to
0
0 in the distribution sense too. However, it’s derivative fm (x) = m cos mx
does not converge pointwise but converges to 0 in the distribution sense.
1.5.2 Product
Now that we have addition of distributions, a natural question is whether one
can define product of any two distributions. The answer is in negation. The
space of distribution cannot be made an algebra which extends the classical
notion of point-wise multiplication. This is called the Schwartz impossibility
result. L. Schwartz himself showed that it is not possible to define product
of Dirac distributions, i.e., δ0 · δ0 . However, one may define product of a
CHAPTER 1. THEORY OF DISTRIBUTIONS 60
≤ CK C0 kφkNK .
(ii) For any φ ∈ D(R), since 0 = xT (φ) = T (xφ). By Exercise 32, any
solution T ∈ D0 (R) satisfies T (ψ) = 0 for all ψ ∈ D(R) such that
ψ(0) = 0. Obviously, T ≡ δ0 is a solution. In fact, for any scalar
λδ0 ∈ D0 (R) is also a solution. We need to show that any solution is a
scalar multiple of δ0 . Let T ∈ D0 (R) be a solution of xT = 0. Hence
T (ψ) = 0 for all test functions on R such that ψ(0) = 0. Let χ ∈ D(R)
such that χ(0) = 1, then for any φ ∈ D(R), ψ(x) := φ(x) − φ(0)χ(x) is
a test function such that ψ(0) = 0. Therefore,
Thus, T (φ) = λδ0 (φ) where the choice of λ depends on the choice of χ.
(iii) By Exercise 32, any solution T ∈ D0 (R) satisfies T (ψ) = 0, for all
ψ ∈ D(R), such that ψ (k) (0) = 0 for all k = 0, 1, 2, . . . , 2012. Therefore,
(k)
δ0 is a solution for all k = 0, 1, 2, . . . , 2012. Thus,
2012
X (k)
T ≡ λk δ0
k=0
In light of the above example, we come back to the issue of not being
able to have a product in the space of distributions. Suppose we give a
notion of product of distributions, the product cannot be associative because
P V (1/x) · (x · δ0 ) = 0 6= δ0 = (P V (1/x) · x) · δ0 . Further, recall that
f (x) = |x|−1/2 ∈ L1loc (R), but g(x) = |x| 6∈ L1loc (R). How do we give a
notion of product, in a suitable way, such that (Tf )2 coincides with classical
pointwise product. The failure to define a suitable notion of product is
what makes the theory of distributions unsuitable for nonlinear differential
equations.
Exercise 35. For any given ψ ∈ C ∞ (R), find the solutions of the equation
DT + ψT = 0 in D0 (R).
Exercise 36. For any given ψ ∈ C ∞ (R) and S ∈ D0 (R), find the solutions of
the equation DT + ψT = S in D0 (R).
Recall the topology on C ∞ (Ω) described in §1.3.3. The sequential de-
scription of the topology is the uniform convergence on compact sets. Thus,
a sequence ψm is said to converge to 0 in C ∞ (Ω) if Dα ψm , for all α, uniformly
converges to 0 on all compact subsets of Ω.
1.5.3 Support
The fact that we cannot have a product on D0 (Ω), generalising point-wise
multiplication, that makes D0 (Ω) an algebra leads us to look for other ways
to make D0 (Ω) an algebra. One such choice is the convolution operation and
whether this can be extended to distributions. We shall see later than one
can define a notion of convolution for distributions provided one of them
has compact support! This motivates us to understand the support of a
distribution which coincides for classical functions. Classically, support of a
function is complement of the largest open set on which the function vanishes.
Theorem 1.5.17. The dual space E 0 (Ω) is the collection of all distributions
with compact support.
CHAPTER 1. THEORY OF DISTRIBUTIONS 66
1.5.6 Convolution
Recall the definition of convolution of functions given in Definition 1.3.20.
Thus, for any x ∈ Rn ,
Z
(f ∗ g)(x) := f (x − y)g(y) dy.
Rn
T ∗(φ1 +φ2 )(x) = T (τx (φ1 +φ2 )ˇ) = T (τx φˇ1 +τx φˇ2 ) = (T ∗φ1 +T ∗φ2 )(x).
(ii) Consider
(S + T ) ∗ φ = (S + T )(τx φ̌) = S ∗ φ + T ∗ φ.
CHAPTER 1. THEORY OF DISTRIBUTIONS 70
Definition 1.5.27. Let S, T ∈ D0 (Rn ) such that one of them has compact
support, i.e. either T or S is in E 0 (Rn ). We define the convolution S ∗ T as,
(S ∗ T ) ∗ φ = S ∗ (T ∗ φ) ∀φ ∈ D(Rn ).
Exercise 44. For any three distributions T1 , T2 , T3 ∈ D0 (Rn ) such that at least
two of them have compact support then
T1 ∗ (T2 ∗ T3 ) = (T1 ∗ T2 ) ∗ T3 .
Proof. Since δ0 has compact support {0}, the convolution makes sense. By
commutativity, T ∗ δ0 = δ0 ∗ T . For φ ∈ D(Rn ), consider (T ∗ δ0 )(φ) =
(T ∗ (δ0 ∗ φ̌))(0). Set ψ := δ0 ∗ φ̌. Then (T ∗ δ0 )(φ) = T (ψ̌). Note that
(1) (1)
Show that (T ∗ δ0 ) ∗ H 6= T ∗ (δ0 ∗ H).
Proof. Consider
Z
(1) (1)
(1 ∗ δ0 )(φ) =1∗ (δ0 ∗ φ̌)(0) = (1 ∗ ψ)(0) = T1 (ψ̌) = ψ̌(x) dx.
R
CHAPTER 1. THEORY OF DISTRIBUTIONS 72
(1) (1)
But ψ̌(x) = δ0 ∗ φ̌(−x) = δ0 (τ−x φ) = −τ−x φ0 (0) = −φ0 (x). Hence (1 ∗
(1)
δ0 )(φ) = − R φ0 (x) dx = 0. Further, 0 ∗ H = 0. Now, consider
R
and if p = ∞,
Hence Tˆf = Tfˆ and, hence, both definitions of the Fourier transform coincide
on S(Rn ).
Remark 1.6.4. Since S(Rn ) ⊂ L2 (Rn ) ⊂ S 0 (Rn ), then there are two defi-
nitions of Fourier transform on L2 (Rn ) inherited from S 0 (Rn ) and the other
by extending from S(Rn ). It turns out that both are same. Consider the
weak-* convergence of tempered distributions, i.e., Tm → T weak-* sense if,
for every φ ∈ S(Rn ), Tm (φ) → T (φ). If Tm → T in S 0 (Rn ) then Tˆm → T̂ in
S 0 because
T̂m (φ) = Tm (φ̂) → T (φ̂) = T̂ (φ).
Let f ∈ L2 (Rn ) and let fk ∈ S(Rn ) such that fk → f in L2 (Rn ). Since the
inclusion of L2 (Rn ) ⊂ S 0 (Rn ) is continuous, fk → f in S 0 (Rn ), as well. Hence
fˆk → fˆ in S 0 (Rn ). On the other hand, F(fk ) → F(f ) in L2 (Rn ) and hence in
S 0 (Rn ). But as fk ∈ S(Rn ), we know that F(fk ) = fˆk . Now, by uniqueness
of the weak-* limit fˆ = F(f ).
Dα T̂ = (−ı)|α| (x
\ αT )
and
\
(D α T ) = (ı)|α| ξ α T̂ .
\
(x α T )(f ) = (xα T )(fˆ) = T (xα fˆ(x))
1 1
= |α|
T (Ddαf ) = T̂ (Dα f )
(ı) (ı)|α|
1
= |α|
Dα T̂ (f ).
(−ı)
Thus, the first relation is proved. The second relation can be proved similarly.
\
∂δ0
Example 1.40. δ̂0 = 1, ∂xk
(ξ) = ıξk and 1̂ = δ0 . Let φ ∈ S(Rn ). Then
Z
δˆ0 (φ) = δ0 (φ̂) = φ̂(0) = φ(x) dx.
Rn
CHAPTER 1. THEORY OF DISTRIBUTIONS 75
Thus, δˆ0 = 1, i.e., the distribution induced by the constant function 1. Fur-
ther,
\
∂δ0
(ξ) = ıξk δˆ0 = ıξk .
∂xk
The 1̂ = δ0 follows by the Fourier inversion formula.
(i) L is hypoelliptic.
Sobolev Spaces
In this chapter we develop a class of functional spaces which form the right
setting to study partial differential equations. Recall that Ω ⊂ Rn is open
and not necessarily bounded. The situations where Ω needs to be bounded
will be specified clearly. Before we describe Sobolev spaces in detail, we
introduce a classical functional space very useful in the theory of Sobolev
spaces.
However, it is not always necessary that the supremum is finite. Note that
the modulus in the numerator and denominator are in R and Rn , respectively.
Definition 2.1.1. Let γ ∈ (0, 1]. We say a function u : Ω → R is Hölder
continuous of exponent γ at x0 ∈ Ω, if
|u(x) − u(x0 )|
pγ (u)(x0 ) := sup < +∞.
x∈Ω |x − x0 |γ
x6=x0
77
CHAPTER 2. SOBOLEV SPACES 78
and is denoted by C 0,γ (Ω). The quantity pγ (u) is called the γ-th Hölder
coefficient of u. If the Hölder coefficient is finite on every compact subsets of
Ω, then u is said to be locally Hölder continuous with exponent γ, denoted
0,γ
as Cloc (Ω).
Note that pγ (u) = supx0 ∈Ω pγ (u)(x0 ). In other words, uniformly Hölder
continuous function do not include all those that are Hölder continuous at
all points of Ω. However, bounded locally Hölder continuous functions are
Hölder continuous at all points of Ω.
Example 2.1. Note that, for all β < 0, u(x) = |x|β on the open ball B1 (0) ⊂
Rn do not belong to C(B1 (0)). For β ∈ [0, ∞), |x|β ∈ C(B1 (0)) and, for
β ∈ [2, ∞) ∪ {0}, |x|β ∈ C 1 (B1 (0)). For β ∈ (0, 1], |x|β is Hölder continuous,
with exponent γ, for all 0 < γ ≤ β, but is not Hölder continuous for γ > β. In
particular, |x|β ∈ C 0,β (B1 (0)) for each β ∈ [0, 1]. Thus, we have one example
from each of the space C 0,β (B1 (0)). We first check the Hölder continuity at
x0 = 0. The γ-th Hölder coefficient is
|x|β
sup γ
= |x|β−γ ≤ 1 for γ ≤ β.
x∈B1 (0) |x|
For γ > β the γ-th Hölder coefficient blows up. More generally, for x 6= 0
and |x| > |y| (wlog),
β
β−γ |y|
|x| 1 −
|x|β − |y|β |x|
= γ .
|x − y|γ x
− y
|x| |x|
CHAPTER 2. SOBOLEV SPACES 79
β
|y| |y| |y|
Since 0 < 1 − |x|
< 1 and 0 < β ≤ 1, 1 − |x|
≤ 1− |x|
. Thus,
β
|y|
|x|β−γ 1 − |x| |x|β−γ 1 − |y|
|x| ||x| − |y||
γ ≤ γ ≤ .
x
− y x
− y |x|1−γ |x − y|γ
|x| |x| |x| |x|
The last inequality is true for γ ≤ β and we have used |x|β−γ ≤ 1. For one
dimension, the last quantity is equal to 1 because
|y|
||x| − |y|| |x| 1 − |x| 1 − xy
= ≤ = 1.
|x|1−γ |x − y|γ |x||1 − xy |γ |1 − xy |
1 pγ (u)(x) γ
|uxi (x)| = lim |u(x+tei )−u(x)| ≤ lim |t| = lim pγ (u)(x)|t|γ−1 = 0.
t→0 |t| t→0 |t| t→0
≤ diam(Ω)δ−γ pδ (u).
Thus, pγ (u) < +∞ and u ∈ C 0,γ (Ω). The inclusions are strict by Exam-
ple 2.1. If γ <√δ, then u(x) = |x|γ is in C 0,γ (−1, 1) but not in C 0,δ (−1, 1).
In particular, x ∈ C 0,1/2 (−1, 1) but is not Lipschitz, i.e. do not belong to
C 0,1 (−1, 1).
A natural question at this juncture is: what is the relation between the
spaces C 1 (Ω) and C 0,1 (Ω). Of course |x| ∈ C 0,1 (−1, 1) and not in C(−1, 1).
Given the inclusion relation in above exercise, we would tend to believe that
C 1 (Ω) is contained in C 0,1 (Ω), but this is not true. In fact, only a subclass
of C 1 (Ω) belong to Lipschitz class.
Theorem 2.1.3. Let Ω be a convex domain. If u ∈ C 1 (Ω) and Dα u is
bounded on Ω, for each |α| = 1, then u ∈ C 0,1 (Ω).
Proof. Let u ∈ C 1 (Ω) such that there is a C0 > 0 such that |∇u(x)| ≤ C0
for all x ∈ Ω. For any two given points x, y ∈ Ω, we define the function
F : [0, 1] → R as F (t) = u((1 − t)x + ty). F is well defined due to the
convexity of Ω. Since u is differentiable, F is differentiable in (0, 1). Thus, for
all t ∈ (0, 1), F 0 (t) = ∇u[(1−t)x+ty]·(y −x). Therefore, |F 0 (t)| ≤ C0 |y −x|.
By mean value theorem, there is a ξ ∈ (0, 1) such that F 0 (ξ) = F (1) − F (0).
Thus,
Hence, ũ satisfies the Hölder estimate with exponent γ and is in C 0,γ (Ω).
The above result is very special to uniformly continuous spaces and,
hence, Hölder spaces. For instance, a similar result is not true between
C(Ω) and C(Ω). If Ω is compact then every function in C(Ω) is bounded,
where as C(Ω) might have unbounded functions. An interesting point in the
proof of above result is that the Hölder coefficient are same for the extension,
i.e., pγ (u) = pγ (ũ). An advantage of above theorem is that, without loss of
generality, we can use the space C 0,γ (Ω) and C 0,γ (Ω) interchangeably.
Exercise 50. Show that the Hölder coefficient, pγ , defines a semi-norm on
C 0,γ (Ω).
CHAPTER 2. SOBOLEV SPACES 82
Theorem 2.1.5. For any bounded open set Ω ⊂ Rn , the space C 0,γ (Ω) is a
Banach space with norm k · kC 0,γ (Ω) .
Proof. We need to prove the completeness of the space C 0,γ (Ω) w.r.t the norm
k · kC 0,γ (Ω) . Let {um } be a Cauchy sequence in C 0,γ (Ω), then {um } ⊂ C(Ω)
is Cauchy w.r.t the supremum norm. Thus, there is a u ∈ C(Ω) such that
kum − uk∞ → 0, as m → ∞. We first show that u ∈ C 0,γ (Ω). For x, y ∈ Ω
with x 6= y, consider
Since {um } is Cauchy, limm kum kC 0,γ (Ω) < ∞. Hence u ∈ C 0,γ (Ω). Finally,
we show that the sequence {um } converges to u in C 0,γ (Ω) w.r.t the γ-Hölder
norm. Consider,
|um (x) − u(x) − um (y) + u(y)| |um (x) − uk (x) − um (y) + uk (y)|
= lim
|x − y|γ k |x − y|γ
≤ lim sup pγ (um − uk ).
k
Therefore, limm pγ (um − u) ≤ limm lim supk pγ (um − uk ) and the RHS con-
verges to 0 since the sequence is Cauchy. Hence, kum − ukC 0,γ (Ω) → 0.
Exercise 53. The space C 0,γ (Ω) norm is not separable.2
For an unbounded open set Ω, we consider a sequence of exhaustion
compact sets, as described in §1.3.3, and make the space C 0,γ (Ω) a Fréchet
space.
Theorem 2.1.8. Let Ω be a bounded open subset of Rn . For any 0 < γ <
δ ≤ 1, the inclusion map I : C 0,δ (Ω) → C 0,γ (Ω) is continuous and compact.
Further, the inclusion map I : C 0,γ (Ω) → C(Ω) is compact, for all 0 < γ ≤ 1.
can assume that kum kC 0,δ (Ω) ≤ 1. Therefore, pδ (um ) ≤ 1 for all m which
implies that {um } is an equicontinuous sequence in C(Ω). By Arzelà-Ascoli
theorem, there exists a subsequence {umk } of {um } and a u ∈ C(Ω) such that
kumk − uk∞ → 0, as k → ∞. The u ∈ C(Ω) is, in fact, in C 0,δ (Ω) ⊂ C 0,γ (Ω)
because
|u(x) − u(y)| |umk (x) − umk (y)|
δ
= lim ≤1
|x − y| k |x − y|δ
and, hence, pδ (u) ≤ 1. We now show that the umk converges to u in C 0,γ (Ω).
For simplicity, set vk := umk − u and we will show that {vk } converges to 0 in
C 0,γ (Ω). Obviously, kvk k∞ → 0 thus it is enough to show that pγ (vk ) → 0.
Note that, for every given ε > 0, we have pγ (vk ) ≤ Skε + Tkε , where
ε |vk (x) − vk (y)|
Sk = sup
x,y∈Ω |x − y|γ
x6=y;|x−y|≤ε
and
|vk (x) − vk (y)|
Tkε = sup .
x,y∈Ω |x − y|γ
|x−y|>ε
Consider
|vk (x) − vk (y)|
Skε = sup |x − y|δ−γ ≤ εδ−γ pδ (vk ) ≤ 2εδ−γ .
x,y∈Ω |x − y|δ
x6=y;|x−y|≤ε
Similarly,
Tkε ≤ 2ε−γ kvk k∞ .
Hence, lim supk pγ (vk ) ≤ 2εδ−γ + 2ε−γ lim supk kvk k∞ = 2εδ−γ + 0. Since ε
can be made as small as possible, we have pγ (vk ) → 0.
We extend the Hölder continuous functions in to all differentiable class of
functions. We denote by C k,γ (Ω) the space of all C k (Ω), k-times continuously
differentiable, functions such that Dα u is Hölder continuous with exponent
γ, i.e., Dα u ∈ C 0,γ (Ω) for all |α| = k. For a bounded open set Ω, we give the
γ-th Hölder norm on C k,γ (Ω) as
k
X X
kukC k,γ (Ω) := kDα uk∞ + pγ (Dα u). (2.1.3)
|α|=0 |α|=k
Exercise 54. For any bounded open set Ω ⊂ Rn , show that k · kC k,γ (Ω) is
a norm on C k,γ (Ω). and the space C k,γ (Ω) is a Banach space with norm
k · kC k,γ (Ω) .
Proof. Let {um } be a Cauchy sequence in C k,γ (Ω). Then, {um } ⊂ C k (Ω)
is Cauchy in the supremum norm. Thus, there is a u ∈ C k (Ω) such that
kum − ukC k (Ω) → 0, as m → ∞. The fact that u ∈ C k,γ (Ω) and that u is a
limit, in the γ-Hölder norm, of the Cauchy sequence is similar to case k = 0
proved in Theorem 2.1.5.
Note that for, 1 < p < +∞, q is the number for which 1/p + 1/q = 1. In
this section, we shall define spaces that are analogous to C k (Ω) using the
generalised notion of derivative, viz. weak derivative (cf. Definition 1.5.2).
We know that every u ∈ Lp (Ω) being locally integrable induces a distribu-
tion Tu . Further, the distribution Tu is differentiable for all multi-indices α.
However, we have already seen that Tu need not be weakly differentiable.
For a fixed multi-index α, if there exists a vα ∈ Lp (Ω) such that Tvα = Dα Tu ,
then we denote the vα as Dα u. We know that such a vα is unique up to a set
of measure zero. As usual, Ω is an open subset of Rn .
Equivalently,
With this convention, we have W 0,p (Ω) = Lp (Ω). Note that, by definition,
if u ∈ W k,p (Ω) then every function in the equivalence class of u is also in
W k,p (Ω) and they all have their derivative as vα upto a set of measure zero.
Example 2.3. Let Ω = (−a, a) ⊂ R, for a positive real number a. Recall that
the function u(x) = |x| is not in C 1 (−a, a). But u ∈ W 1,p (−a, a). Consider
the v ∈ Lp (−a, a) defined as
(
1 x ∈ [0, a)
v(x) =
−1 x ∈ (−a, 0).
We shall show that v is the weak derivative of u. Consider, for φ ∈ D(−a, a),
Z a Z 0 Z a Z 0 Z a Z a
0 0
vφ dx = − φ dx + φ dx = xφ dx − xφ dx = − |x|φ0 dx.
−a −a 0 −a 0 −a
Example 2.4. The function v defined in above example is not in W 1,p (−a, a).
The argument is similar to the function w defined as
(
1 x ∈ [0, a)
w(x) =
0 x ∈ (−a, 0).
Note that w ∈ Lp (−a, a), for all p. However, the distributional derivative
of w (and v) is the Dirac measure at 0, δ0 , which is a singular distribution.
Hence w 6∈ W k,p (Ω), for all k > 0.
Example 2.5. The function u defined as
(
x x ∈ (0, a)
u(x) =
0 x ∈ (−a, 0]
is in W 1,p (−a, a), for all p ∈ [1, ∞], since both u and its distributional deriva-
tive Du = w are in Lp (−a, a) (cf. previous example). But u 6∈ W k,p (Ω), for
all k ≥ 2.
Note that we have the inclusion W k,p (Ω) ( W `,p (Ω) for all ` < k. The
inclusion is strict as seen from above examples.
Exercise 55. Find values of β ∈ R such that |x|β ∈ W 1,p (B1 (0)) for a fixed
p ∈ [1, ∞) and B1 (0) is the open unit ball in Rn .
Proof. Let u(x) = |x|β . Note that Dei u(x) = βxi |x|β−2 and |∇u(x)| =
β|x|β−1 . Consider
Z Z 1
k∇ukpp =β p p(β−1)
|x| p
dx = β ωn rpβ−p+n−1 dr.
Ω 0
The last quantity is finite iff p(β −1)+n−1 > −1, i.e., β > 1− np (also β = 0,
if not already included in the inequality condition) and |x|β ∈ W 1,p (Ω).
k
X k Z
X 1/p
α α p
kukk,p,Ω := kD ukp = |D u| .
|α|=0 |α|=0 Ω
k
X
kukk,∞,Ω = ess supΩ |Dα u|.
|α|=0
Proof. Let {um } be a Cauchy sequence in W k,p (Ω). Thus, by the definition
of the norm on W k,p (Ω), {um } and {Dα um }, for 1 ≤ |α| ≤ k are Cauchy in
Lp (Ω). Since Lp (Ω) is complete, there exist function u, vα ∈ Lp (Ω) such that
um → u and Dα um → vα ∀1 ≤ |α| ≤ k.
Remark 2.2.3. The proof above uses an important fact which is worth
noting. If um → u in Lp (Ω) and Dα um , for all |α| = 1, is bounded in Lp (Ω)
then u ∈ W 1,p (Ω).
CHAPTER 2. SOBOLEV SPACES 90
The first term with 1/m goes to zero, the second term tends to zero by
Theorem 1.3.29 and the last term goes to zero by Dominated convergence
3
The case k = 0 is precisely the result of Theorem 1.3.30
CHAPTER 2. SOBOLEV SPACES 91
theorem. Since
k
X
kum − ukk,p = kDα um − Dα ukp
|α|=0
k
X X
= Cα (Dα−β φm )(ρm ∗ Dβ u) − Dα u
|α|=0 β≤α p
and each term in the sum goes to zero as m increases, we have um converges
to u in W k,p (Rn ) norm.
The real trouble in approximating W k,p (Ω) by Cc∞ (Ω) functions is that
the approximation approach boundary continuously, if Ω has a boundary.
Proof. Let u ∈ W k,p (Ω). Fix ω relatively compact subset of Ω and a α such
that 1 ≤ |α| ≤ k. Let ũ denote the extension of u by zero in Ωc . Choose a
φ ∈ Cc∞ (Ω) such that φ ≡ 1 on ω and 0 ≤ φ ≤ 1 in Ω. Then supp(φu) ⊂ Ω
and we extend φu to all of Rn by zero in Ωc and denote it as φu.
f Set v = φuf
n k,p n
on R . Note that v ∈ W (R ). By Theorem 2.3.1, we have the sequence
vm := φm (ρm ∗ v) ∈ Cc∞ (Rn ) converging to v in α Sobolev norm. Since on ω,
v = u, we have
kvm − ukk,p,ω ≤ kvm − vkk,p,Rn → 0.
This is true for all relatively compact subset ω of Ω.
Note that the restriction ‘relatively compact set’ is only for k ≥ 1. For the
case k = 0, the restriction to Ω works, as seen in Theorem 1.3.30. The density
of Cc∞ (Ω) in W k,p (Ω) may fail to generalise for an arbitrary proper subset
Ω ⊂ Rn , because a “bad” derivative may be introduced at the boundary
while extending by zero outside Ω.
Example 2.6. Let Ω = (0, 1) ⊂ R and u ≡ 1 on Ω. Then u ∈ W 1,p (0, 1).
Setting ũ = 0 in R \ (0, 1), we see that ũ ∈ Lp (R) but not in W 1,p (R).
Because Dũ = δ0 − δ1 is not in L1loc (R).
CHAPTER 2. SOBOLEV SPACES 92
Recall the inclusion Cc∞ (Ω) ⊂ C ∞ (Ω) ⊂ C ∞ (Ω). The C ∞ (Ω) denotes all
functions in C ∞ (Ω) such that all its derivatives can be extended continuously
to Ω. For an arbitrary subset Ω ⊆ Rn the best one can do is the following
density result.
Theorem 2.3.3 (Meyers-Serrin). Let 1 ≤ p < ∞ and Ω ⊆ Rn . For u ∈
W k,p (Ω) and any ε > 0, there is a φ ∈ C ∞ (Ω) such that kφkk,p,Ω < ∞ and
ku − φkk,p,Ω < ε.
Proof. For each m ∈ N, consider the sets
1
ωm := {x ∈ Ω | |x| < m and dist(x, ∂Ω) > }
m
and set ω0 = ∅. Define the collection of open sets {Um } as Um := ωm+1 ∩
(ω m−1 )c . Note that Ω = ∪m Um is an open covering of Ω. Thus, we choose the
C ∞ locally finite partition of unity {φm } ⊂ Cc∞ (Ω) such that supp(φm ) ⊂ Um ,
k,p
P
0 ≤ φm ≤ 1 and m φm = 1. For the given u ∈ W (Ω), note that
φm u ∈ W (Um ) with supp(φm u) ⊂ Um . We extend φm u to all of Rn by zero
k,p
outside Um , i.e.,
(
φm u(x) x ∈ Um
φg
m u(x) =
0 x ∈ Rn \ Um .
k,p
Observe that φg mu ∈ W (Rn ). Let ρδ be the sequence of mollifiers and
∞ n
consider the sequence ρδ ∗φgm u in C (R ). Support of ρδ ∗φm u ⊂ Um +B(0; δ).
g
Note that for all x ∈ Um , 1/(m + 1) < dist(x, ∂Ω) < 1/(m − 1). Thus,
for all 0 < δ < 1/(m + 1)(m + 2)4 , supp(ρδ ∗ φg c
m u) ⊂ ωm+2 ∩ (ω m−2 ) ,
which is compactly contained in Ω. Since φm u ∈ W k,p (Ω), we can choose a
subsequence {δm } going to zero in (0, 1/(m + 1)(m + 2) such that
ε
kρδm ∗ φm u − φm ukk,p,Ω = kρδm ∗ φgm u − φm ukk,p,Rn < m .
2
∞
P
Set φ = m ρδm ∗ φm u. Note that φ ∈ C (Ω) and kφkk,p,Ω < ∞. Since every
Um intersects Um−1 and Um+1 , we at most have three non-zero terms in the
sum for each x ∈ Um , i.e.,
1
X
φ(x) = (ρδm+i ∗ φm+i )(x) x ∈ Um .
i=−1
4 1 1
The choice of this range for δ is motivated from the fact that m+1 − (m+1)(m+2) =
1
(m+2)
CHAPTER 2. SOBOLEV SPACES 93
Therefore,
X X
ku − φkk,p,Ω = k (uφm − ρδm ∗ φm u)kk,p,Ω ≤ kφm u − ρδm ∗ φm ukk,p,Ω < ε.
m m
Then for any fixed ε > 0 there exists no φ ∈ C 1 (Ω) such that ku − φk1,p < ε.
Example 2.8. Let Ω := {(r, θ) ∈ R2 | 1 < r2 < 2 and θ 6= 0}. Consider
u(r, θ) = θ. Then there exists no φ ∈ C 1 (Ω) such that ku − φk1,1 < 2π.
The trouble with domains in above examples is that they lie on both
sides of the boundary ∂Ω which becomes the main handicap while trying to
approximate W k,p (Ω) by C ∞ (Ω) functions.
CHAPTER 2. SOBOLEV SPACES 94
Then its distributional derivative is u0 (x) = 1(0,∞) . Let φ ∈ C ∞ (Ω) such that
kφ0 − u0 k∞ < ε. Thus, if x < 0, |φ0 (x)| < ε and if x > 0, |φ0 (x) − 1| < ε.
In particular, φ0 (x) > 1 − ε. By continuity, φ0 (0) < ε and φ0 (0) > 1 − ε
which is impossible if ε < 1/2. Hence, u cannot be approximated by smooth
functions in W 1,∞ (Ω) norm.
5
(k,p) polar sets
CHAPTER 2. SOBOLEV SPACES 95
and P u |Ω = u a.e. for every u ∈ W k,p (Ω). If P is same for all 1 ≤ p < ∞
and 0 ≤ m ≤ k, then P is called strong k-extension operator. If P is a strong
k-extension operator for all k then P is called total extension operator.
Example 2.10. There is a natural extension operator P : W0k,p (Ω) → W k,p (Rn )
which is the extension by zero. Define
(
u(x) x ∈ Ω
P u := ũ =
0 x ∈ Rn \ Ω.
Thus, Dα ũ = D
g α u and therefore
k
X k
X k
X
α
kũkk,p,Rn = kD ũkp,Rn = kD
g α uk
p,Rn = kDα ukp,Ω = kukk,p,Ω .
|α|=0 |α|=0 |α|=0
CHAPTER 2. SOBOLEV SPACES 97
Hence, Z
α ∗
D u (φ) = − uDα ψ(x) dx, (2.5.1)
Rn
+
The RHS in above equation is same as the RHS obtained in (2.5.1). Hence,
α ∗
R α
we have D u (φ) = Rn D u(x)ψ(x) dx. By setting
+
(
Dα u(x0 , xn ) xn > 0
(Dα u)∗ (x0 , xn ) := α 0
D u(x , −xn ) xn < 0,
Hence, Z
en ∗
D u (φ) = − uDen ψ(x) dx, (2.5.2)
Rn
+
where ψ(x0 , xn ) = φ(x0 , xn )−φ(x0 , −xn ) when xn > 0. Note that ψ(x0 , 0) = 0.
Thus, by Mean value theorem, |ψ(x0 , xn )| ≤ C|xn |. As before, in general,
ψ∈ / D(Rn+ ), so ζm (xn )ψ(x) ∈ D(Rn+ ). One such choice of ζm is by choosing
ζm (t) = ζ(mt) where (
0 if t < 1/2
ζ(t) =
1 if t > 1.
Consider,
Z
en
D u(ζm ψ) = − u(x)Den (ζm (xn )ψ(x)) dx
Rn
+
Z Z
0
= − u(x)ζm (xn )ψ(x) dx − u(x)ζm (xn )Den ψ(x) dx.
Rn
+ Rn
+
0
R
Therefore, limm→∞ Rn
u(x)ζm (xn )ψ(x) dx = 0 and hence
+
Z Z
en
D u(x)ψ(x) dx = − u(x)Den ψ(x) dx.
Rn
+ Rn
+
The RHS in above equation is same as the RHS obtained in (2.5.2). Hence,
en ∗
R en
we have D u (φ) = Rn D u(x)ψ(x) dx. By setting
+
(
Den u(x0 , xn ) xn > 0
(Den u)] (x0 , xn ) :=
−Den u(x0 , −xn ) xn < 0,
Note that the dual space is considered for W0k,p (Ω) and not for W k,p (Ω). The
reason is that D(Ω) is dense in W0k,p (Ω) and hence W0k,p (Ω) will have a unique
continuous extension (by Hahn-Banach) for any continuous linear functional
defined on D(Ω).
Example 2.11. In general the dual of W k,p (Ω) may not even be a distribution.
Note that, in general, D(Ω) is not dense in W k,p (Ω). Thus, its dual [W k,p (Ω)]∗
is not in the space of distributions D0 (Ω). Of course, the restriction to D(Ω)
of every T ∈ [W k,p (Ω)]∗ is a distribution but this restriction may not identify
with T . For instance, consider f ∈ [L2 (Ω)]n with |f | ≥ c > 0 a.e. and
div(f ) = 0. Define Z
T (φ) := f · ∇φ dx.
Ω
Since |T (φ)| ≤ kf k2 k∇φk2 , we infer that T ∈ [H 1 (Ω)]∗ . However, the restric-
tion of T to D(Ω) is the zero operator of D0 (Ω) because
hT, φi = −hdiv(f ), φi = 0 ∀φ ∈ D(Ω).
Theorem 2.6.1 (Characterisation of X1,p (Ω)). Let 1 ≤ p < ∞ and let
F ∈ X1,p (Ω). Then there exist functions f0 , f1 , . . . , fn ∈ Lq (Ω) such that
Z n Z
X ∂u
F (u) = f0 u dx + fi ∀u ∈ W01,p (Ω)
Ω i=1 Ω
∂x i
and kF kX1,p (Ω) = max0≤i≤n kfi kq . Further, if Ω is bounded, one may assume
f0 = 0.
Proof. Recall that (cf. proof of Theorem 2.2.2) T : W01,p (Ω) → (Lp (Ω))n+1
defined as
∂u ∂u
T (u) = u, ,...,
∂x1 ∂xn
CHAPTER 2. SOBOLEV SPACES 102
p 1/p
Pn+1
is an isometry. The norm in (Lp (Ω))n+1 is defined as kuk := i=1 kui kp ,
where u = (u1 , u2 , . . . , un+1 ). Let E = T (W0 (Ω)) ⊂ (L (Ω)) . Observe
1,p p n+1
Now, since D(Ω) is dense in W01,p (Ω), the extension of F to W01,p (Ω) should
be unique. Henceforth, we shall identify any element F ∈ X1,p (Ω) with the
distribution n
X ∂fi
f0 − .
i=1
∂x i
CHAPTER 2. SOBOLEV SPACES 103
The above remark motivates the right notation for the space Xk,p (Ω).
∂u
Observe that if u ∈ W k,p (Ω), then the first derivatives of ∂x i
, for all i, are
k−1,p
in W . To carry forward this feature in our notation, the above remark
motivates to rewrite Xk,p (Ω) as W −k,q (Ω), where q is the conjugate exponent
corresponding to p.
Let us observe that the representation of F in terms fi is not unique. Let
1,q
Pn ∂ ∂g
g ∈ W (Ω) such that ∆g = 0, i.e., i=1 ∂xi ∂xi = 0. Then
n ∂ f + ∂g
X i ∂xi
F = f0 −
i=1
∂xi
is also a representation of F .
Exercise 65. Show that
!1/q
n
X
kF kW −1,q (Ω) = inf kfi kqq .
P n ∂fi
F =f0 − i=1 ∂xi
i=0
with the obvious norm. For any positive real number s, set k := bsc, integral
part and σ := s is the fractional part. Note that 0 < σ < 1.
W s,p (Ω) = {u ∈ W k,p (Ω)|Dα u ∈ W σ,p (Ω) for all |α| = k}. (2.7.2)
0
We denote by W0s,p (Ω), the closure of D(Ω) in W s,p (Ω) and W −s,p (Ω) is dual
of W0s,p (Ω).
We begin by giving a characterisation of the space H 1 (Rn ) in terms of
Fourier transform. Recall that for any u ∈ L1 (Rn ) its Fourier transform û is
defined as Z
û(ξ) = e−ı2πξ·x u(x) dx. (2.7.3)
Rn
CHAPTER 2. SOBOLEV SPACES 104
Recall that û ∈ C0 (Rn ) and if û ∈ L1 (Rn ), one can invert the Fourier trans-
form to obtain u from û by the following formula:
Z
u(x) = eıξ·x û(ξ) dξ. (2.7.4)
Rn
∂u
dm
(ξ) = 2πıξk ûm (ξ). (2.7.7)
∂xk
∂um ∂u
Since um → u in the H 1 (Rn ) norm, um → u in L2 (Rn ) and ∂xk
→ ∂xk
in
L2 (Rn ), we can extract a subsequence um` such that
and
∂u
d m` ∂u
d
(ξ) → (ξ) for a.e. ξ ∈ Rn .
∂xk ∂xk
CHAPTER 2. SOBOLEV SPACES 105
∂u
d
(ξ) = 2πıξk û(ξ) for a.e. ξ ∈ Rn . (2.7.8)
∂xk
∂u
Now, if u ∈ H 1 (Rn ) then ∂x k
∈ L2 (Rn ), for all k = 1, 2, . . . , n. Thus, the
∂u
Fourier transform of ∂x k
is well-defined and
∂u
d
(ξ) = 2πıξk û(ξ).
∂xk
Proof. Note that |ξ|2 = ξ12 + . . . + ξn2 and |ξ α | = |ξ1 |α1 . . . |ξn |αn . By induction
argument on k, we can see that same powers of ξ occur in (1 + |ξ|2 )k and
α 2
P
|α|≤k |ξ | , albeit with different coefficients, which depend only on n and
k. Since the number of terms is finite and depends again only on n and k,
the result follows.
CHAPTER 2. SOBOLEV SPACES 106
Owing to the above lemma one can define the space H k (Rn ) as follows:
k
H k (Rn ) = {u ∈ L2 (Rn )|(1 + |ξ|2 ) 2 û(ξ) ∈ L2 (Rn )}
and Z 21
2 k 2
kukH k (Rn ) = (1 + |ξ| ) |û(ξ)| dξ .
Rn
is finite, for any s ∈ R, since as makes sense for any s ∈ R when a > 0.
In this case, a = 1 + |ξ|2 is positive. Thus, we are motivated to give the
following definition.
For s > 0, we define H −s (Rn ) as the dual of H s (Rn ). The negative order
Sobolev spaces has the following characterization:
Theorem 2.7.5. Let s ∈ (0, ∞). Then
s
H −s (Rn ) = {u ∈ S 0 (Rn ) | (1 + |ξ|2 )− 2 û(ξ) ∈ L2 (Rn )} (2.7.9)
Proof. We shall give the proof for s = 1. If u ∈ H −1 (Rn ) then
n
X ∂fi
u = f0 + , f0 , f1 . . . , fn ∈ L2 (Rn ).
i=1
∂xi
Hence, u is a tempered distribution and
n
X
û = fˆ0 + (2πı)ξi f̂i .
i=1
1
Then (1 + |ξ|2 )− 2 û ∈ L2 (Rn ), proving one inclusion in (2.7.9). To prove the
1
reverse inclusion, consider u ∈ S 0 (Rn ) such that (1 + |ξ|2 )− 2 û(ξ) ∈ L2 (Rn ).
Let φ ∈ D(Rn ). Then there exists ψ ∈ S(Rn ) such that φ = ψ̂. Set
1 1
k(ξ) := (1 + |ξ|2 ) 2 and k−1 (ξ) := (1 + |ξ|2 )− 2 .
Note that both k and k−1 are in C ∞ (Rn ), we write
u(φ) = u(ψ̂) = û(ψ) = (kk−1 )û(ψ) = k−1 û(kψ).
But k−1 û ∈ L2 (Rn ) and, hence,
Z
1 1
u(φ) = (1 + |ξ|2 )− 2 û(ξ)(1 + |ξ|2 ) 2 ψ(ξ) dξ.
Rn
Therefore,
1 1
|u(φ)| ≤ |(1 + |ξ|2 )− 2 û(ξ)|0,Rn |(1 + |ξ|2 ) 2 ψ(ξ)|0,Rn .
But
2
Z
1
2
(1 + |ξ| ) ψ(ξ) 2 = (1 + |ξ|2 )ψ 2 (ξ) dξ
0,Rn Rn
Z
= (1 + |ξ|2 )ψ 2 (−ξ) dξ
Z Rn
ˆ
= (1 + |ξ|2 )ψ̂ 2 (ξ) dξ
n
ZR
= (1 + |ξ|2 )(φ̂(ξ))2 dξ = kφk2H 1 (Rn ) .
Rn
CHAPTER 2. SOBOLEV SPACES 108
For an open subset Ω of Rn , we may define the Sobolev spaces H s (Ω), for
real s, as the restrictions to Ω of elements of H s (Rn ).
Example 2.12. If δ0 is the Dirac distribution, we know that δˆ0 ≡ 1 and, hence,
s
δ0 ∈ H −s (Rn ) if and only if (1 + |ξ|2 )− 2 ∈ L2 (Rn ). This is true for s > n2
since the integral in polar coordinates is
∞
rn−1
Z
dr
0 (1 + r2 )s
Recall that the Lp spaces are actually equivalence classes of functions with
equivalence relation being “equality almost everywhere”. This motivates the
following definition.
Proof. Since u ∈ W 1,p (a, b), u is weakly differentiable and u0 ∈ Lp (a, b). For
each x ∈ (a, b), we define v : (a, b) → R as
Z x
v(x) := u0 (t) dt.
a
1 λ
kukr ≤ C k∇ukp
λn/r λn/p
1+ n −n
kukr ≤ Cλ r p k∇ukp .
The above obtained inequality being true for λ > 0 would contradict (2.8.1)
except when 1 + nr − np = 0. Consequently, to expect an inequality of the
kind (2.8.1), we need to have 1r = p1 − n1 . Therefore, p1 − n1 > 0 and hence
np
1 ≤ p < n and r = n−p . This discussion motivates the definition of Sobolev
conjugate, p? , corresponding to the index p.
np
p? :=
n−p
1 1
Equivalently, p?
= p
− n1 . Also, p? > p.
Yn Z n0 /n
n
= |fi | dx1 . . . dxn
i=1 Rn−1
n
0
Y
= kfi knn,Rn−1
i=1
CHAPTER 2. SOBOLEV SPACES 112
Yn Z 1/n−1
= |fi | dx
i=1 Rn−1
Yn n
Y
1/n−1 1/n−1
= kfi k1,Rn−1 = kDei φk1,Rn .
i=1 i=1
Hence the result proved for p = 1. Let ψ := |φ|γ , where γ > 1 will be chosen
appropriately during the subsequent steps of the proof. Also, if φ ∈ D(Rn ),
then ψ ∈ Cc1 (Rn ). We shall apply the p = 1 result to ψ. Therefore,
Since we want only the gradient term on the RHS, we would like to bring the
nγ
q norm term to LHS. If we choose γ such that n−1 = (γ − 1)q, then we can
CHAPTER 2. SOBOLEV SPACES 114
Note that the second inequality in the statement of the above corollary
involves the W 1,p -norm of u and not the Lp -norm of the gradient of u. How-
ever, for bounded sets one can hope to get the inequality involving gradient
of u.
Proof. Let 1 ≤ p < n and u ∈ W01,p (Ω). Then, by previous corollary, there
is a constant C > 0 (depending on p and n) kukp? ≤ Ck∇ukp . For any
1 ≤ r ≤ p? , there is a constant C > 0 (depending on r and Ω) such that
kukr ≤ Ckukp? (since Ω is bounded). Therefore, we have a constant C > 0
(depending on p, n, r and Ω) kukr ≤ Ck∇ukp for all r ∈ [1, p? ].
Proof. Let a > 0 and suppose Ω = (−a, a)n . Let u ∈ D(Ω) and x = (x0 , xn ) ∈
Rn . Then
Z xn
∂u 0
u(x) = (x , t) dt.
−a ∂xn
CHAPTER 2. SOBOLEV SPACES 117
∂u
kukp,Ω ≤ 2a , ∀u ∈ D(Ω).
∂xn p,Ω
Thus,
∂u
kukp,Ω ≤ 2a ≤ 2ak∇ukp,Ω , ∀u ∈ D(Ω).
∂xn p,Ω
By the density of D(Ω) in W01,p (Ω), we get the result for all u ∈ W01,p (Ω).
Now, suppose Ω is not of the form (−a, a), then Ω ⊂ (−a, a) for some
a > 0, since Ω is bounded. Then any u ∈ W01,p (Ω) can be extended to
W01,p (−a, a) and use the result proved above.
Remark 2.8.11. Poincaré inequality is not true for u ∈ W 1,p (Ω). For in-
stance, if u ≡ c, a constant, then ∇u = 0 and hence k∇ukp = 0 while
kukp > 0. However, if u ∈ W 1,p (Ω) such that u = 0 on Γ ⊂ ∂Ω, then
Poincaré inequality remains valid for such u’s.
We now extend the results to proper subsets of Rn . The proofs are similar
to the equivalent statements from previous section.
where γ := 1 − n/p.
Proof. Let u ∈ D(Rn ) and let E be a cube of side a containing the origin
and each of its sides being parallel to the coordinate axes of Rn . Let x ∈ E.
We have
Z 1 Z 1
d
|u(x) − u(0)| = (u(tx)) dt = ∇u(tx) · x dt
0 dt 0
Z 1 Z 1 n
X ∂
≤ |∇u(tx)| · |x| dt = |xi | u(tx) dt
0 0 i=1
∂xi
Z 1 n Z 1 n
X ∂ X ∂
≤ a u(tx) dt = a u(tx) dt
0 i=1
∂xi 0 i=1
∂xi
Consider,
Z Z
1 1
|u − u(0)| = u(x) dx − u(0) ≤ n |u(x) − u(0)| dx
an E a E
Z Z 1X n
a ∂
≤ u(tx) dt dx
an E 0 i=1 ∂xi
n Z
1 X 1
Z
∂
= n−1
u(tx) dx dt (Fubini’s Theorem)
a i=1 0 E ∂xi
n Z
1 X 1 −n
Z
∂
= n−1
t u(y) dy dt (Change of variable)
a i=1 0 tE ∂xi
n Z
1 X 1 −n ∂u
≤ t (|tE|)1/q dt
an−1 i=1 0 ∂xi p,E
(By Hölder inequality and tE ⊂ E for 0 ≤ t ≤ 1)
n/q Z 1
a
= n−1 k∇ukp,E t−n tn/q dt (q is conjugate exponent of p)
a 0
Z 1
a1−n/p
= a1−n/p k∇ukp,E t−n/p dt = k∇ukp,E .
0 1 − n/p
The above inequality is then true for any cube E of side length a with sides
parallel to axes, by translating it in Rn . Therefore, for any cube E of side a
and x ∈ E, we have
aγ
|u − u(x)| ≤ k∇ukp,E (2.8.5)
γ
where γ := 1 − n/p. Consider,
1
|u(x)| = |u(x) − u + u| ≤ k∇ukp,E + kuk1,E
γ
1
≤ k∇ukp,E + kukp,E (by Hölder’s inequality)
γ
where we have used (2.8.5), in particular, for a unit cube E. Then
1
kuk∞,Rn ≤ k∇ukp,Rn + kukp,Rn .
γ
Therefore,
kuk∞,Rn ≤ Ckuk1,p,Rn , (2.8.6)
CHAPTER 2. SOBOLEV SPACES 121
2γ+1 γ 2γ+1
|u(x) − u(y)| ≤ |x − y| k∇ukp,E ≤ |x − y|γ k∇ukp,Rn .
γ γ
Thus, u is Hölder continuous and its Hölder seminorm pγ (.) (cf.(2.1.2)) is
bounded as below,
2γ+1
pγ (u) ≤ k∇ukp,Rn .
γ
and this together with (2.8.6) gives the bound for the γ-th Hölder norm,
Remark 2.8.17. As usual, the results can be extended to W 1,p (Ω) for Ω
bounded with C 1 smooth boundary and to W01,p (Ω) for any open subset Ω.
Example 2.14. Let p < n. Consider the function |x|δ , for any choice of δ in
1 − np < δ < 0, is in W 1,p (Ω) which has no continuous representative.
Example 2.15. Let p = n and n ≥ 2. We shall give an example of a function in
W 1,n (Ω) which has no continuous representative. We shall given an example
for the case n = 2. Let Ω := {x ∈ Rn : |x| < R} and u(x) = (− ln |x|)δ for
x 6= 0. We have, using polar coordinates,
Z Z R
n n−1
u dx = R ωn (− ln r)nδ r dr.
BR (0) 0
CHAPTER 2. SOBOLEV SPACES 122
and, therefore,
|∇u| = δ(− ln |x|)δ−1 |x|−1 .
Thus, using polar coordinates, we get
Z Z R
n n−1 n
|∇u| dx = R ωn |δ| | ln r|nδ−n r1−n dr.
BR (0) 0
1. If p < n/k, then W k,p (Rn ) ,→ Lr (Rn ) for all r ∈ [p, np/(n − pk)].
3. If p > n/k, then W k,p (Rn ) ,→ L∞ (Rn ) and further there is a represen-
tative of u, say u? , whose k-th partial derivative is Hölder continuous
with exponent γ and there is a constant C > 0 (depending only on p, k
and n) such that
where γ := k − n/p − [k − n/p] and [l] is the largest integer such that
[l] ≤ l.
Proof. 1. If k = 1 we know the above results are true. Let k = 2 and
p < n/2. Then, for any u ∈ W 2,p (Rn ), both u, D1 u ∈ W 1,p (Rn ) and
since p < n/2 < n, using Sobolev inequality, we have both u, D1 u ∈
? ?
Lp (Rn ) with p? = np/(n − p). Thus, u ∈ W 1,p (Rn ). Now, since
p < n/2, we have p? < n. Thus, using Sobolev inequality again again,
? ?
we get u ∈ L(p ) (Rn ). But 1/(p? )? = 1/p − 2/n. Extending similar
arguments for each case, we get the result. Note that when we say D1 u,
we actually mean Dα u for each |α| = 1. Such convention will be used
throughout this proof.
Remark 2.8.20. As usual, the results can be extended to W 1,p (Ω) for Ω
bounded with C k smooth boundary and to W01,p (Ω) for any open subset Ω.
CHAPTER 2. SOBOLEV SPACES 124
kgi − gj krr,R = kgi krr,Ii + kgj krr,Ij + kgi − gj krr,R\(Ii ∪Ij ) = 2cr .
Thus, the sequence is not Cauchy in Lr (R) (as seen by choosing ε < 21/r c
for all i, j). The arguments can be generalised to Rn .
Proof. It is enough to prove the inequality for u ∈ D(Rn ) due to the density
of D(Rn ) in W 1,p (Rn ). Consider
Proof. (i) We first prove the case p < n. Let B be the unit ball in W 1,p (Ω).
We shall verify conditions (i) and (ii) of Theorem 1.3.31. Let 1 ≤ q < p∗ .
Then choose α such that 0 < α ≤ 1 and
1 α 1−α
= + .
q 1 p∗
We choose h small enough such that C|h|α < ε. This will verify condi-
tion (i) of Theorem 1.3.31. Now, if u ∈ B and Ω0 ⊂⊂ Ω, it follows by
Hölder’s inequality that
0 ?)
kukq,Ω\Ω0 ≤ kukp? ,Ω\Ω0 |Ω \ Ω |1−(q/p
0 ?)
≤ C|Ω \ Ω |1−(q/p
(ii) Assume for the moment that the result is true for p < n. Notice that
as p → n, p? → ∞. Since Ω is bounded, W 1,n (Ω) ⊂ W 1,n−ε (Ω), for
every ε > 0. Since n − < n, using the p < n case, we get W 1,n−ε (Ω)
is compactly imbedded in Lr (Ω) for all r ∈ [1, (n − ε)? ). Note that as
ε → 0, (n − ε)? → ∞. Therefore, for any r < ∞ we can find small
enough ε > 0 such that 1 ≤ r < (n − )∗ . We deduce that W 1,n (Ω) is
compactly imbedded in Lr (Ω) for any 1 ≤ r < ∞.
(iii) For p > n, the functions of W 1,p (Ω) are Hölder continuous. If B is the
unit ball in W 1,p (Ω) then the functions in B are uniformly bounded and
equicontinuous in C(Ω̄). Thus B is relatively compact in C(Ω̄) by the
Ascoli-Arzela Theorem.
Remark 2.8.23. Note that the continuous inclusion for the r = p? case is
not compact. The above result can be extended to W01,p (Ω) provided Ω is
bounded and is a connected open subset (bounded domain) of Rn .
(i) A is bounded in W 1,p (Rn ), i.e., supf ∈A kf kW 1,p (Rn ) < +∞.
R
(ii) A is Lp -equi-integrable at infinity, i.e., limr→+∞ {|x|>r} |f (x)|p dx = 0
uniformly with respect to f ∈ A. Then, A is relatively compact in
Lp (Rn ).
Proof. Suppose the inequality was false, then for each positive integer m we
have a um ∈ W 1,p (Ω) such that
r
converges in R(Ω), for some measure µ ∈ R(Ω). Set s = r−1 and denote by
1,s n
B the closed unit ball in W0 (Ω). Since 1 ≤ r < n−1 , we have s > n and so
N (ε)
B is compact in C0 (Ω). Thus, given ε > 0 there exist {φi }i=1 ⊂ C0 (Ω) such
that
min kφ − φi kC(Ω) < ε
1≤i≤N (ε)
]
Since the space C ∞ (Ω) is dense in W 1,p (Ω) (cf. Corollary 2.5.6) and Lp (∂Ω)
is complete, there exists a unique continuous linear extension of (cf. The-
orem ??) γ from C ∞ (Ω) to W 1,p (Ω). This extension, still denoted as γ, is
]
called the trace operator and each γ(u) ∈ Lp (∂Ω) is called the trace of the
function u ∈ W 1,p (Ω).
Further, if 1 < p < n then the trace operator γ : W 1,p (Ω) → Lq (∂Ω) is
compact for all q ∈ [1, p] ).
1 1
ab ≤ ap + bq
p q
CHAPTER 2. SOBOLEV SPACES 130
1 1
with p
+ q
= 1 to the following situation:
∂u 0
a= (x , xn ) and b = |u(x0 , xn )|p−1 .
∂xn
We obtain
Z +∞ p
0 1 ∂u 0 1
|u(x , 0)| ≤ p p
(x , xn ) + |u(x0 , xn )|(p−1)q dxn .
0 p ∂xn q
Hence
1
ku(·, 0)kLp (Rn−1 ) ≤ p p kukW 1,p (Rn+ ) .
Proof. By the regularity property of the boundary ∂Ω, we know that D(Ω̄)
is dense in W 1,p (Ω). For any v ∈ D(Ω̄), we can define the restriction of u to
∂Ω, setting γ(u) := u|∂Ω . Suppose we prove that
Thus, we just need to prove that γ is continuous. By Lemma 2.9.2, the result
is true for half-space Ω = Rn+ . Let Ω̄ ⊂ ki=0 Gi with Ḡ0 ⊂ Ω, Gi open for all
S
i = 0, . . . , k, while φi : B(0, 1) → Gi , i = 1, 2, . . . , k, are the local coordinates
{α
Pk0 , . . . , αi }, is an associated partition of unity, i.e., αi ∈ D(Gi ), αi ≤ 0,
i=0 αi = 1 on Ω̄.
(
(αi v) ◦ φi on B+ ,
wi =
0 on Rn+ \B+ .
By using classical differential calculus rules (note that all the functions αi , v, φi
are continuously differentiable), one obtains the existence, for any i = 1, . . . , k,
of a constant Ci such that
We now use the definition of the Lp (∂Ω) norm which is based on the use of
local coordinates. One can show that an equivalent norm to the Lp (∂Ω) norm
can be obtained by using local coordinates: denoting by ∼ the extension by
zero outside of Rn−1 \{y ∈ Rn−1 | |y| < 1}, we have that
Lp (∂Ω) = {u : ∂Ω → R | (α
] p
i u) ◦ φi (·, 0) ∈ L (R
n−1
), 1 ≤ i ≤ k}
CHAPTER 2. SOBOLEV SPACES 132
and
k
! p1
X
u 7→ αi u) ◦ φi kpLp (Rn−1 )
k(g (2.9.4)
i=1
1
Second step: To show that v(x0 , 0) = w(x0 ) ∈ H 2 (Rn−1 ) we need to show
1
e 0 )|2 is integrable. But
that (1 + |ξ 0 |2 ) 2 |w(ξ
Z
1
(1 + |ξ 0 |2 ) 2 |w(ξ
e 0 )|2 dξ 0
Rn−1
Z Z ∞ 2
0 2 12
= (1 + |ξ | ) v̂(ξ) dξn dξ 0
Rn−1 −∞
Z Z ∞ 2
1 −1 1
= (1 + |ξ 0 |2 ) 2 (1 + |ξ|2 ) 2 dξn dξ 0
v̂(ξ)(1 + |ξ|2 ) 2
n−1
ZR Z−∞ ∞ Z ∞
0 2 21
≤ (1 + |ξ | ) 2 2
(1 + |ξ| )|v̂(ξ)| dξn 2 −1
(1 + |ξ| ) dξn dξ 0
Rn−1 −∞ −∞
Z
= π (1 + |ξ|2 )|v̂(ξ)|2 dξ = πkvk2H 1 (Rn )
Rn
since
Z ∞ Z ∞
2 −1 dξn 1
(1 + |ξ| ) dξn = 0 2 2
= π(1 + |ξ 0 |2 )− 2 .
−∞ −∞ 1 + |ξ | + ξn
π/2 π/2
sec2 θ
Z Z
1 −1 −1
1 2
dθ = (1 + |ξ 0 |2 ) 2 dθ = π(1 + |ξ 0 |2 ) 2 .
(1 + |ξ 0 |2 ) 2 −π/2 1 + tan θ −π/2
1
Thus, if v ∈ D(Rn ), v(x0 , 0) ∈ H 2 (Rn−1 ) and by density the result follows
for v ∈ H 1 (Rn+ ).
1 1
Third step: We now show that γ is onto H 2 (Rn−1 ). Let h(x0 ) ∈ H 2 (Rn−1 ).
h(ξ 0 ) be its Fourier transform. We define u(x0 , xn ) by
Let e
0
e(ξ 0 , xn ) = e−(1+|ξ |)xn e
u h(ξ 0 ). (2.9.5)
We must first show that u then belongs to H 1 (Rn+ ). To see this extend u by
CHAPTER 2. SOBOLEV SPACES 134
−(1 + |ξ 0 |)e
h (ξ 0 )
\
∂u
(ξ) = .
∂xn 1 + |ξ 0 | + 2πıξn
Then
2
(1 + |ξ 0 |2 )|e
h (ξ 0 )|2
Z \ Z
∂u
(ξ) dξ ≤ 2 dξ
Rn ∂xn Rn 1 + |ξ 0 |2 + ξn2
Z
1
= 2π (1 + |ξ 0 |2 ) 2 |e
h (ξ 0 )|2 dξ 0 < ∞.
Rn
CHAPTER 2. SOBOLEV SPACES 135
∂u ∂u
Thus ∂x n
(extended by zero) is in L2 (Rn ) and so ∂x n
∈ L2 (Rn+ ) and hence
u ∈ H 1 (Rn+ ). Now (by the Fourier inversion formula)
e(ξ 0 , 0) = e
u h(ξ 0 )
k
Definition 2.9.7. We shall denote the range of the map T to be W 2 ,p (∂Ω).
k k
For p = 2, we denote W 2 ,2 as H 2 (∂Ω).
Thus, instead of defining the fractional power Sobolev spaces using Fourier
transform, one can define them as range of trace operator γ.
k
Theorem 2.9.8. W 2 ,p (∂Ω) is dense in Lp (∂Ω).
Proof. We first show the inclusion W01,p (Ω) ⊂ kerγ. Take v ∈ W01,p (Ω). By
definition of W01,p (Ω), there exists a sequence of functions (vn )n∈N , vn ∈ D(Ω)
such that vn → v in W 1,p (Ω). Since γ0 (vn ) = vn |∂Ω = 0, by continuity of γ0
we obtain that γ0 (v) = 0, i.e., v ∈ kerγ0 .
The other inclusion is a bit more involved. We shall prove it in a sequence
of lemma for H 1 (Rn+ ). The idea involved is following: Take v ∈ W 1,p (Rn+ )
such that γ0 (v) = 0. Prove that v ∈ W01,p (Rn+ ). Let us first extend v by zero
outside of Rn+ . By using the information γ0 (v) = 0 one can verify that the
so-obtained extension ṽ belongs to W 1,p (Rn ). Then, let us translate ṽ, and
consider for any h > 0
Finally, one regularizes by convolution the function τh ṽ. We have that for ε
sufficiently small, ρε ? (τh ṽ) belongs to D(Rn+ ) and ρε ? (τh ṽ) tends to v in
W 1,p (Rn+ ) as h → 0 and ε → 0. Hence v ∈ W01,p (Rn+ ).
Z Z Z
∂v ∂u
u =− v− γ0 (u)γ0 (v). (2.9.7)
Rn
+
∂x n Rn
+
∂xn Rn−1
by the above corollary. Let h > 0 and consider h̄ = hen ∈ Rn where en is the
unit vector (0, 0, ..., 0, 1). Consider the function τh̄ ve, where ve is the extension
by zero outside Rn+ of v ∈ ker(γ0 ). Then τh̄ ve vanishes for all x ∈ Rn such
that xn < h. (Recall that τh̄ ve(x) = ve(x − h̄).)
Lemma 2.9.13. Let 1 ≤ p < ∞ and h̄ ∈ Rn . Then if f ∈ Lp (Rn )
ke
v − ζk vek1,Rn < η,
where η > 0 is a given positive quantity. Again we can choose h small enough
so that if h̄ = hen , then
Now τh̄ (ζk ve) has compact support in Rn+ and vanishes for all x ∈ Rn with
xn < h. Let {ρ } be the family of mollifers. If > 0 is chosen small enough
then ρ ∗ τh̄ (ζk ve) will have support contained in the set
and we know that as ↓ 0, ρ∗τh̄ (ζk ve) → τh̄ (ζk ve). Thus we can choose small
enough such that
kρ ∗ τh̄ (ζk ve) − τh̄ (ζk ve)k1,Rn < η.
Thus we have found a function φη ∈ D(Rn+ ),
such that
kφη − vk1,Rn+ ≤ kφη − vek1,Rn < 3η .
Thus, as η is arbitrary, it follows that
where (νi )ni=1 denotes the unit outer normal vector field along ∂Ω.
Proof. Let us now turn to the case of a bounded open set Ω of class C 1 . Let
{Uj , Tj }kj=1 be an associated local chart for the boundary ∂Ω and let {ψj }kj=1
be a partition of unity subordinate to the cover {Uj } of ∂Ω. If u ∈ H 1 (Ω),
then (ψj u|Uj∩ Ω)o Tj ∈ H 1 (Rn+ ) and so we can define its trace asTan element of
H 21 (Rn−1 ). Coming back by Tj−1 we can define the trace on Uj ∂Ω. Piecing
these together we get the trace γ0 u in L2 (∂Ω) and the image (by definition of
1
the spaces) will be precisely H 2 (∂Ω). Similarly if the boundary is smoother
we can define the higher order traces γj .
Now that we have given a meaning to the functions restricted to the
boundary of the domain, we intend to generalise the classical Green’s identi-
ties (cf. Appendix ??) to H 1 (Ω). The trace theorem above helps us to obtain
Green’s theorem for functions in H 1 (Ω), Ω of class C 1 . If ν(x) denotes the
unit outer-normal vector on the boundary ∂Ω (which is defined uniquely a.e.
on ∂Ω), we denote its components along the coordinate axes by νi (x). Thus
we write generically,
ν = (ν1 , · · · , νn ).
For example, if Ω = B(0; 1) then ν(x) = x for all |x| = 1. Thus νi (x) = xi
in this case. If Ω = B(0; R), then ν(x) = Rx . If Ω has a part of its boundary,
say, xn = 0, then the unit outer normal is ±en depending on the side on
which Ω lies.
Theorem 2.9.18 (Green’s Theorem). . Let Ω be a bounded open set of Rn
set of class C 1 lying on the same side of its boundary ∂Ω. Let u, v ∈ H 1 (Ω).
Then for 1 ≤ i ≤ n,
Z Z Z
∂v ∂u
u =− v+ (γ0 u) (γ0 v) νi . (2.9.11)
Ω ∂xi Ω ∂xi ∂Ω
Remark 2.9.21. The Green’s formula holds for u ∈ W 1,p (Ω) and a ∈
W 1,q (Ω) with p1 + 1q = 1
CHAPTER 2. SOBOLEV SPACES 142
n
!1/2
X
kΦk∞ = sup Φ2i (x) .
x∈Ω
i=1
143
CHAPTER 3. BOUNDED VARIATION FUNCTIONS 144
1. uk → u in L1 (Ω) and
2. V (uk , Ω) → V (u, Ω) as k → ∞
Theorem 3.0.6. Let Ω be an open subset of Rn . For any function u ∈
Cc∞ (Ω), the total variation of u satisfies the identity V (u, Ω) = k∇uk1,Ω .
1,1
Corollary 3.0.7. Let u ∈ Wloc (Ω). Then ∇u ∈ L1 (Ω; Rn ) iff V (u, Ω) < ∞.
Further, in that case, k∇uk1,Ω = V (u, Ω).
Theorem 3.0.8 (Sobolev inequality for BV). There exists a constant C > 0
such that
n
kuk n−1
n
,Rn ≤ CV (u, R ) ∀u ∈ BV(Rn ).
1,1
Thus, kuk n−1
n
,Rn ≤ lim inf k→∞ kuk k n−1
n
,Rn . Since uk ∈ W (Rn ), applying
GNS inequality for p = 1, we obtain
kf k n−1
n
,Rn ≤ lim Ck∇uk k1,Rn .
k→∞
147
CHAPTER 4. SECOND ORDER ELLIPTIC PDE 148
as long as the integrals make sense. Note that, in the above problem, it
is enough for u to be in C 1 (Ω). Also, in particular, any classical solution
satisfies the identity Z Z
|∇u|2 dx = f u dx.
Ω Ω
Since C 1 (Ω) which vanishes on boundary is dense in H01 (Ω), we have the
classical solution u will also solve
Z Z
∇u · ∇v dx = f v dx ∀v ∈ H01 (Ω),
Ω Ω
as long as the integrals make sense. The integrals will make sense if f ∈
L2 (Ω). In fact, more generally, if f ∈ H −1 (Ω) then the integral on the
right side can be replaced with the duality product. The arguments above
motivates the notion of weak solution
in H01 (Ω).
CHAPTER 4. SECOND ORDER ELLIPTIC PDE 149
Proof. Recall that H 1 (Ω) is a Hilbert space endowed with the inner-product
Z Z
(v, w) := vw dx + ∇v∇w dx.
Ω Ω
Since Ω is bounded, by Poincaré inequality (cf. (2.8.4)), there is a constant
C > 0 such that
kvk2 ≤ Ck∇vk2 ∀v ∈ H01 (Ω).
Thus, H01 (Ω) is a Hilbert space endowed with the equivalent norm kvkH01 (Ω) :=
k∇vk2 and the corresponding inner product is given as
Z
((v, w)) := ∇v · ∇w dx.
Ω
We shall use the Lax-Milgram result (cf. Theorem ??). We define the map
a : H01 (Ω) × H01 (Ω) :→ R as
Z
a(v, w) := ((v, w)) = ∇v · ∇w dx.
Ω
Note that a is a bilinear. We need to show that a is continuous and coercive.
Consider,
Z
|a(v, w)| ≤ |∇v||∇w| dx
Ω
Z 1/2 Z 1/2
2 2
≤ |∇v| dx |∇w| dx
Ω Ω
= kvkH01 (Ω) kwkH01 (Ω) .
Thus, a is a bilinear continuous form. Now, consider
Z
a(v, v) = |∇v|2 dx = k∇vk22 .
Ω
Hence a is coercive, bilinear continuous form. By Lax-milgram theorem,
there is a u ∈ H01 (Ω) such that a(u, v) = hf, viH −1 (Ω),H01 (Ω) . This is equivalent
to (4.1.2) and also u minimizes the functional
Z
1
= |∇v|2 dx − hf, viH −1 (Ω),H01 (Ω) .
2 Ω
Proof. For any given g ∈ H 1/2 (Ω), there is a vg ∈ H 1 (Ω). Note that the
choice of vg is not unique. We have already seen (cf. Theorem 4.1.3) the
existence of weak solution for the homogeneous Dirichlet problem. Thus,
there is a unique w ∈ H01 (Ω) satisfying (4.2.2). Then u := w + vg is a weak
solution of (4.2.1). However, the uniqueness of u does not follow from the
uniqueness of w, since the choice of vg is not unique. If u1 , u2 ∈ Kg are two
weak solutions of (4.2.1) then, by (4.2.3),
Z
∇(u1 − u2 ) · ∇v dx = 0 ∀v ∈ H01 (Ω).
Ω
Equivalently,
(Hence the name divergence form). On the other hand, a second order op-
erator L is said to be in non-divergence form, if L acting some u has the
form
n
X ∂ 2u
Lu := − aij (x) + b(x) · ∇u + c(x)u.
i,j=1
∂x i ∂x j
Observe that the operator L will make sense in the divergence form only
if aij (x) ∈ C 1 (Ω). Thus, if aij (x) ∈ C 1 , then a divergence form equation can
be rewritten in to a non-divergence form because
n n
!
X ∂ 2u X ∂aij
∇ · (A(x)∇u) = aij (x) + · ∇u.
i,j=1
∂x i ∂x j j=1
∂x j
i
∂a
Now, by setting ebi (x) = bi (x) − nj=1 ∂xijj , we have written a divergence L in
P
non-divergence form. Also, note that for A(x) = I, b(x) = 0 and c(x) = 0,
L = −∆ is the usual Laplacian.
We shall now extend the notion of weak solution introduced for the Lapla-
cian operator to a general second order elliptic equation in divergence form.
We remark that for the integrals in the defintion of weak solution to make
sense, the minimum hypotheses on A(x), b and c is that aij , bi , c ∈ L∞ (Ω).
CHAPTER 4. SECOND ORDER ELLIPTIC PDE 153
whenever
Z Z Z
A∇u·∇v dx+ (b·∇u)v dx+ cuv dx = hf, viH −1 (Ω),H 1 (Ω) , ∀v ∈ H01 (Ω).
0
Ω Ω Ω
(4.3.2)
Lemma 4.3.4. If aij , bi , c ∈ L∞ (Ω) then the bilinear map a(·, ·) is continuous
on H01 (Ω) × H01 (Ω), i.e., there is a constant c1 > 0 such that
Proof. Consider,
Z Z
|a(v, w)| ≤ |A(x)∇v(x) · ∇w(x)| dx + |(b(x) · ∇v(x))w(x)| dx
ΩZ Ω
+ |c(x)v(x)w(x)| dx
Ω
≤ max kaij k∞ k∇vk2 k∇wk2 + max kbi k∞ k∇vk2 kwk2 + kck∞ kvk2 kwk2
i,j i
≤ mk∇vk2 (k∇wk2 + kwk2 ) + mkvk2 kwk2 ,
where m = max max kaij k∞ , max kbi k∞ , kck∞
i,j i
in H01 (Ω).
Proof. It is already noted in Lemma 4.3.4 that one can find a c3 > 0 such
that a(v, v) + c3 kvk22 is coercive in H01 (Ω). Thus, by Theorem 4.3.5, there is
a unique u ∈ H01 (Ω) such that
Z Z
a(u, v) + c3 uv dx = f v dx ∀v ∈ H01 (Ω).
Ω Ω
(iii) If c ≡ 0 and f ≡ 0 then inf y∈∂Ω u(y) ≤ u(x) ≤ supy∈∂Ω u(y) for all
x ∈ Ω.
Proof. Recall that if u ∈ H 1 (Ω) then |u|, u+ and u− are also in H 1 (Ω).
(i) If u ≥ 0 on ∂Ω then u = |u| on ∂Ω. Hence, u− ∈ H01 (Ω). Thus, using
v = u− in (4.3.2), we get
Z Z Z
− − − 2
− A∇u · ∇u dx − c(x)(u ) dx = f (x)u− (x) dx
Ω Ω Ω
Thus,
Z Z Z
A(x)∇(T f ) · ∇v(x) dx + c(x)(T f )(x)v(x) dx = f (x)v(x) dx.
Ω Ω Ω
because range of T is H01 (Ω). Hence, φm satisfies (4.3.3). It now only remains
to show that φm ∈ C ∞ (Ω). For any x ∈ Ω, choose Br (x) ⊂ Ω. Since
φm ∈ L2 (Br (x)) and solves the eigen value problem, by interior regularity
(cf. Theorem 4.3.7), φm ∈ H 2 (Br (x)). Arguing similarly, we obtain φm ∈
H k (Br (x)) for all k. Thus, by Sobolev imbedding results, φm ∈ C ∞ (Br (x)).
Since x ∈ Ω is arbitrary, φm ∈ C ∞ (Ω).
Remark 4.3.10. Observe that if H01 (Ω) is equipped with the inner product
R −1/2
Ω
∇u·∇v dx, then λm φm is an orthonormal basis for H01 (Ω) where (λm , φm )
is the eigen pair corresponding to A(x) = I and c ≡ 0. With the usual inner
product Z Z
uv dx + ∇u · ∇v dx
Ω Ω
CHAPTER 4. SECOND ORDER ELLIPTIC PDE 159
in H01 (Ω), (λm + 1)−1/2 φm forms an orthonormal basis of H01 (Ω). The set
{φm } is dense in H01 (Ω) w.r.t both the norms mentioned above. Suppose
f ∈ H01 (Ω) is such that hf, φmRi = 0 in H01 (Ω), for all m. Then, from the
eigenvalue problem, we get λm Ω φm f dx = 0. Since φm is a basis for L2 (Ω),
f = 0.
Definition 4.3.11. The Rayleigh quotient map R : H01 (Ω) \ {0} → [0, ∞)
is defined as
R R
Ω
A(x)∇v · ∇v dx + Ω c(x)v 2 (x) dx
R(v) = .
kvk22,Ω
Theorem 4.4.1. Let Y be unit open cell and let aij ∈ L∞ (Ω) such that the
matrix A(y) = (aij (y)) is elliptic with ellipticity constant α > 0. For any
f ∈ (Wper (Y ))? , there is a unique weak solution u ∈ Wper (Y ) satisfying
Z
A∇u · ∇v dx = hf, vi(Wper (Y ))? ,Wper (Y ) ∀v ∈ Wper (Y ).
Y
as long as the integrals make sense. Note that, in the above problem, it
is enough for u to be in C 1 (Ω). Also, in particular, any classical solution
satisfies the identity
Z Z Z Z
2
A∇u · ∇u dx + c(x)u dx = f u dx + gu dσ.
Ω Ω Ω ∂Ω
1 1
By the density of C (Ω) in H (Ω), the classical solution u will also solve
Z Z Z Z
A∇u · ∇v dx + c(x)uv dx = f v dx + gv dσ ∀v ∈ H 1 (Ω),
Ω Ω Ω ∂Ω
where v inside the boundary integral is in the trace sense, as long as the
integrals make sense. The integrals will make sense if f ∈ L2 (Ω) and g ∈
L2 (∂Ω). In fact, one can consider g ∈ H −1/2 (∂Ω), if we choose to replace
the boundary integral with the duality product h·, ·iH −1/2 (∂Ω),H 1/2 (∂Ω) . The
arguments above motivates the notion of weak solution for (4.5.1).
CHAPTER 4. SECOND ORDER ELLIPTIC PDE 162
in H 1 (Ω).
Proof. Recall that H 1 (Ω) is a Hilbert space endowed with the inner-product
Z Z
(v, w) := vw dx + ∇v∇w dx.
Ω Ω
We shall use the Lax-Milgram result (cf. Theorem ??). We define the map
a : H 1 (Ω) × H 1 (Ω) :→ R as
Z Z
a(v, w) := A∇v · ∇w dx + cvw dx.
Ω Ω
Note that a is a bilinear. We need to show that a is continuous and coercive.
Consider,
Z Z
|a(v, w)| ≤ max kaij k∞ |∇v||∇w| dx + kck∞ |vw| dx
i,j Ω Ω
≤ C (k∇vk2 k∇wk2 + kvk2 kwk2 )
(where C := max max kaij k∞ , kck∞ )
i,j
Hence a is coercive, bilinear continuous form. Also, note that the map
Z
v 7→ f v dx + hg, viH −1/2 (∂Ω),H 1/2 (∂Ω)
Ω
and the linearity and continuity of the trace map T . Thus, by Lax-milgram
theorem, there is a u ∈ H 1 (Ω) such that
Z
a(u, v) = f v dx + hg, viH −1/2 (∂Ω),H 1/2 (∂Ω) .
Ω
in H 1 (Ω).
The term coercive refers to the fact that both A and c are positive definite.
In contrast to the Dirichlet boundary condition, observe that the Neumann
boundary condition is not imposed a priori, but gets itself imposed naturally.
In other words, if u is a weak solution of (4.5.1) and is in H 2 (Ω), then for
any v ∈ H 1 (Ω),
Z Z Z Z
A∇u · ∇v dx + c(x)uv dx = f v dx + gv dσ
Ω
Z Z ZΩ ZΩ Z∂Ω
∂u
− ∇ · (A∇u)v dx + v dσ + c(x)uv dx = f v dx + gv dσ
Ω ∂Ω ∂νA Ω Ω ∂Ω
CHAPTER 4. SECOND ORDER ELLIPTIC PDE 164
implies that
Z
∂u
v −g dσ = 0 ∀v ∈ H 1 (Ω).
∂Ω ∂νA
But v |∂Ω ∈ H 1/2 (∂Ω), since v ∈ H 1 (Ω). Now, by the density of H 1/2 (∂Ω) in
∂u
L2 (∂Ω) (cf. Theorem 2.9.8), we have ∂ν A
= g in L2 (∂Ω). Thus, note that we
have obtained the Neumann boundary condition as a consequence of u being
a weak solution to the Neumann problem on the domain Ω. Thus, Dirichlet
conditions are called essential boundary condition and Neumann conditions
are called natural boundary conditions.
An astute reader should be wondering that if H01 (Ω) was the precise solu-
tion space for the Dirichlet problem (4.1.1), then the corresponding solution
space for the Neumann problem (4.5.1) should be the closure of
Proof. We shall use the Lax-Milgram result (cf. Theorem ??). We define the
map a : V × V :→ R as
Z
a(v, w) := A∇v · ∇w dx.
Ω
and the linearity and continuity of the trace map T . Thus, by Lax-milgram
theorem, there is a u ∈ V such that
Z
a(u, v) = f v dx + hg, viH −1/2 (∂Ω),H 1/2 (∂Ω) .
Ω
in V .
The term semi-coercive refers to the fact that A is positive definite and
c is identically zero.
We introduce a different approach to prove the existence of weak solution
for the semi-coercive Neumann problem (4.6.1). This approach is called
the Tikhonov Regularization Method. The basic idea is to obtain the semi-
coercive case as a limiting case of the coercive Neumann problem.
Hence u minimizes J in V .
We skip the proof of above theorem because the arguments are similar
to those already introduced while studying homogeneous Dirichlet problem.
The Poincaré inequality is valid in Vm (since u vanishes on section of the
boundary, i.e., Γ1 ). Thus, all earlier arguments can be carried over smoothly.
Similarly, one can also study the mixed Dirichlet-Neumann problem with
inhomogeneous condition on Γ1 , say h. In this case the arguments are similar
to the Inhomogeneous Dirichlet problem with Vm replaced with its translate
Vm + h.
Let c : ∂Ω → R+ be given. We consider the Robin problem, i.e., given
f : Ω → R, g : ∂Ω → R and aij ∈ L∞ (Ω), find u : Ω → R such that
−div(A(x)∇u(x)) = f (x) in Ω
(4.7.3)
c(x)u + A(x)∇u · ν = g on ∂Ω
CHAPTER 4. SECOND ORDER ELLIPTIC PDE 171
where ν = (ν1 , . . . , νn ) is the unit outward normal. Observe that the Robin
problem incorporates most of the problems introduced in this chapter. For
instance, if c ≡ 0, then we have the semi-coercive Neumann problem. If
c ≡ +∞, then (formally) we have the homogeneous Dirichlet problem. If
c ≡ +∞ on Γ1 ⊂ ∂Ω and c ≡ 0 in ∂Ω \ Γ1 , then we have the mixed Dirichlet-
Neumann problem.
Definition 4.7.3. Given f ∈ L2 (Ω) and g ∈ H −1/2 (∂Ω), we say u ∈ H 1 (Ω)
is a weak solution of (4.7.3) if
Z Z Z
A∇u·∇v dx+ cuv dσ = f v dx+hg, viH −1/2 (∂Ω),H 1/2 (∂Ω) ∀v ∈ H 1 (Ω).
Ω ∂Ω Ω
(4.7.4)
Theorem 4.7.4. Let Ω be a bounded, connected open subset of R with C 1 n
Proof. The only non-trivial part of the proof is to show the coercivity of the
bilinear map Z Z
a(v, w) := A∇v · ∇w dx + cvw dσ
Ω ∂Ω
in H 1 (Ω).
Proof. First prove J is coercive and continuous on W01,p (Ω), then use Theo-
rem ?? to show the existence of a minimizer u ∈ W01,p (Ω).
Secondly, show J is convex, in fact strictly convex on W01,p (Ω) and con-
sequently conclude the uniqueness of minimizer.
Lastly, show that J is Gâteaux differentiable and use (i) of Proposition ??
to show that the minimizer u is a weak solution of (4.8.1).
Note that the strain tensor is 3 × 3 symmetric matrix. Let σij (u) denote the
stress (measure of internal forces arising due to force f acting on Ω) tensor.
Then the constitutive law (Hooke’s law) states that
3
!
X
σij (u) = λ kk (u) δij + 2µij (u),
k=1
where λ ≥ 0 and µ > 0 are called Lame’s coefficients and δij is the Kronecker
delta defined as (
0 i 6= j
δij =
1 i = j.
Note that the stress tensor (σij ) is also a 3 × 3 symmetric matrix. Now, for
each i = 1, 2, 3, the displacement vector is given as a solution of the BVP
P
3 ∂
− j=1 ∂xj (σij (u)) = fi in Ω
P3
j=1 σij (u)νj = gi on Γ1 (4.10.1)
ui = 0 on Γ2
H 1 (Coercive) There is a constant α > 0 such that a(x, ξ).ξ ≥ α|ξ|p , for
every ξ ∈ Rn .
Let H(α, β, h) denote the set of all Carathéodory functions satisfying the
above three hypotheses. For every coercive (hypotheis H1) Carathéodory
function, a(x, 0) = 0 for almost every x ∈ Ω.
whenever Z Z
uf dx = v dµ, ∀f ∈ L∞ (Ω),
Ω Ω
where v solves
−div(At (x)∇v) = f in Ω
(4.12.2)
v =0 on ∂Ω
An equivalent way of stating the truncation at height k is Tk (s) = max(−k, min(k, s)).
Let T01,p (Ω) be the set all measurable functions u : Ω → R on Ω and
finite almost everywhere in Ω, such that Tk (u) ∈ W01,p (Ω) for every k > 0.
Every function u ∈ T01,p (Ω) will be identified with its p-quasi continuous
representative.
CHAPTER 4. SECOND ORDER ELLIPTIC PDE 176
Lemma 4.12.2 (Lemma 2.1 of [BBG+ 95]). For every u ∈ T01,p (Ω), there
exists a measurable function v : Ω → Rn such that ∇Tk (u) = vχ{|u|≤k} almost
everywhere in Ω, for every k > 0. Also v is unique up to almost everywhere
equivalence.
177
Appendix A
Let us construct the Cantor set which plays a special role in analysis.
Consider C0 = [0, 1] and trisect C0 and remove the middle open interval
to get C1 . Thus, C1 = [0, 1/3] ∪ [2/3, 1]. Repeat the procedure for each
interval in C1 , we get
C0 ⊃ C1 ⊃ C2 ⊃ . . . ⊃ Ci ⊃ Ci+1 ⊃ . . . .
179
APPENDIX A. CANTOR SET AND CANTOR FUNCTION 180
Proof. If x, y ∈ C are such that z ∈ C for all z ∈ (x, y), then we have the
open interval (x, y) ⊂ C. It is always possible to find i, j such that
j j+1
, ⊆ (x, y)
3i 3i
Proof. Use Cantor’s diagonal argument to show that the set of all sequences
containing 0 and 2 is uncountable.
1
This is true for any positional system. For instance, 1 = 0.99999 . . . in decimal system
APPENDIX A. CANTOR SET AND CANTOR FUNCTION 181
Cantor Function
We shall now define the Cantor function fC : C → [0, 1] as,
∞
! ∞
X ai X ai −i
fC (x) = fC i
= 2 .
i=1
3 i=1
2
Hence, C1 = C11 ∪ C12 . Note that C1i are sets of length a1 carved out from
the end-points of C0 . We repeat step one for each of the end-points of C1i of
length a1 a2 . Therefore, we get four sets
C21 := [0, a1 a2 ] C22 := [a1 − a1 a2 , a1 ],
C23 := [1 − a1 , 1 − a1 + a1 a2 ] C24 := [1 − a1 a2 , 1].
Define C2 = ∪4i=1 C2i . Each C2i is of length a1 a2 . Note that a1 a2 < a1 .
Repeating the procedure successively for each term in the sequence {ak },
we get a sequence of sets Ck ⊂ [0, 1] whose length is 2k a1 a2 . . . ak . The
“generalised” Cantor set C is the intersection of all the nested Ck ’s, C =
2k
∩∞ i
k=0 Ck and each Ck = ∪i=1 Ck . Note that by choosing the constant sequence
ak = 1/3 for all k gives the Cantor set defined in the beginning of this
Appendix. Similar arguments show that the generalised Cantor set C is
compact. Moreover, C is non-empty, because the end-points of the closed
intervals in Ck , for each k = 0, 1, 2, . . ., belong to C.
Lemma A.0.4. For any x, y ∈ C, there is a z ∈
/ C such that x < z < y.
Lemma A.0.5. C is uncountable.
We show in example ??, that C has length 2k a1 a2 . . . ak .
The interesting fact about generalised Cantor set is that it can have non-
zero “length”.
Proposition A.0.6. For each α ∈ [0, 1) there is a sequence {ak } ⊂ (0, 1/2)
such that
lim 2k a1 a2 . . . ak = α.
k
Proof. Choose a1 ∈ (0, 1/2) such that 0 < 2a1 −α < 1. Use similar arguments
to choose ak ∈ (0, 1/2) such that 0 < 2k a1 a2 . . . ak − α < 1/k.
185
BIBLIOGRAPHY 186
187