Diricklet Problem For Laplace
Diricklet Problem For Laplace
Stefano Meda
Università di Milano-Bicocca
A Francesco
Contents
iii
iv CONTENTS
6 Dirichlet L2 155
6.1 Poincare’s inequality . . . . . . . . . . . . . . . . . . . . . . . 155
6.2 Solution to the minimization problem . . . . . . . . . . . . . . 157
6.3 The dual of H01 (Ω) . . . . . . . . . . . . . . . . . . . . . . . . 162
6.4 Weak solutions and regularity . . . . . . . . . . . . . . . . . . 166
Part I
1
Chapter 1
1.1 Introduction
Suppose that Ω is a bounded open set in Rn . We denote by ∆ the Laplace
operator, which acts on a function f in C 2 (Ω) by
n
X
∆f (x) = ∂j2 f (x) ∀x ∈ Ω.
j=1
3
4 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
where the Laplacian ∆ acts on the x variable. Denote by g(x) the tempera-
ture at which the heating and refrigeration system keeps the point x ∈ ∂Ω.
Then we must have
Of course, the body has an initial temperature u(x, 0) at each point x ∈ ∂Ω.
Given that the boundary temperature is kept at a steady state, it is plau-
sible that the system will evolve towards an equilibrium state u(x), that is
Since for all T > 0 the function uT (x, t) := u(x, T + t) satisfies the boundary
value problem (1.1.1)–(1.1.2), it is reasonable to expect that the same will
happen to u. Since u does not depend on t, u will satisfy
(
∆u = 0 in Ω
u|∂Ω = g.
complex plane and the unit disc D := {z ∈ C : |z| < 1}. By this we
mean that ϕ is a holomorphic bijection between Ω and D. Suppose that Ω
has smooth boundary. We look for a homeomorphism ϕ : Ω → D, which,
restricted to Ω, is a biolomorphism between Ω and D. We may assume
that 0 ∈ Ω. By possibly composing ϕ with a Möbius transformation, we
may assume that ϕ(0) = 0. If such a conformal mapping ϕ exists, then
z 7→ ϕ(z)/z is nonvanishing (recall that ϕ0 (z) 6= 0 for all z ∈ Ω, for ϕ is
conformal), whence there exists a holomorphic function f on Ω such that
ϕ(z)
= ef (z) ∀z ∈ Ω.
z
Therefore
log ϕ(z) = log |z| + Re f (z) ∀z ∈ Ω.
Clearly Re f is harmonic in Ω, for it is the real part of a holomorphic function,
and it is a solution to the boundary value problem
(
∆u = 0 in Ω
u|∂Ω = − log |z|.
{(x, y) ∈ R2 : 0 ≤ x ≤ π, y ≥ 0}.
which we write as −c, where c is, for the time being, a complex number.
Then, we are led to solve the ordinary differential equations
X 00 + c X = 0 and Y 00 − c Y = 0.
All the solutions to this problem are multiples of e−ky . Thus, we are led to
consider the functions
They satisfy
Unless the datum f is one of the functions sin(kx), none of the functions uk
will match the boundary condition uk (x, 0) = f (x). The idea to circumvent
this difficulty is to consider superpositions of the functions uk , i.e., to see
whether functions of the form
∞
X
ck e−ky sin(kx) (1.1.6)
k=1
satisfy the given Dirichlet problem. Of course, this idea is suggested by the
fact that the Laplace operator v 7→ ∆v is linear. At least formally, if the
function above satisfies the Dirichlet problem, then its value at (x, 0) must
be equal to f (x). In other words, f should have the following Fourier sine
series expansion
∞
X
f (x) = ck sin(kx).
k=1
Exercise 1.1.2 Show that the method of separation of variables for the
Dirichlet problem discussed above leads, in the case where f (x) = 1, to the
solution
4 h −y 1 1 i
u(x, y) := e sin x + e−3y sin(3x) + e−5y sin(5x) + · · ·
π 3 5
2 sin x
= arctan .
π sinh y
Show that u satisfies the boundary conditions except at the corners. Draw
the isothermals for small x and y. Hint. To find the sum of the series within
square brackets, write z = x + iy and observe that the series is the imaginary
part of a power series in the variable eiz , whose sum is (1/2) log[(1+eiz )/(1−
eiz )].
(x0 , xn ) 7→ x0 , xn − ψ(x0 )
x0 7→ x0 , ψ(x0 ) ,
∇φ(x)
ν(x) = ± .
∇φ(x)
This formula shows that the normal field x 7→ ν(x) is of class C k−1 on S.
We say that a domain Ω has C k boundary if ∂Ω is a hypersurface of class
k
C . We recall the classical divergence theorem.
In the applications, the boundary does not always satisfy the assumptions
of Theorem 1.2.1. There is a generalisation of the divergence theorem which
covers domains with a very general boundary. It is beyond the scope of these
notes to deal with such generalisations, for which the reader is referred to the
book of Evans and Gariepy [EG]. We just make a few comments concerning
domains with Lipschitz boundary.
Recall that a map ϕ : Rn−1 → R is Lipschitz if there exists a constant L
such that
ϕ(x) − ϕ(y) ≤ L |x − y| ∀x, y ∈ Rn−1 .
1.2. BACKGROUND AND PRELIMINARY RESULTS 9
The infimum of all the constants L such that the above inequality holds is
called the Lipschitz constant of ϕ.
Lipschitz functions are possibly not differentiable at some points (for in-
stance, x 7→ |x| is Lipschitz with constant 1, but it is not differentiable at
the origin), but the set where it is not differentiable is small in the measure
theoretic sense. In fact, the following nontrivial result holds. We refer the
reader to [EG] for the proof of this result.
(i) Z Z
∆u dV = ∂ν u dσ;
Ω ∂Ω
Exercise 1.2.6 Prove that the measure of the unit ball in Rn is ωn /n.
Hint: use the divergence theorem for an appropriate vector field.
Exercise 1.2.7 By following the steps below, prove that the measure of the
unit ball in Rn is π n/2 /Γ(n/2 + 1) (here Γ denotes Euler’s Gamma function):
(ii) compute
Z
I := p θ1 , . . . , θn−1 dθ1 · · · dθn−1 ,
R
by computing Z
exp − a|x|2 dx
Rn
1
divx J (F ◦ φ) [(φ0 )t ]−1 (x);
(ii) divy F(y) =
J(x)
1 −1
divx J ∇x (f ◦ φ) (φ0 )t φ0
(iii) ∆y f (y) = (x);
J(x)
(vi) suppose that φ is a conformal mapping, i.e., that φ0 (x) = %(x) U (x),
where %(x) is a scalar factor and U (x) is an orthogonal matrix. Prove
that
1 n−2
∆y f (y) = div x % ∇x (f ◦ φ) (x);
%(x)n
where w = φ(z).
Then find all radial solutions to the equation ∆u = 0 in R3 \ {0}. Prove also
that the differential operator within square brackets in the formula above is
the Laplace–Beltrami operator on the sphere S2 with respect to the Rieman-
nian metric induced by the standard Euclidean metric of R3 .
Exercise 1.3.2 Prove that the real and the imaginary parts of a holomor-
phic function in a domain Ω are harmonic in Ω. Prove that the functions
x2 − y 2 and xy are harmonic in R2 , and so are the functions rj cos(jθ) and
rj sin(jθ) for each nonnegative integer j.
Proof. We shall prove (i). The proof of (ii) is similar and is omitted. Part
(iii) is a straightforward consequence of (i) and (ii).
To prove (i), suppose that B = BR (y), that 0 < ρ < R, and observe that,
by Corollary 1.2.5 (i),
Z
0≤ ∂ν u dσ
∂Bρ (y)
ω0
Z
= ∇u(y + ω 0 ) · 0 dσρ (ω 0 )
∂Bρ (0) |ω |
Z
= ρn−1 ∇u(y + ρω) · ω dσ1 (ω)
∂B1 (0)
Z
n−1 d
=ρ u(y + ρω) dσ1 (ω)
∂B1 (0) dρ
Z
n−1 d
=ρ u(y + ρω) dσ1 (ω)
dρ ∂B1 (0)
h 1 Z
n−1 d
i
=ρ u dσ .
dρ ρn−1 ∂Bρ (y)
Now we integrate both sides with respect to ρ between ε and R and obtain
Z Z
1 1
0 ≤ n−1 u dσ − n−1 u dσ.
R ∂BR (y) ε ∂Bε (y)
Since u is continuous in y,
Z
1
lim u dσ = ωn u(y).
ε→0 εn−1 ∂Bε (y)
1.3. SUBHARMONIC FUNCTIONS 15
where Ha,b is the harmonic function on [a, b] such that Ha,b (a) = u(a) and
Ha,b (b) = u(b).
Ωε := {x : B ε (x) ⊂ Ω},
16 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
Now, the left hand side can be differentiated infinitely many times (see Propo-
sition ?? (iii), whence u is in C ∞ (Ωε ). Since ε is arbitrary, u is in C ∞ (Ω).
It remains to show that u is harmonic. Suppose that B r (x) ⊂ Ω. By
Corollary 1.2.5 (i),
Z Z
∆u dV = ∂ν u dσ
Br (x) ∂Br (x)
Z
ω
= grad u(x + ω) · dσ(ω)
|ω|=r |ω|
Z
=r n−1
grad u(x + rω 0 ) · ω 0 dσ(ω 0 )
|ω|=1
Z
n−1 d
=r u(x + rω 0 ) dσ(ω 0 ).
dr |ω0 |=1
which, in turn is equal to ωn u(x), for u possesses the mean value property
by assumption. Thus, we have proved that
Z
d
∆u dV = ωn rn−1 u(x) = 0
Br (x) dr
Deduce that if u satisfies the mean value inequality of Theorem 1.3.7 (i),
then u is subharmonic.
Proof. We prove (i). The proof of (ii) is similar and is omitted. Part (iii) is
a direct consequence of (i) and (ii).
Set M := supΩ u, and
ΩM := {x ∈ Ω : u(x) = M }.
sup u = sup u;
Ω ∂Ω
inf u = inf u;
Ω ∂Ω
inf u ≤ u ≤ sup u.
∂Ω ∂Ω
Proof. To prove (i), observe that if supΩ u > sup∂Ω u, then there exists a
point y in Ω for which u(y) = supΩ u. Consequently u is constant in Ω by
Theorem 1.4.1 (i). Since u is continuous on Ω, supΩ u = sup∂Ω u, thereby
contradicting the assumption.
The proof of (ii) is almost verbatim the same as the proof of (i), and is
omitted.
Part (iii) follows directly from (i) and (ii). 2
1.5. GREEN’S FUNCTION 19
Exercise 1.4.3 Suppose that u and v are in C 2 (Ω) ∩ C(Ω) and that
u|∂Ω = v|∂Ω .
Prove that if u is harmonic and v is subharmonic, then v ≤ u on Ω. This
result justifies the term subharmonic. State and prove a corresponding result
for superharmonic functions.
Theorem 1.4.4 Suppose that u and v are functions in C 2 (Ω) ∩ C(Ω) that
solve the Dirichlet problem
(
∆u = f in Ω
.
u|∂Ω = g
Then u = v.
∆u = f,
(iv) suppose that x ∈ Ω and that there exists a function hx ∈ C 1 (Ω) which is
harmonic on Ω and satisfies (hx )|∂Ω = −N (x − ·)|∂Ω . Denote by G(x, ·)
the function hx + N (x − ·). Then
Z Z
u(x) = u(y) ∂ν G(x, y) dσ(y) + G(x, y) ∆u(y) dV (y).
∂Ω Ω
Proof. We prove the theorem in the case where n ≥ 3. The case n = 2 is left
to the reader.
First we prove (i). Consider the domain Ω \ B r (x), for r small. Apply
second Green’s identity (see Corollary 1.2.5 (iii)) with N (x − ·) in place of v
and Ω \ B r (x) in place of Ω. We obtain
Z
N (x − y) ∆u(y) dV (y)
Ω\B r (x)
Z
= N (x − y)∂ν u(y) − u(y) ∂ν N (x − y) dσ(y)
∂Ω Z
+ N (x − y)∂ν u(y) − u(y) ∂ν N (x − y) dσ(y).
∂Br (x)
The required formula will follow from this and the dominated convergence
theorem once we prove that
Z
lim N (x − y) ∂ν u(y) dσ(y) = 0 (1.5.3)
r↓0 ∂Br (x)
and that Z
lim u(y) ∂ν N (x − y) dσ(y) = u(x). (1.5.4)
r↓0 ∂Br (x)
22 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
Since u is smooth,
|u(y) − u(x)| ≤ r max |∇u| ∀y ∈ ∂Br (x).
Ω
Note that, by second Green’s identity and the fact that hx is harmonic,
Z Z Z
∂ν u hx dσ = u ∂ν hx dσ + hx ∆u dV,
∂Ω ∂Ω Ω
The last part of the theorem above suggest that the function G defined
therein may play an important role in the theory. This justifies the following
definition.
Note that we do not assert the existence of a Green’s function for a generic
domain Ω. In fact, there are domains which do not admit a Green’s function.
Exercise 1.5.7 Prove that the Green’s function for Ω, if it exists, is unique.
Definition 1.5.9 Suppose that G is the Green function for the domain Ω.
The function (x, y) 7→ ∂ν G(x, y) (here the normal derivative is taken with
respect to the y variable) is called a Poisson kernel for the domain Ω.
Hints: (i) for every z in Ω and every (small) ε > 0, denote by Ωε (z) the set Ω\
B ε (z). Then write the harmonic function G(x, ·) by using the representation
formula in Theorem 1.5.5 (iii) with Ωε (x) in place of Ω and similarly write
the harmonic function G(y, ·) by using the representation formula with Ωε (y)
in place of Ω. Then use second Green’s identity and the fact that G(x, ·) =
N (x − ·) + hx (·);
(ii) note that G(x, y) tends to −∞ if y tends to x. Therefore, G < 0
near x. Apply the maximum principle to the domain Ω \ Bε (x);
(iii) observe that, by the minimum principle, hx > 0 in Ω and use (ii)
and the definition of G;
(iv) use (iii) and the Lebesgue dominated convergence theorem.
First we need to compute the Green’s function for Rn+ . Fix a point x =
(x0 , xn ) in the upper half space, and consider the Newtonian potential N (x−·)
with pole x. We must find a function hx in C 1 (Rn+ ) that is harmonic on Rn+
and such that N (x − ·) + hx vanishes identically when xn = 0. Observe
that the Newtonian potential is, up to a constant, the potential generated
by a unit negative charge placed at the point x∗ := (x0 , −xn ). By symmetry,
its values on the hyperplane xn = 0 are the same of those of a Newtonian
potential generated by a unit negative charge placed at the point x∗ . Thus,
the Green’s function G of Rn+ is given by
1 log |x − y|
∗ − y|
if n = 2
G(x, y) = 2π |x (1.6.1)
1 ∗
2−n 2−n
(2−n) ωn
|x − y| − |x − y| if n ≥ 3.
Exercise 1.6.1 Check that the function G defined above is the Green’s
function for the upper half space.
First we consider the case where n ≥ 3. By definition, for every x ∈ Rn+ and
y 0 ∈ Rn−1 the Poisson kernel P (x, y 0 ) is then given by
P (x, y 0 ) = ∂ν G(x, y 0 )
= −∂yn G(x, y 0 )
1 h y n − xn y n + xn i
=− − .
ωn |x − y|n |x∗ − y|n |yn =0
Observe that |x − y| = |x∗ − y| when yn = 0. Hence
2 xn
P (x, y 0 ) =
ωn |x − y|n
2 xn (1.6.2)
= ∀xn > 0, x0 , y 0 ∈ Rn−1 .
ωn |x0 − y 0 |2 + x2 n/2
n
A similar computation shows that the formula above holds also in the case
where n = 2.
Exercise 1.6.2 Compute the Poisson kernel for a quadrant of the plane.
Hint: refine the method of images illustrated above for the half plane.
may have more than one solution. In particular, if g vanishes identically, then
the null function and the function u(x0 , xn ) = xn both solve the Dirichlet
problem above. Thus, if we want to recover uniqueness, we must impose
further restriction on the solution. A typical statement which holds in this
case is the following.
Observe that kpxn k1 = kpk1 , thereby justifying the terminology. With this
notation, formula (1.6.3) may be rewritten as follows
u(x) = g ∗ pxn (x0 ) ∀x ∈ Rn+ , (1.6.4)
where ∗ stands for convolution on Rn−1 .
Exercise 1.6.5 Prove that (1.6.3) gives, indeed, a solution to the Dirichlet
problem on the upper half space with boundary datum g.
Proof. For each ζ ∈ ∂Ω the point ϕ(ζ) is in ∂(ϕ(Ω)). Therefore for each
z∈Ω
Gϕ(Ω) (ϕ(z), ϕ(ζ) = 0,
because Gϕ(Ω) is the Green’s function for ϕ(Ω). For the same reason
∆w Gϕ(Ω) (w, ϕ(ζ)) = 0.
Since ϕ is conformal,
∆w Gϕ(Ω) (w, ϕ(ζ)) = |ϕ0 (z)|2 ∆z Gϕ(Ω) (ϕ(z), ϕ(ζ)),
by (1.2.9) (viii), whence GΩ (·, ζ) is harmonic for every ζ in Ω.
Finally,
1
GΩ (z, ζ) − log |z − ζ|
2π
1
= Gϕ(Ω) (ϕ(z), ϕ(ζ) −
log |z − ζ|
2π
1 1
= log |ϕ(z) − ϕ(ζ)| + hϕ(z) ϕ(ζ) − log |z − ζ|
2π 2π
1 ϕ(z) − ϕ(ζ)
= log + hϕ(z) ϕ(ζ)
2π z−ζ
28 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
We apply this result to the case where the domain Ω is just the unit disc
D := {z ∈ C : |z| < 1} and
1+z
ϕ(z) = i ∀z ∈ D.
1−z
It is not hard to check that ϕ maps D conformally onto the upper half
plane Π, and ∂D onto R ∪ ∞. In the last section we have proved that the
Green’s function for the upper half plane is given by
1 |w − ω|
GΠ (w, ω) := log .
2π |w − ω|
1
z 7→ ,
z
which is the inversion with respect to the circle {|z| = 1}. Observe that
1 |ϕ(z) − ϕ(ζ)|
= log
2π |ϕ(1/z) − ϕ(ζ)|
1 | 1+z
1−z
− 1+ζ
1−ζ
|
= log 1+(1/z) 1+ζ
2π | 1−(1/z) − 1−ζ |
1 |z − ζ|
= log .
2π |1 − z ζ|
1.7. THE POISSON KERNEL FOR THE BALL 29
The aim of the following series of exercises is to determine the Green’s func-
tion, hence the Poisson kernel, of certain domains in the complex plane that
are conformally equivalent to the unit disc.
Exercise 1.7.2 Compute the Green’s function and the Poisson kernel of
the half disc
F := {z ∈ C : |z| < 1, Im(z) > 0}.
Exercise 1.7.3 Prove that the map z 7→ ez is a conformal map of the strip
S := {z ∈ C : 0 < Im(z) < π}
onto the upper half plane. Then compute the Green’s function and the
Poisson kernel of the strip S.
Exercise 1.7.4 Prove that the map z 7→ z α/π is a conformal map of the
upper half plane onto the sector
Γ := {z ∈ C : 0 < arg(z) < α}.
Then compute the Green’s function and the Poisson kernel of the sector Γ.
30 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
Lemma 1.7.5 Suppose that Y ∈ ∂B1 (0) and that x ∈ B1 (0) \ {0}. Then
x
|Y − x| = |x| Y − .
|x|
(i) the function, defined for every x ∈ B1 (0) and for every y ∈ B1 (0) by
1 2−n
|x − y|2−n − |x|−1 x − |x| y
= ,
(2 − n) ωn
R2 − |x|2
PR (x, Y ) = ∀x ∈ BR (0) ∀Y ∈ ∂BR (0). (1.7.2)
ωn R |x − Y |n
Proof. To prove (i) observe that the first equation shows that G1 (x, ·) is
harmonic in B1 (0), and, by Lemma 1.7.5, the second proves that G1 (x, y) =
N (x − y) + hx (y), where hx is harmonic in B1 (0) and hx (Y ) = −N (x − Y )
when Y ∈ ∂B1 (0). Since the Green’s function, if it exists, is unique, G1 is
the Green’s function of B1 .
Next we prove (ii). It is straightforward to check that the function
1
GR (x, y) = G1 (x/R, y/R)
Rn−2
1.7. THE POISSON KERNEL FOR THE BALL 31
is the Green’s function for BR (0). Indeed, by Lemma 1.7.5, if Y ∈ ∂BR (0)
and x ∈ BR (0), then
|x| x
|Y − x| = Y −R . (1.7.3)
R |x|
1 2−n
h x Y 2−n R x |x| Y 2−n i
GR (x, Y ) = R − − −
(2 − n) ωn R R |x| R R R
1 h
2−n R |x| 2−n i
= |x − Y | − x− Y
(2 − n) ωn |x| R
= 0,
where the last equality follows from (1.7.3). Denote by ν the outward normal
of ∂BR (0) at the point Y . By the definition of Poisson kernel,
PR (x, Y ) = ∂ν GR (x, Y )
|x| R
1 h Y −x Y |x| R
Y − |x|
x Y i
= n
· − n ·
ωn |x − Y | |Y | R R
x− |x|
Y |Y |
|x| R
R2 − |x|2
= ,
ωn R |x − Y |n
as required. 2
Then
R2 − |y|2
Z
g(X)
u(y) = dσ(X) ∀y ∈ Ω.
ωn R ∂BR (0) |y − X|n
In addition, we can prove that a solution of the Dirichlet problem above with
continuous boundary data exists. This is the main result of this section.
32 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
Theorem 1.7.8 Suppose that g ∈ C(∂BR (0)). Then the Dirichlet problem
(
∆u = 0 in Ω
u|∂Ω = g
and that
R2 − |y|2
Z
dσ(X)
=1 ∀y ∈ BR (0). (1.7.6)
ωn R ∂BR (0) |y − X|n
These formulae will imply the required conclusion (1.7.4). The second for-
mula is a direct consequence of the representation formula of Corollary 1.7.7
(with the constant function 1 in place of g and u).
To prove (1.7.5), we proceed as follows. Since g is continuous in X0 , for
ε > 0 there exists δ > 0 such that
We split the integral over ∂BR (0) as the sum of the integral over ∂BR (0) ∩
B2δ (X0 ) and the integral over ∂BR (0) \ B2δ (X0 ), and estimate them sepa-
rately.
To estimate the first, observe that
g(X) − g(X0 )
Z Z
n
dσ(X) < ε |y − X|−n dσ(X)
∂BR (0)∩B2δ (X0 ) |y − X|
Z∂BR (0)∩B2δ (X0 )
≤ε |y − X|−n dσ(X)
∂BR (0)
ωn R
=ε ;
R2 − |y|2
we have used (1.7.6) in the equality above. Hence
R2 − |y|2 g(X) − g(X0 )
Z
dσ(X) < ε ∀y ∈ BR (0).
ωn R ∂BR (0)∩B2δ (X0 ) |y − X|n
Moreover
g(X) − g(X0 )
Z Z
n
dσ(X) ≤ 2kgk∞ |y − X|−n dσ(X)
∂BR (0)\B2δ (X0 ) |y − X| ∂BR (0)\B2δ (X0 )
is subharmonic in Ω.
We leave the reader the task of stating and proving the corresponding prop-
erties for superharmonic functions.
Proof. We first prove (i). Suppose that y is a point in Ω such that
u(y) = sup u.
Ω
M := sup (u − v).
Ω
M = (u − v)(x0 ).
(u − v)(x0 ) = 0.
α
n |α| |α|
sup D u ≤ sup |u|.
K dist(K, ∂Ω) Ω
Proof. We prove the required estimate in the case where |α| = 1. Since ∂j u
is harmonic in Ω (because ∂j and ∆ commute), we may use the mean value
property for harmonic functions and write
Z
1
∂j u(x) = ∂j u dV
V BR (x) BR (x)
Z
1
= u νj dσ.
V BR (x) ∂BR (x)
We have used the formula of integration by parts in the last equality. Here
R < dist(K, ∂Ω), so that BR (x) ⊂ Ω. Therefore
1
∂j u(x) ≤ σR ∂BR (x) sup |u|
V BR (x) Ω
n
= sup |u|
R Ω
The required estimate follows by taking the supremum of both sides with
respect to R < dist(K, ∂Ω). 2
is harmonic in Ω.
It may very well happen that {v` } is unbounded from below. However, the
sequence {max(vn , inf ϕ)} is bounded from below, and still
u(y) = lim max(v` (y), inf ϕ),
`→∞
40 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
so that we may assume that {v` } is bounded. Now, we choose R such that
BR (y) ⊂⊂ Ω. We consider the harmonic lifting HBR (y) (v` ) of v` . Recall that
HBR (y) (v` ) is subharmonic in Ω, and harmonic in BR (y).
Clearly HBR (y) (v` ) is a subfunction relative to ϕ and
u(y) = lim HBR (y) (v` )(y).
`→∞
Observe that {HBR (y) (v` )} is a sequence of uniformly bounded harmonic func-
tions on BR (y). By Corollary 1.8.9, there exists a subsequence {`j } such that
{HBR (y) (v`j )} is uniformly convergent on BR/2 (y) to a function v, which is
harmonic in BR/2 (y).
Obviously v ≤ u in BR/2 (y). We claim that v = u in BR/2 (y).
We argue by contradiction. Suppose that there exists z ∈ BR/2 (y) such
that
v(z) < u(z).
By the definition of the function u, there exists a subfunction ve relative to ϕ
in Ω such that
v(z) < ve(z) ≤ u(z). (1.8.1)
Consider the sequence
wj := max ve, HBR (y) (v`j ) .
Clearly wj is a subfunction relative to ϕ. Its harmonic lifting HBR (y) (wj ) is
harmonic in BR (y),
u(y) = lim HBR (y) (wj )(y),
j→∞
v ≤ w ≤ u on BR/2 (y), v(z) < ve(z) ≤ w(z) and v(y) = w(y) = u(y).
Observe that v − w is harmonic in BR/2 (y) and that (v − w)|∂BR/2 (y) ≤ 0.
Since (v − w)(y) = 0, v − w = 0 in BR (y) by the maximum principle. This
contradicts (1.8.1). Therefore v = u on BR/2 (y). This proves the claim.
The proof of the theorem is complete. 2
u = sup v ≤ w in Ω.
v∈Sϕ
Thus, w = u in Ω, as required. 2
(i) w is superharmonic in Ω;
m := inf w > 0.
BR (X)\B
is a barrier at X relative to Ω.
Indeed, w
e is clearly continuous on Ω, and is superharmonic in Ω by
e > 0 on Ω \ {X} and w(X) = 0.
Lemma 1.8.4 (iv). Furthermore, w
Proof. Set M := sup |ϕ| and choose ε > 0. Since X is a regular point, there
exists a barrier w at X relative to Ω. Since ϕ is continuous at X, there exists
δ > 0 such that
|ϕ(Y ) − ϕ(X)| < ε
1.8. PERRON’S METHOD 43
Now, the problem becomes to give criteria which imply that every point of
the boundary of a given domain are regular. A simple criterion is discussed
in the following exercise.
44 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
Exercise 1.8.22 Prove that every bounded Lipschitz domain satisfies the
exterior cone condition.
Proof. Fix a point X ∈ ∂Ω. Since Ω satisfies the uniform exterior cone
condition, there exist constants α and h such that the cone Γ(X, α, h) is
such that Γ(X, α, h) ∩Ω = {X}. We normalise our setting by assuming that
Γ(X, α, h) ∩ X + Sn−1 6= ∅. To construct a barrier at X we argue as follows.
Consider all functions v of the form
w(x) = rλ ϕ0 (ω),
where x = r ω (polar coordinates centred at X), λ > 0, and ϕ0 is the
eigenfunction associated to the lowest eigenvalue µ of the eigenvalue problem
∆Sn−1 U = −µ U
on Sn−1 \ Γ(0, α, h) with Dirichlet boundary conditions. A well known result
in spectral theory, which we do not prove in these notes, asserts that µ > 0
and that ϕ0 is a smooth function, which is strictly positive in Sn−1 \Γ(0, α, h).
By using the formula for the Laplacian ∆ in polar coordinates, we find that
∆w(x) = rλ−2 ϕ0 α2 + (n − 2) α − µ .
p
It is straightforward to check that if α = (1/2) (n − 2)2 + 4µ − (n − 2) ,
then w is harmonic in Ω. Furthermore, w(X) = 0, and w > 0 in Ω \ {X}.
Thus, w is a barrier at X. Hence X is a regular point relative to Ω.
Since this holds for every point in ∂Ω, the classical Dirichlet problem
is solvable by Lemma 1.8.18. The uniqueness follows from the maximum
principle. 2
1.9. THE LEBESGUE SPINE 45
(see Section 5.1 and, in particular, Theorem 5.1.2, for more on the connection
between measures and continuous linear functionals on Cc (Rn )). Note that
the support of µ is contained in the segment
Σ := {(0, 0, z) ∈ R3 : z ∈ [0, 1]}.
We consider the potential associated to the measure µ, i.e., the function u
given by
u = N ∗ µ.
Since µ has compact support, the convolution on the right hand side of the
formula above makes perfect sense (see Definition 5.4.9 below). We shall
compute u explicitly. For every ϕ ∈ Cc∞ (R3 )
hϕ, N ∗ µi = ϕ ∗ Ň , µ
Z 1
= ϕ ∗ Ň (0, 0, τ ) τ dτ
0
Z 1 Z
= τ dτ ϕ(x, y, z) N (x, y, z − τ ) dx dy dz
0 R3
Z Z 1
1 τ
= ϕ(x, y, z) dx dy dz p dτ.
4π R3 0 x + y + (z − τ )2
2 2
46 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
Therefore
Z 1
1 τ
N ∗ µ(x, y, z) = p dτ.
4π 0 x2 + y 2 + (z − τ )2
The right hand side can be computed explicitly. We find that, up to a
constant factor,
p √
N ∗ µ(x, y, z) = r2 + (z − 1)2 − r2 + z 2
√ p
+ z log r2 + z 2 + z r2 + (z − 1)2 + 1 − z
− 2z log r ∀(x, y, z) ∈ Ω such that z > 0.
Ωε := {(x, y, z) ∈ Ω : r2 > ε2 },
It may be worth noticing that there are domains with inward cusps for which
the classical Dirichlet problem is solvable. For instance, it may be shown that
this is the case for the domain
for every positive integer k. The proof of this fact uses the theory of capacity,
which we do not tackle in this notes. The interested reader is referred to [K].
Similar results hold in Rn for n ≥ 3. However, the case of bounded domains
in the plane is quite different. Indeed, it is straightforward to show that if
Ω is a planar domain and z0 is a boundary point for which the following is
true: there exists a straight line segment [z0 , z1 ] entirely contained in R2 \ Ω,
then z0 is a regular point relative to Ω.
This criterion implies that the two dimensional analogue of the Lebesgue
spine is a regular point relative to Ω.
48 CHAPTER 1. THE CLASSICAL DIRICHLET PROBLEM
Part II
49
51
then Z
u(x) = g(Y ) ∂ν(Y ) G(x, Y ) dσ(Y ) ∀x ∈ Ω
∂Ω
(see Corollary 1.5.8), where G is the Green’s function of Ω, and ν(Y ) denotes
the outward unit normal to ∂Ω at the point Y . Furthermore,
The idea that leads to the method of integral equations is to replace the
Green’s function of Ω with the Newtonian potential N in the integral above,
and to consider the operator that associates to each reasonable function g on
∂Ω the function Dg, defined by
Z
Dg(x) = g(Y ) ∂ν(Y ) N (x − Y ) dσ(Y ) ∀x ∈ Ω.
∂Ω
Clearly Dg will not be the solution of the Dirichlet problem with datum g
for the obvious reason that N (x − Y ) does not agree with G(x, Y ). However,
the operator D is related to the solution of the Dirichlet problem, as we shall
explain below. The operator g 7→ Dg is called the double layer potential
and the function Dg is called the double layer potential with density g.
Its properties will play a key role in the theory.
We shall prove that for ϕ ∈ C(∂Ω), the integral
Z
ϕ(Y ) ∂ν(Y ) N (X − Y ) dσ(Y )
∂Ω
makes sense also when X ∈ ∂Ω. Thus, we may define the operator K on
C(∂Ω) by
Z
Kϕ(X) := ϕ(Y ) ∂ν(Y ) N (X − Y ) dσ(Y ) ∀X ∈ ∂Ω.
∂Ω
52
Thus, the boundary value of the double layer potential Dϕ with density ϕ is
the datum g provided that ϕ satisfies the following integral equation
1
g(X) = ϕ(X) + Kϕ(X) ∀X ∈ ∂Ω.
2
It is a nontrivial fact, which will be proved below, that for every contin-
uous function g on ∂Ω the integral equation above has a unique solution
ϕ ∈ C(∂Ω). In the language of spectral theory, this statement may be refor-
mulated by saying that −1/2 is in the resolvent set of the operator K.
The operator K is a compact operator. This is the reason why in the
following we shall study in some detail the class of compact operators and
their spectra.
Clearly B ⊃ A.
We consider two more examples of linear operators, which are one dimen-
sional models of the Dirichlet and Neumann Laplacian. Suppose that
53
54 CHAPTER 2. LINEAR OPERATORS ON BANACH SPACES
−∞ < a < b < ∞. Let HD and HN be the operators on L2 ((a, b)) defined by
It is not hard to check that the domains of HD and HN are dense in L2 ((a, b)).
kAf k
sup < ∞.
f 6=0 kf k
Exercise 2.2.3 Prove that for a linear operator A on B the following are
equivalent:
(iii) A is continuous at 0.
Denote by H01 (I) the completion of Cc∞ (I) with respect to the norm above.
The following proposition illustrates the role of Exercise 2.2.4 in proving the
boundedness of certain operators.
Proposition 2.2.6 Suppose that I is a bounded interval of the real line. The
identity map j : ϕ 7→ ϕ, initially defined on Cc∞ (I), extends to a bounded
linear operator from H01 (I) to L2 (I).
and the required estimate follows by integrating the squares of both sides
on I. 2
It is worth observing that the space H01 (I) may be realised as a (nonclosed)
subspace of L2 (I).
Indeed, to prove that the map j in Proposition 2.2.6 is an immersion, we
have to show that if f ∈ H01 (I) satisfies j(f ) = 0, then f = 0. Note that
this is trivial if f ∈ Cc∞ (I), because j(f ) = f as functions. However, we do
not know a priori how the map j acts on elements of H01 (I), so we need to
prove that j(f ) = 0 implies f = 0. Since H01 (I) is the completion of Cc∞ (I),
56 CHAPTER 2. LINEAR OPERATORS ON BANACH SPACES
given f ∈ H01 (I), there exists a sequence {fn } of functions in Cc∞ (I) such
that kfn − f kH01 (I) → 0 as n tends to ∞. Note that kfn k2 → 0, because j is
continuous by Proposition 2.2.6, and fn = j(fn ) → j(f ) = 0.
Furthermore, {fn0 } is a Cauchy sequence in L2 (I) by the definition of the
norm in H01 (I). Since L2 (I) is complete, there exists a function g ∈ L2 (I) such
that kfn0 − gk2 → 0 as n tends to ∞. By possibly passing to a subsequence,
we deduce that {fn } is convergent to 0 a.e. By the fundamental theorem of
calculus Z x
fn (x) = fn0 (s) ds,
0
so that, by taking the limit of both sides,
Z x
0= g(s) ds
0
Exercise 2.2.7 Prove that the operators A and B defined just below Def-
inition 2.1.2 do not admit bounded extensions to L2 (R).
Proposition 2.2.8 The space L(B) is a Banach space with respect to the
operator norm.
Consequently,
|||A − An ||| ≤ ε ∀n ≥ ν.
(i) A is bounded;
Proof. Exercise. 2
Consider the operator A defined just below Definition 2.1.2. There exists a
(purely algebraic) extension of A to a linear operator on B 1 . Note that this
extension cannot be bounded for we have proved in Exercise 2.2.7 that A
does not admit a bounded extension to all of B.
This extension provides an example of an unbounded linear operator
which is defined everywhere.
and C, defined on `1 by
x x
2 3
C(x1 , x2 , x3 , . . .) = x1 , , , . . . .
2 3
The main reason for introducing the residual spectrum of an operator is that
self adjoint operators have empty residual spectrum.
= sup (Af, g)
kf k≤1
= |||A||| kgk,
60 CHAPTER 2. LINEAR OPERATORS ON BANACH SPACES
and that |||A∗ ||| ≤ |||A|||. By reversing the role of A and A∗ , we find that
|||A||| ≤ |||A∗ |||, whence |||A||| = |||A∗ |||.
Tf = f ∗ ϕ ∀f ∈ L2 (R),
L(x1 , x2 , . . .) = (x2 , x3 , . . .)
for every sequence {xn } in `2 . Compute L∗ . Do the same for the right shift
operator R, defined by
= sup |(Af, Af )|
kf k=1
= sup |(f, A∗ Af )|
kf k=1
≤ |||A∗ A|||,
and the required equality follows. 2
The properties of the involution ∗ contained in the proposition above and
the fact that L(H) is complete are usually summarised by saying that L(H)
is a C ∗ -algebra (with involution ∗).
Proof. Denote by ηA the right hand side of the equality above. The inequality
ηA ≤ |||A||| is trivial and its proof is left to the reader.
It remains to prove that |||A||| ≤ ηA . We use the following polarisation
identity: for every f , g in H
which is easily verified. Observe that for each ϕ in H the number (Aϕ, ϕ) is
real, for
(Aϕ, ϕ) = (ϕ, A∗ ϕ) = (ϕ, Aϕ) = (Aϕ, ϕ).
Here we have used the assumption that A is self adjoint. From (2.4.2) we
then deduce that
whence
4 Re(Af, g) ≤ ηA kf + gk 2 + kf − gk 2
(2.4.3)
= 2ηA kf k 2 + kgk 2 .
The norm of B is the infimum of all M for which the above inequality holds.
= sup B(f, g)
kgk=1
≤ M kf k,
where we have used the boundedness of the form B in the last inequality.
Therefore the operator A is bounded: A will be called the bounded linear
operator associated to the form B.
Thus, bounded bilinear forms and bounded operators on H are somewhat
“interchangeable”. Furthermore, if A is self adjoint, then the associated
bilinear form is symmetric, i.e.
U ∗ U = U U ∗ = I.
(i) U is unitary;
The next proposition relates the kernel and the range of the bounded oper-
ators A and A∗ .
(x, A∗ y) = 0 ∀x ∈ H,
which is equivalent to
(Ax, y) = 0 ∀x ∈ H,
Note that Ker(A) and Ker(A∗ ) are always closed subspaces of H, but this is,
in general, no longer true for Ran(A) and Ran(A∗ ). Therefore the statements
It is a well known fact (that we shall not prove here) that the closed unit ball
in a Banach space B is compact if and only if B is finite dimensional. As a
consequence, the identity operator I on a Banach space B is compact if and
only if B is finite dimensional.
Exercise 2.5.2 Suppose that T is in L(B). Prove that the following are
equivalent:
(iii) for every bounded sequence {fk } in B, there exists a subsequence {fnk }
such that T fnk is convergent in B.
Clearly, finite rank operators are compact, because the image of the unit
ball B in B is a bounded set in the finite-dimensional Banach space Ran(T ),
and therefore it is relatively compact.
Compact operators need not have finite rank. Indeed, consider the inter-
val I = [0, 1] and the operator T : C(I) → C(I), defined by
Z x
T f (x) = f (s) ds ∀f ∈ C(I) ∀x ∈ I.¯
0
Exercise 2.5.4 Prove that the vector space K(B) of all compact linear
operators on B is a two sided closed ideal in L(B).
Indeed,
|||C − Pn C||| = sup k(I − Pn )Cxk ≤ sup k(I − Pn )yk,
x∈B y∈K
as required. 2
Hence X
kAf k 2 = |cj |2 |λj |2 .
j>N
For every ε > 0 there exists ν such that |λj | < ε for every N ≥ ν. Therefore
X
|||A − PN A||| ≤ sup ε2 |cj |2
kf k=1,f ∈Ran(PN )⊥ j>N
= sup ε kf k 2 ,
2
kf k=1,f ∈Ran(PN )⊥
m
X
gm = cm
j ϕj .
j=1
Thus,
m
X
cm
m ϕm = gm − cm
j ϕj . (2.6.2)
j=1
Therefore
Agm − λm gm ∈ Hm−1 .
Clearly {gm } is a bounded sequence in H. We shall prove that {Agm } does
not admit a convergent subsequence, thereby contradicting the compactness
of A.
Indeed, suppose that m < n. Then
kAgn − Agm k2 = kλn gn − vk2 ,
⊥
where, by (2.6.2), v ∈ Hn−1 . Since gn ∈ Hn−1 ,
kAgn − Agm k2 ≥ kλn gn k2
= |λn |2 kgn k2
≥ µ2 .
Hence {Agm } has no Cauchy subsequences, as required.
2
Operators of the form λ I − A, where A is compact, arise also in the theory of
second order elliptic equations, as we shall see. Fredholm studied some of the
properties of these operators, and the resulting theory is now now commonly
referred to as Fredholm theory. The following theorem is the main result
of Fredholm theory. We point out that Fredholm theory was extended to
Banach spaces, mainly by Riesz and Schauder. See [Br] or [Y] for excellent
expositions of this generalisation.
0 = λ lim wj − z.
j
λz − Az = λ lim(λwj − Awj ) = 0.
j
λ lim vj = g + ζ,
j
and
(λ I − A)((g + ζ)/λ) = lim (λ I − A)vj = lim gj = g.
j→∞ j→∞
Since Aλ∗ is not injective, Ker(A∗λ ) is not trivial, whence Ran(Aλ ) is a proper
(closed) subspace of H. Set Rjλ := Ran(Ajλ ).
We claim that
Rjλ ⊃ Rj+1
λ j = 1, 2, 3, . . . (proper inclusion). (2.6.4)
For otherwise there is an index j such that Rjλ = Rj+1λ . Then there are
j−1 j j
v ∈ Rλ \Rλ and w ∈ Rλ such that Aλ v = Aλ w. But this and the injectivity
of Aλ imply that v = w, which contradicts the fact that v ∈ Rj−1
λ \ Rjλ .
By (i), each Rjλ is a closed subspace of H, hence a proper closed subspace
of Rj−1 j j+1
λ . Therefore there exists a unit vector fj ∈ Rλ , orthogonal to Rλ .
S := A + Λ ◦ P.
λu − Au = 0 and Λ(P u) = 0.
λI − A = (λ − A∗ )∗ ,
prove that δ ≤ δ ∗ .
This concludes the proof of (v), and of the theorem. 2
(i) 0 (which may or may not be an eigenvalue) and a finite number (pos-
sibly 0) of nonzero eigenvalues with finite multiplicities;
An interesting fact, which we shall use in the sequel without any further
reference, is that all the results in this section hold, with different proofs, for
compact operators on Banach spaces (see the books [Br, Y] for the corre-
sponding proofs).
Aϕk = λk ϕk ,
76 CHAPTER 2. LINEAR OPERATORS ON BANACH SPACES
Step III: for each µ > 0, the set Eµ of all the eigenvalues λ of A
with |λ| > µ is finite, and for each λ ∈ Eµ the nullspace Ker(λ I − A)
is finite dimensional.
This is just Theorem 2.6.3 above, which holds for all compact operators.
For the sake of definiteness, assume that the first of the two equalities above
holds. Clearly, there exists a sequence {fk } ∈ H with kfk k = 1, such that
Then we take the limit of both sides as k tends to infinity, and obtain
From (2.7.2) and the fact that the left hand side is nonnegative, we deduce
that |||A||| = kgk and, consequently, that
Step V: conclusion.
We argue by reductio ad absurdum. Suppose that the closure S of the
span of all the eigenvectors of A is not dense in H. Then S ⊥ is a nontrivial
closed subspace of H, hence an Hilbert space itself. It is straightforward
to check that AS ⊆ S and that AS ⊥ ⊆ S ⊥ . Furthermore, A, restricted to
S ⊥ , is clearly a bounded self adjoint operator on S ⊥ . Therefore, by Step IV
(applied to A|S ⊥ ), it has at least one nonzero eigenvalue. This clearly would
be also an eigenvalue of the original operator A, which contradicts the fact
that S is the closure of the span of all the eigenvalues of A.
This concludes the proof of Step V, and of the theorem. 2
Note that 0 may or may not be an eigenvalue of a bounded self adjoint oper-
ator. Furthermore, if 0 is an eigenvalue, then the corresponding eigenspace
may be finite or infinite dimensional. The following examples should help
clarifying the situation.
78 CHAPTER 2. LINEAR OPERATORS ON BANACH SPACES
Example 2.7.7 Denote I the interval [0, 1], and by A : L2 (I) → L2 (I) the
operator defined by
Af (x) := x2 f (x) ∀x ∈ I.
79
80 CHAPTER 3. DIRICHLET VIA INTEGRAL EQUATIONS
and Z
C1 := sup |K(x, y)| dµ(x) < ∞.
y∈M M
We now compute the p norm of both sides (with respect to the variable x),
and obtain
Z Z 1/p
1/p0
kTK f kp ≤ C0 dµ(x) |K(x, y)| |f (y)|p dµ(y)
M M
1/p0 1/p
≤ C0 C1 kf kp ,
as required. 2
3.1. KERNELS OF TYPE α 81
(iv) there exist positive constants C0 and δ0 such that for each X ∈ ∂Ω there
exists j ∈ {1, . . . , N } with B4δ0 (X) ∩ ∂Ω ⊂ Uj , and
Φj Bδ (X) ∩ ∂Ω ⊆ Bδ Φj (X) ⊆ Φj BC0 δ (X) ∩ ∂Ω ∀δ ∈ (0, δ0 ],
equivalently
Bδ (X) ∩ ∂Ω ⊆ Ψj Bδ (Φj (X)) ⊆ BC0 δ (X) ∩ ∂Ω ∀δ ∈ (0, δ0 ].
Proof. For each point Z ∈ ∂Ω there is a local chart (UZ , ΦZ ) such that Z
is in UZ , the map ΦZ is an homeomorphism between UZ and Φ(UZ ) =: VZ
and its inverse ΨZ := Φ−1 2
Z is of class C (VZ ). By possibly shrinking the open
set UZ , and remaining the variables, we may assume that ΨZ is of the form
ξ 7→ (ξ, fZ (ξ)), where fZ is a function of class C 2 (VZ ), and that VZ is a ball
n−1
S
with centre 0 ∈ R . Clearly Z∈∂Ω UZ = ∂Ω. Since ∂Ω is compact, we can
choose a finite subcovering UZ1 , . . . , UZN , which clearly satisfies (i)-(iii). In
the sequel we shall write Uj instead of UZj .
Now we prove (iv). For each X ∈ ∂Ω and j ∈ {1, . . . , N }, set
(
0 if X ∈ / Uj
τj (X) :=
sup{r > 0 : Br (X) ∩ ∂Ω ⊆ Uj } if X ∈ Uj .
82 CHAPTER 3. DIRICHLET VIA INTEGRAL EQUATIONS
Then set
τ (X) := max{τ1 (X), . . . , τN (X)}
If τ (X) = τk (X), then we say that X is related to Uk , and write X ∼ Uk . .
Since {U1 , . . . , UN } is a covering of ∂Ω, each X belongs to at least to one of
the sets U1 , . . . , UN , so that τ (X) > 0 for every X ∈ ∂Ω. Define
τ := inf τ (X). (3.1.2)
X∈∂Ω
Exercise 3.1.4 Prove that the convex hull of a compact set in Rn is com-
pact.
This proves the required result for δ ≤ δ0 . Suppose now that δ > δ0 . Then
Z
|X − Y |−α dσ(Y )
Bδ (X)∩∂Ω
Z
≤ |X − Y |−α dσ(Y ) + δ0−α σ(∂Ω)
Bδ0 (X)∩∂Ω
ωn−1
≤ C0 δ0n−1−α + δ0−α σ(∂Ω) (3.1.12)
n − 1 −h α
ωn−1 i
≤ δ n−1−α C0 + δ 1+α−n δ0−α σ(∂Ω)
hn − ω
1−α i
n−1
≤ δ n−1−α C0 + δ01+α−n δ0−α σ(∂Ω)
n−1−α
n−1−α
= C1 δ ,
where
ωn−1
C1 := C0 + δ01−n σ(∂Ω). (3.1.13)
n−1−α
This concludes the proof of the lemma. 2
when α = 0;
(iii) if α ∈ [0, n − 1), then TK is a compact operator on L2 (∂Ω).
Proof. To prove (i), we apply Proposition 3.1.2. We shall prove the required
result in the case where α > 0. The case where α = 0 is left to the reader.
Observe that for every δ > 0
Z Z
|K(X, Y )| dσ(Y ) ≤ kAk∞ |X − Y |−α dσ(Y )
∂Ω ∂Ω
≤ kAk∞ C1 δ n−1−α ,
3.1. KERNELS OF TYPE α 85
as required.
Finally we prove (iii). For each ε > 0 we set
Dε := {(X, Y ) ∈ ∂Ω × ∂Ω : |X − Y | ≤ ε}.
Correspondingly, we define
The required conclusion then follows from the fact that the ideal of compact
operators is closed in L(L2 (∂Ω)) (see Exercise 2.5.4).
To prove (3.1.16) observe that TKε is the integral operator whose kernel
vanishes in Dεc and is equal to K(X, Y ) in Dε . By (ii),
and
lim K(Y, Z) − K(X, Z) = 0 ∀Z ∈ ∂Ω \ B2δ (X).
Y →X
TK f (Y ) − TK f (X)
Z
n−1−α
≤ kAk∞ kf k∞ C1 δ + kf k∞ K(Y, Z) − K(X, Z) dσ(Z).
∂Ω\B2δ (X)
By taking the infimum of both sides over all δ > 0, we obtain that
lim TK f (Y ) − TK f (X) = 0,
Y →X
so that
∞
−1
X TK 1 jε
u=λ g.
j=0
λ
Note that, by (i), each summand of the series above is a continuous function
on ∂Ω. Since
X∞
kTKj ε1 gk∞ X∞
|||TKε1 |||j
≤ kgk ∞ < ∞,
j=0
λj j=0
λj
(the last inequality follows from (3.1.19)) the sum of the series above is
continuous on ∂Ω, whence u is continuous on ∂Ω, as required to conclude
the proof of (ii), and of the proposition. 2
the series being convergent in L(B). Hint: prove that the series above is
convergent in L(B) by using the analogue
P∞ of the classical Weierstrass test for
series of elements in B. Then set B := j=0 Aj , and show that (I − A) B =
I = B (I − A).
1 (Y − x) · ν(Y )
∂ν(Y ) N (x − Y ) = ∀Y ∈ ∂Ω ∀x 6= Y. (3.2.1)
ωn |x − Y |n
Definition 3.2.1 The double layer potential with density g is the func-
tion Z
Dg(x) = g(Y ) ∂ν N (x − Y ) dσ(Y ) ∀x ∈ Rn \ ∂Ω.
∂Ω
Proof. The required property follows from the remark above by differentiat-
ing under the integral sign. The reader should check that differentiation is
permitted. 2
Now, we study Dg(X) when X ∈ ∂Ω. The difficulty here is that the ker-
nel of the double layer potential has a singularity on ∂Ω, so that even the
convergence of the integral
Z
g(Y ) ∂ν(Y ) N (X − Y ) dσ(Y )
∂Ω
must be proved for every X ∈ ∂Ω. As a first attempt, we examine the order
of magnitude of the kernel. By (3.2.1),
∂ν(Y ) N (X − Y ) ≤ C |X − Y |1−n ∀X 6= Y.
Observe that the vectors ∂1 Ψ(η), . . . , ∂n−1 Ψ(η) are tangent to ∂Ω at the
point Y , whence they are orthogonal to ν(Y ). Therefore
n−1
hX i
∂` Ψ(η) (ξ` − η` ) · ν(Y ) = 0, (3.2.2)
`=1
and
|(X − Y ) · ν(Y )|
= | Ψ(η) − Ψ(0) · ν(Y )|
n−1
1 X 2 2
≤ [∂ ψ1 (ζ1 ), . . . , ∂k,` ψn (ζn )]T · ν(Y ) (ξk − ηk ) (ξ` − η` )
2 k,`=1 k,`
2
≤ C max max sup |∂k,` ψh (ω)| |ξ − η|2
h=1,...,n k,`=1,...,n−1 ω∈κ
2
≤ C max max sup |∂k,` ψh (ω)| C02 |X − Y |2 .
h=1,...,n k,`=1,...,n−1 ω∈κ
Proof. We give only a sketch of the proof. The details are left to the interested
reader.
Clearly F is of class C 1 . Furthermore dF (X, 0) is nonsingular for every
X ∈ S. By the inverse function theorem, F can be inverted in a neighbour-
hood WX of the point (X, 0), yielding a map
FX−1 : WX → (S ∩ WX ) × − ε(X), ε(X)
(i)
Z 1
∀x ∈ Ω
K(x, Y ) dσ(Y ) = 1/2 ∀x ∈ ∂Ω
∂Ω
0 ∀x ∈ Ωc ;
(ii) Z
sup K(x, Y ) dσ(Y ) < ∞;
x∈Rn \∂Ω ∂Ω
Proof. First, we prove (i). Recall that the Newtonian potential with pole x
is harmonic in Rn \ {x}. Suppose that x ∈ Ωc . Then N (x − ·) is harmonic
in a neighbourhood of Ω, and
Z Z
∂ν(Y ) N (x − Y ) dσ(Y ) = ∆Y N (x − Y ) dV (Y ) = 0,
∂Ω Ω
as required.
Now suppose that x ∈ Ω. For ε small, consider the domain
Ωε := Ω \ Bε (x).
Then N (x − ·) is harmonic in a neighbourhood of Ω, and, by a standard
consequence of the divergence theorem
Z Z
∂ν(Y ) N (x − Y ) dσ(Y ) = ∆Y N (x − Y ) dV (Y ) = 0.
∂Ωε Ωε
Observe that
Z
∂ν(Y ) N (x − Y ) dσ(Y )
∂Ωε
Z Z
= ∂ν(Y ) N (x − Y ) dσ(Y ) − ∂ν(Y ) N (x − Y ) dσ(Y ).
∂Ω ∂Bε (x)
3.2. THE DOUBLE LAYER POTENTIAL 93
as required.
Finally suppose that x ∈ ∂Ω. For ε > 0, we define
and
∂Bε (x)00 := {y ∈ ∂Bε (x) : ν(x) · y < 0}.
Since Y 7→ ∂ν(Y ) N (x − Y ) is integrable (see Lemma 3.1.5),
Z Z
∂ν(Y ) N (x − Y ) dσ(Y ) = lim ∂ν(Y ) N (x − Y ) dσ(Y ).
∂Ω ε↓0 Sε
1 2n−1 1−n
∂ν(Y ) N (x − Y ) ≤ |x − Y |1−n ≤ δ ,
ωn−1 ωn−1
whence
2n−1 1−n
Z
∂ν(Y ) N (x − Y ) dσ(Y ) ≤ δ σ(∂Ω).
∂Ω ωn−1
(II) Suppose that d(x, ∂Ω) < δ/2. By Lemma 3.2.7, there exists a unique
point X ∈ ∂Ω such that
x = X + |x − X| ν(X). (3.2.4)
We write
Z
∂ν(Y ) N (x − Y ) dσ(Y )
∂ΩZ Z
= ∂ν(Y ) N (x − Y ) dσ(Y ) + ∂ν(Y ) N (x − Y ) dσ(Y ).
∂Ω\Bδ (X) Bδ (X)
(3.2.5)
We denote the integrals on the right hand side by J1 and J2 , respectively,
and estimate them separately.
To estimate the first, observe that if Y ∈
/ (Bδ (X) ∩ ∂Ω), then
|x − Y | ≥ |X − Y | − |X − x| ≥ δ/2,
so that
2n−1 1−n
∂ν(Y ) N (x − Y ) ≤ δ ,
ωn−1
and
2n−1 1−n
Z
∂ν(Y ) N (x − Y ) dσ(Y ) ≤ δ σ(∂Ω).
∂Ω\Bδ (X) ωn−1
Thus, to conclude the proof of (ii), it remains to estimate
Z
∂ν(Y ) N (x − Y ) dσ(Y ). (3.2.6)
Bδ (X)
3.2. THE DOUBLE LAYER POTENTIAL 95
1
|x − Y |2 ≥ |Y − X|2 + |x − X|2 .
(3.2.8)
2
Given the claim, we show how to conclude the proof of (ii). The claim and
(3.2.7) imply that there exists a constant C such that
ωn ∂ν(Y ) N (x − Y )
|x − X| |Y − X|2
≤C n/2 + C n/2
|Y − X|2 + |x − X|2 |Y − X|2 + |x − X|2
|x − X| 2−n
≤C n/2 + C |Y − X| .
|Y − X|2 + |x − X|2
|x − X|
Z Z
C n/2 dσ(Y )+C |Y − X|2−n dσ(Y ).
Bδ (X)∩∂Ω |Y − X|2 + |x − X|2 Bδ (X)∂Ω
x − Y = X − Y + |x − X| ν(X)
(∇fj (η), −1)
= ξ − η, fj (ξ) − fj (η) + p |x − X|
1 + |∇fj (ξ)|2
Then
2
|x − Y |2 = (ξ − η, fj (ξ) − fj (η)
2
+p |x − X| ∇fj (η)(ξ − η) − fj (ξ) + fj (η)
1 + |∇fj (ξ)|2
+ |x − X|2
Now,
2
≥ |ξ − η|2 ,
(ξ − η, fj (ξ) − fj (η)
and, by Taylor’s formula,
≤ C |ξ − η|2 .
∇fj (η)(ξ − η) − fj (ξ) + fj (η)
Therefore
2
p |x − X| ∇fj (η)(ξ − η) − fj (ξ) + fj (η)
1 + |∇fj (ξ)|2
≤ |x − X| sup |||Hfj (ω)||| |ξ − η|2 .
ω∈κj
Therefore
Z
lim |J | ≤ 2ε sup |K(x, Y )| dσ(Y ) = C ε.
x→X,x∈Ω x∈Rn ∂Ω
We have used (ii) in the last equality. By taking the infimum of both sides
with respect to ε > 0, we obtain that
lim |J | = 0,
x→X,x∈Ω
Definition 3.3.1 The single layer potential with density g is the func-
tion Z
Sg(x) := g(Y ) N (x − Y ) dσ(Y ) ∀x ∈ Rn \ ∂Ω.
∂Ω
Proof. The required properties follow from the remark above by differentiat-
ing under the integral sign. The reader should check that differentiation is
permitted. 2
x−Y
Z
1
∂ν (Sϕ)(x) = · ν(X) ϕ(Y ) dσ(Y ) ∀x ∈ V \ ∂Ω. (3.3.1)
ωn ∂Ω |x − Y |n
3.3. THE SINGLE LAYER POTENTIAL 99
and
∂ν + (Sϕ)(X) := lim+ ∂ν (Sϕ)(X + tν(X)), (3.3.4)
t→0
whenever the limits above exist.
Some of the main properties of the single layer potential are summarised in
the next proposition.
100 CHAPTER 3. DIRICHLET VIA INTEGRAL EQUATIONS
(i) Sϕ is continuous in Rn ;
1
∂ν − (Sϕ)(X) = − ϕ(X) + TK ∗ ϕ(X) (3.3.5)
2
and
1
∂ν + (Sϕ)(X) = ϕ(X) + TK ∗ ϕ(X). (3.3.6)
2
Furthermore, the convergence of ∂ν (Sϕ)(X + tν(X)) to ∂ν − (Sϕ)(X)
as t tends to 0− and of ∂ν (Sϕ)(X + tν(X)) to ∂ν + (Sϕ)(X) as t tends
to 0+ is uniform on ∂Ω.
u(x) − u(X)
Z
1
≤ |x − Y |2−n |ϕ(Y )| dσ(Y )
(n − 2) ωn Bδ (X)∩∂Ω
Z
1
+ |X − Y |2−n |ϕ(Y )| dσ(Y )
(n − 2) ωn Bδ (X)∩∂Ω
Z
1
+ |X − Y |2−n − |X − Y |2−n |ϕ(Y )| dσ(Y ).
(n − 2) ωn ∂Ω\Bδ (X)
lim J3 = 0, (3.3.8)
x→X,x∈Ω
Then
x − Y = ζ − Y − |x − ζ| ν(ζ).
We distinguish between two cases.
(I) If Y ∈ Bδ (X) ∩ ∂Ω \ Bδ/5 (ζ), then
Thus,
Z
|x − Y |2−n |ϕ(Y )| dσ(Y ) ≤ C kϕk∞ δ 2−n σ Bδ (X) ∩ ∂Ω
Bδ (X)∩∂Ω
|x − Y | ≥ c |ζ − Y |.
Therefore
Z Z
2−n
|x − Y | |ϕ(Y )| dσ(Y ) ≤ C kϕk∞ |ζ − Y |2−n dσ(Y )
Bδ/5 (ζ)∩∂Ω Bδ/5 (ζ)∩∂Ω
≤ C kϕk∞ δ.
Thus,
|J3 | ≤ C kf k∞ δ. (3.3.9)
By combining (3.3.9), (3.3.7), and (3.3.8), we see that for all δ > 0 small
enough
lim |u(x) − u(X)| ≤ C kf k∞ δ.
x→X,x∈Ω
Next we prove the first formula in (ii). The proof of the second formula
is similar and is omitted. Set
1
Et (X) := ∂ν (Sϕ)(X + tν(X)) + ϕ(X) − TK ∗ ϕ(X).
2
We must show that
lim sup Et (X) = 0. (3.3.10)
t→0− X∈∂Ω
102 CHAPTER 3. DIRICHLET VIA INTEGRAL EQUATIONS
1
ϕ(X)
2
X −Y X + t ν(X) − Y i
Z h
1
= − · ν(Y ) ϕ(X) dσ(Y )
ωn ∂Ω |X − Y |n |X + t ν(X) − Y |n
X −Y X + t ν(X) − Y i
Z h
1
= − · ν(Y ) [ϕ(X) − ϕ(Y )] dσ(Y )
ωn ∂Ω |X − Y |n |X + t ν(X) − Y |n
X −Y X + t ν(X) − Y i
Z h
1
+ − · ν(Y ) ϕ(Y ) dσ(Y )
ωn ∂Ω |X − Y |n |X + t ν(X) − Y |n
(3.3.11)
Furthermore
X −Y
Z
1
TK ∗ ϕ(X) = · ν(X) ϕ(Y ) dσ(Y ) (3.3.12)
ωn ∂Ω |X − Y |n
and
X + t ν(X) − Y
Z
1
∂ν (Sϕ)(X + tν(X)) = · ν(X) ϕ(Y ) dσ(Y )
ωn ∂Ω |X + t ν(X) − Y |n
(3.3.13)
Therefore
Et (X)
X −Y X + t ν(X) − Y i
Z h
1
= − · ν(Y ) [ϕ(X) − ϕ(Y )] dσ(Y )
ωn ∂Ω |X − Y |n |X + t ν(X) − Y |n
X −Y X + t ν(X) − Y i
Z h
1
− − · [ν(X) − ν(Y )] ϕ(Y ) dσ(Y )
ωn ∂Ω |X − Y |n |X + t ν(X) − Y |n
(3.3.14)
t 2
We denote the two integrals above by It (X) and It (X), respectively.
Fix ε > 0, and choose δ > 0 such that
for all X and Y in ∂Ω such that |X − Y | < 2δ. This is possible for both ν
and ϕ are continuous, hence uniformly continuous, functions on the compact
set ∂Ω.
We estimate It1 (X). We write
Z Z
1
It (X) = + .
B2δ (X)∩∂Ω ∂Ω\B2δ (X)
3.4. SOLVABILITY 103
Observe that
Z Z
≤ ε ωn K(X, Y ) + K(X + tν(X), Y ) dσ(Y )
B2δ (X)∩∂Ω ∂Ω
Z (3.3.16)
≤ 2 ε ωn sup K(x, Y ) dσ(Y ),
x∈Rn ∂Ω
where K(x, Y ) is the kernel of the double layer potential. Recall that, by
Proposition 3.2.8 (ii), the supremum above is finite.
Next, we apply the mean value theorem, and obtain that
X −Y X + t ν(X) − Y
sup n
− n
≤ C |t| δ −n ∀t ∈ (−δ, 0),
Y ∈∂Ω\B2δ (X) |X − Y | |X + t ν(X) − Y |
By combining (3.3.16) and (3.3.17), we obtain that for every ε > 0 there
exists δ > 0 such that
whence
lim sup It1 (X) ≤ C ε.
t→0− X∈∂Ω
Therefore
lim sup It1 (X) = 0.
t→0− X∈∂Ω
The last two formulae imply (3.3.10), as required to conclude the proof of
(ii), and of the proposition. 2
and Z
HR := Sf (X) ∂ν(X) (Sf )(X) dσ(X).
∂BR (0)
We claim that limR↑∞ HR = 0. Indeed, for all x such that d(x, ∂Ω) ≥ |x|/2
kf k∞
Z
|Sf (x)| ≤ |x − Y |2−n dσ(Y )
(n − 2) ωn ∂Ω
kf k∞
≤ d(x, ∂Ω)2−n σ(∂Ω)
(n − 2) ωn
kf k∞
≤ 2n−2 |x|2−n σ(∂Ω).
(n − 2) ωn
Similarly, for all x such that d(x, ∂Ω) ≥ |x|/2
kf k∞
Z
|∂ν(X) Sf (x)| ≤ |x − Y |1−n dσ(Y )
ωn ∂Ω
kf k ∞
≤ 2n−2 |x|1−n σ(∂Ω).
ωn
Thus,
kf k2∞
|HR | ≤ 22(n−2) R3−2n σ(∂Ω)2 ωn Rn−1
(n − 2) ωn2
kf k2∞
= 22(n−2) R2−n σ(∂Ω)2 ,
(n − 2) ωn
which tends to 0 as R tends to infinity (we are assuming n ≥ 3), as claimed.
We claim that limt↓0 Jt = 0. This follows from Proposition 3.3.5 (i)-(ii).
We leave the verification of this fact to the reader.
The two claims above and (3.4.1) imply that I = 0. Therefore |∇(Sf )| =
0 a.e. on Rn \ Ω, hence |∇(Sf )| = 0 everywhere on Rn \ Ω, for |∇(Sf )| is
continuous on Rn \ Ω. Since Rn \ Ω is connected, Sf is constant on Rn \ Ω.
We have already proved that lim|x|→∞ Sf (x) = 0, whence Sf = 0 on Rn \ Ω.
By (i), Sf is continuous on Rn , so that Sf = 0 on ∂Ω. By the maximum
principle for Sf on Ω, it follows that Sf = 0 on Rn . Thus,
∂ν − (Sf ) = 0 on ∂Ω.
By (ii),
1
∂ν − (Sf )(X) = − f (X) + TK ∗ f (X) ∀X ∈ ∂Ω.
2
Therefore
1 1
f (X) = f (X) + TK ∗ f (X) − − f (X) + TK ∗ f (X)] = 0,
2 2
106 CHAPTER 3. DIRICHLET VIA INTEGRAL EQUATIONS
1
thereby proving the injectivity of 2
I + TK ∗ , and concluding the proof of the
lemma. 2
107
109
(i) does a solution to the minimization problem (4.0.2) solve the Dirichlet
problem (4.0.1)?
(ii) does the solution to the the Dirichlet problem (4.0.1) solve the mini-
mization problem (4.0.2)?
D(u) ≤ D(w) ∀w ∈ Yg .
In fact, the proposition above does not establish the existence of a solution to
the Dirichlet problem (4.0.1). Rather, it converts (4.0.1) into the problem of
minimizing the Dirichlet integral, under the assumption that g is the “trace”
on the boundary of some functions in C 2 (Ω).
In order to avoid interruptions in the flow of information, we postpone
all the proofs of the results stated in this introduction to sections below (see
Subsection 4.0.3 for details).
Weierstrass pointed out that it was not clear (and had not been proved)
that a minimizing function for the Dirichlet integral exists, so there might
simply be no winner to the implied competition in Proposition 4.0.1. He ar-
gued by analogy with a simpler one-dimensional problem: that of minimizing
the integral Z 1
2
I(ϕ) := xϕ0 (x) dx
−1
1
amongst all C functions ϕ on [−1, 1] that satisfy ϕ(−1) = −1 and ϕ(1) = 1.
Exercise 4.0.2 Prove that the infimum of I is equal to 0. Prove that the
infimum is not a minimum.
This suggests the possibility that the functional D may reach its infimum in
a set of competitors larger than Yg , but possibly not in Yg , and demands a
proof that a minimizer, in fact, exists. After all, the problem
min {x2 : x ∈ Q, x2 ≥ 2}
is a norm on C 2 (Ω), where the latter is the space of all functions ϕ in C 2 (Ω)
such that Dα ϕ admits a continuous extension to Ω for each multiindex α
with |α| ≤ 2.
Definition 4.0.5 We define H01 (Ω) to be the completion of Cc∞ (Ω) with
respect to the Sobolev norm k·kH 1 (Ω) .
Since Cc∞ (Ω) is a subspace of C 2 (Ω), the space H01 (Ω) is a (closed) subspace
of H 1 (Ω). Since the restriction of the trace operator γ to Cc∞ (Ω) vanishes, the
restriction of γ to H01 (Ω) vanishes as well. Thus, we may say that elements
of H01 (Ω) have “vanishing boundary values”. The converse statement is also
true. For the proof, see [AF, Thm 5.37].
The following result states that, under some mild conditions, the classical
solution and the solution to the minimization problem (4.0.6) agree.
We shall not prove this result. Its proof is a consequence of some deep
estimates of the oscillation of solutions on balls close to the boundary ∂Ω.
The interested reader is referred to [GT, Ch. 8], especially Corollary 8.28 and
the comments at the beginning of p. 206.
Now we discuss briefly the second question posed at the beginning of
this subsection. We show that the assumption that the boundary datum g
113
In conclusion, there are good reasons to introduce the so-called weak deriva-
tives of L2 functions.
Hence
d
D(u + εϕ)|ε=0
dε Z
= ∇ϕ · ∇u + ∇u · ∇ϕ dV
ΩZ Z
=− ϕ div(∇u) + div(∇u) ϕ dV + ϕ ∂n u + (∂n u) ϕ dσ
ZΩ ∂Ω
=− ϕ div(∇u) + div(∇u) ϕ dV :
Ω
and Z
d
D(u + iεψ)|ε=0 = 2i ψ ∆(Im u) dV ∀ψ ∈ Re Y0 .
dε Ω
Now, if (i) holds, then ∆(Re u) = 0 = ∆(Im u), and the two formulae above
imply that u is a critical point of D, i.e. (ii) holds.
Conversely, if (ii) holds, then
a.e., hence in the classical sense, for u is of class C 2 (Ω). Therefore (i) holds.
116
Next, trivially (iii) implies (ii). To conclude the proof of the proposition,
we need to show that (ii) implies (iii). Observe that
D(w) = D(u + w − u)
Z
= D(u) + ∇u · ∇(w − u) + ∇(w − u) · ∇u dV + D(w − u).
Ω
(4.1.1)
The middle term of the right hand side vanishes. Indeed, and by integrating
by parts Z
∇u · ∇(w − u) + ∇(w − u) · ∇u dV
Ω Z
=− ∆u (w − u) + (w − u) ∆u dV
Ω
= 0;
note that there are no boundary terms arising from the integration by parts,
because w − u is in Y0 , hence it vanishes on the boundary of Ω. Furthermore,
we already know that (ii) is equivalent to (i), whence ∆u = 0, thereby
justifying the last equality above.
Therefore, by (4.1.1),
Now observe that ν and div ν are uniformly bounded in Rn . Therefore there
exists a constant C such that
Z Xn Z
2
|u (∂j u)| + |(∂j u) u| + |u|2 dV
|u| dσ ≤ C
∂Ω j=1 Ω
h Xn i
2
(by Schwartz’s inequality) ≤C 2 kukL2 (Ω) k∂j ukL2 (Ω) + n kukL2 (Ω)
j=1
γ(u) := u|∂Ω .
By Exercise 2.2.4, the map γ extends uniquely to a bounded linear map from
H 1 (Ω) to L2 (∂Ω), for C 2 (Ω) is dense in H 1 (Ω) (see Theorem 5.5.12 below).
This concludes the proof of the proposition. 2
agrees with gα on the boundary of the unit disc in the plane, is harmonic in
the (open) unit disc (the series can be differentiated term by term as long
118
An elementary computation shows that the right hand side is infinite for all
α ≤ 1/2. Thus, for these values of α the classical solution to the Dirichlet
problem with boundary datum gα has infinite Dirichlet integral. 2
Chapter 5
then ∞
X
µ(E) = µ(Ej ),
j=1
the series being absolutely convergent. Note that µ is not defined on un-
bounded Borel sets. It may be shown that, given a complex measure µ,
there exist four nonnegative measures µj , j = 1, . . . , 4, such that
4
X
µ(E) = µj (E).
j=1
119
120 CHAPTER 5. DISTRIBUTIONS AND SOBOLEV SPACES
Thus, for every complex measure µ we have defined a complex linear func-
tional T µ on Cc (Rn ), defined by
Z
(T µ)(ϕ) = ϕ dµ ∀ϕ ∈ Cc (Rn ).
Rn
lim g(x) = 0.
x→∞
P
If g is chosen so that j g(j) = ∞, then the right hand side in the preceding
formula is equal to +∞, thereby contradicting the fact that (T µ)(g) is a
complex number.
Thus, a different topology on Cc (Rn ) is needed. A basic result we shall
use is the following classical result of Riesz. The reader is referred to [Ru,
Ch. 6] for the proof of a slight generalisation thereof. Recall that if X is
a locally compact Hausdorff space, we denote by C0 (X) the completion of
Cc (X) with respect to the uniform norm.
for each f in C0 (X). Moreover, ||F || = |ν|(X), where |ν|(X) denotes the
total variation of the measure ν.
|ν(E)| ≤ |ν|(E)
122 CHAPTER 5. DISTRIBUTIONS AND SOBOLEV SPACES
for every measurable set E. In fact, |ν| is the minimal solution to the problem
of finding a positive measure µ that dominates ν, in the sense that
|ν(E)| ≤ µ(E)
S
where E = j Ej , the measurable sets Ej are pairwise disjoint, and the
supremum is taken with respect to all partitions of E as countable union of
pairwise disjoint measurable subsets of E. For these facts, and many more,
the reader is referred to [Ru, Ch. 6].
where Bj := B(0, j), and then consider the countable family %B j of the
associated seminorms, which we shall denote by %j for the sake of simplicity.
5.1. CONTINUOUS FUNCTIONS AND MEASURES 123
ψj (x) = ψ(x − xj ) ∀x ∈ Rn .
Note that the sequence {%0j (ϕ)} is decreasing. For each decreasing sequence
ε := {εj } of positive real numbers we consider the set
V(f ) := f + V(0).
We claim that τSI is strictly finer than τU C. On the one hand, it is
clear that for each j the neighbourhood U %j ; ε of 0 in the τU C topology
(see (5.1.2)) contains the neighbourhood Vε of 0 in the τSI topology, where
ε = {2−j ε : j = 0, 1, 2, . . .}. Thus, τSI is finer than τU C .
On the other hand, the neighbourhood
(ii) there exists ` such that the support of {fN } is contained in B ` and {fN }
is uniformly convergent to f .
≥ fNj (xnj )
= εnj+1 ,
Corollary 5.1.6 Every Cauchy sequence for the topology τSI on Cc (Rn ) is
convergent to an element of Cc (Rn ).
Proof. Let {fj } be a Cauchy sequence with respect to τSI . Since τSI is
finer than τU C , {fj } is a Cauchy sequence with respect to the topology τU C .
Therefore {fj } is a Cauchy sequence with respect to each of the seminorms
%k . In particular, the restrictions of the functions fj to B k is a Cauchy
sequence in C(B k ) for every k. Since C(B k ) is complete with respect to the
uniform norm, there exists a function gk in C(B k ) such that
g(x) = gk (x) ∀x ∈ B k .
as claimed.
This contradicts the fact that {fN } is convergent to g. Hence we have
proved that the support of g is compact.
The proof that {fN } is convergent to g in the τSI topology is straightfor-
ward and is left to the reader. 2
is continuous on Cc (Rn ), τSI .
We need to show that for every ε > 0, the set
Z
−1
n o
(T µ) Bε (0) = ϕ ∈ Cc (Rn ) : ϕ dµ < ε
Rn
(here Bε (0) denotes the open disc in the complex plane with centre 0 and
radius ε) contains a neighbourhood of the origin in the strict inductive limit
topology, i.e., there exists a decreasing sequence η := {ηj } such that
(T µ)(ϕ) < ε ∀ϕ ∈ Vη .
5.1. CONTINUOUS FUNCTIONS AND MEASURES 127
< ε,
as required.
as required. 2
(ii) there exists a sequence {fN } of functions in Cc∞ (Ω) such that supp (fN ) ⊂
VN +1 and fN = 1 on V N .
The role played above by the space Cc (Rn ) will now be played by the
space Cc (Ω). The relevant topology on Cc (Ω) will be the strict inductive
limit topology τSI (Ω), which is defined similarly to the strict inductive
limit topology τSI on Cc (Rn ), with the role of the seminorms %0j played now
by the seminorms ρ0j , defined by
Note that DK (Ω) and DK (Rn ) are just the same space.
(i) if K0 and K1 are two disjoint compact sets in Rn , show that the con-
tinuous function
d(x, K0 )
η(x) := ∀x ∈ Rn
d(x, K0 ) + d(x, K1 )
(ii) show that it suffices to prove the result for functions g satisfying 0 ≤
g ≤ 1;
5.2. SMOOTH FUNCTIONS AND DISTRIBUTIONS 131
and 0 ≤ Fj ≤ (1/3)(2/3)j−1 on Rn ;
(v) set F = j Fj . Prove that F is continuous on Rn and F|K = g;
P
∞
Observe that τSI (Ω) is not the metrizable topology induced by the countable
family of seminorms %m,KN (here KN = V N ). It is a finer topology, which
induces on DKN (Ω) the topology introduced in Definition 5.2.1.
∞
We might argue that τSI (Ω) may depend on the chosen exhaustion {KN }
of Ω, but, in fact, it is straightforward to prove that two different exhaustion
give rise to the same topology. Henceforth, for every domain Ω we fix once
and for all an exhaustion, and all seminorms are associated to that particular
exhaustion.
(ii) there exists ` such that the support of fN is contained in V ` for ev-
ery N and for every mullti-index α the sequence {Dα fN } is uniformly
convergent to Dα f on V ` .
Proof. The proof is similar to the proof of Proposition 5.1.5, and it is left as
an exercise. 2
∞
Corollary 5.2.6 Every Cauchy sequence for the topology τSI (Ω) on D(Ω) is
convergent to an element of D(Ω).
Proof. The proof is similar to the proof of Corollary 5.1.6, and it is left as
an exercise. 2
It is a natural question to ask whether there are simple criteria which imply
∞
that a linear functional on D(Ω) is also continuous (with respect to the τSI
topology). The following theorem gives a handy characterisation of distribu-
tions.
∞
(i) T is a distribution (i.e., T is continuous with respect to the τSI topol-
ogy);
(ii) for every compact set K ⊂ Ω, there exist a constant CK and a number
mK such that
Proof. Suppose first that T ∈ D0 (Ω). Then for every ε > 0 there exist
sequences η and m such that
%0h,Ω\Vj (ϕ) = 0 ∀j ≥ ` ∀h ∈ N,
≤ max max Dα ψe
|α|≤mj K
Dα ψ
= η` max max
|α|≤mj K %m` ,K (ψ)
< η`
< ηj ,
the last inequality being a consequence of the facts that η is decreasing and
j < `. Therefore
ε > ψ,e T = η` ψ, T ,
%m` ,K (ψ)
134 CHAPTER 5. DISTRIBUTIONS AND SOBOLEV SPACES
whence
ε
ψ,
eT < %m` ,K (ψ),
η`
We must show that T is a distribution on Ω, i.e., that for every ε > 0 there
exist two sequences η and m such that
and
∞
X
1= ω` (x) ∀x ∈ Ω.
`=0
2−`−1 ε
η` < .
CV `+2 C`+2
5.2. SMOOTH FUNCTIONS AND DISTRIBUTIONS 135
Proof. First we prove (i). Suppose that f is in Lp (Ω), and consider the linear
functional Tf on D(Ω), given by
Z
hϕ, Tf i = ϕ f dλ ∀ϕ ∈ D(Ω),
Ω
Of course we need to check that the right hand side of the formula above
defines a distribution, i.e., the linear functional ϕ 7→ − h∂j ϕ, T i is continuous
for the strict inductive limit topology of D(Ω). By Theorem 5.2.8, ∂j T is con-
tinuous if and only if for every compact set K ⊂ Ω, there exist a constant CK
and a nonnegative integer mK such that
Therefore
hϕ, ∂j T i = h∂j ϕ, T i
≤ CK %m0K ,K (∂j ϕ)
≤ CK %m0K +1,K (ϕ) ∀ϕ ∈ DK (Ω),
so that (5.3.1) is satisfied with mK = m0K + 1, as required.
It is clear that we can iterate the reasoning above, and define Dα T for
every distribution T . This will be the distribution defined by
Example 5.3.2 Consider the absolute value function T on the real line
T (x) = |x| ∀x ∈ R.
Clearly T is differentiable in the classical sense off the origin with derivative
T 0 (x) = sgn(x),
hϕ, T 0 i = − hϕ0 , T i
Z ∞
(T is a loc. int. function) =− ϕ0 (x) |x| dx
−∞
Z 0 Z −∞
0
= ϕ (x) x dx − ϕ0 (x) x dx
−∞ 0
Z 0 Z −∞
(by parts) =− ϕ(x) dx + ϕ(x) dx
−∞ 0
Z ∞
= ϕ(x) sgn(x) dx ∀ϕ ∈ D,
−∞
as claimed.
H 0 = δ0
hϕ, H 0 i = − hϕ0 , Hi
Z ∞
(H is a loc. int. function) =− ϕ0 (x) dx
0
= ϕ(0)
= hϕ, δ0 i ∀ϕ ∈ D,
as required. Similarly, for every positive integer k ≥ 2 one can prove that
(k)
ϕ, δ0 = (−1)k ϕ(k) (0) ∀ϕ ∈ D.
Example 5.3.4 Consider the distribution T , which agrees with the locally
integrable function x 7→ x−1/2 1(0,∞) (x). We want to compute T 0 (in the sense
of distributions). Clearly, T 0 is not the function x 7→ −(1/2) x−3/2 1(0,∞) (x),
5.3. DERIVATIVES OF DISTRIBUTIONS 139
hϕ, T 0 i = − hϕ0 , T i
Z ∞
(T is a loc. int. function) =− ϕ0 (x) x−1/2 dx
0
Z ∞
= − lim ϕ0 (x) x−1/2 dx
ε↓0 ε
1 ∞
h Z i
−1/2 −3/2
(by parts) = − lim − ϕ(ε) ε − ϕ(x) x dx
ε↓0 2 ε
1 ∞ ϕ(x) − ϕ(ε)
Z
= − lim dx
ε↓0 2 ε x3/2
1 ∞ ϕ(x) − ϕ(0)
Z
=− dx ∀ϕ ∈ D.
2 0 x3/2
and that Z Z
ϕ(x) 1
dx ≤ kx ϕk∞ dx
I1c x I1c x2
≤ 2R kϕk∞ .
Furthermore
Z Z 1
ϕ(x) ϕ(x) − ϕ(−x)
lim+ dx = lim+ dx
ε→0 I1 \Iε x ε→0 ε x
Z 1 Z x
1
= lim+ ϕ0 (s) ds dx
ε→0 ε x −x
0
≤ 2 kϕ k∞ ∀ϕ ∈ D[−R,R] (R).
Hence
|hϕ, T i| ≤ CR (kϕ0 k∞ + kϕk∞ ) ∀ϕ ∈ D[−R,R] (R).
Therefore the restriction of T to D[−R,R] (R) is continuous, as required.
Exercise 5.3.8 Compute the second derivative (in the sense of distribu-
tions) of the distribution T , which agrees with the locally integrable function
x 7→ x−1/2 1(0,∞) (x).
in the sense of distributions. Here sn denotes the surface measure of the unit
sphere.
2
Exercise 5.3.16 Denote by p f the distribution which agrees on R \{(0, 0)}
2 2
with the function (x, y) 7→ x + y . Compute the partial derivatives of f
up to the second order, and, then the Laplacian of f .
P (D)E = δ0 .
for all ϕ ∈ Cc∞ (Rn ) such that 0 does not belong to the support of ϕ.
Now suppose that 0 belongs to the support of ϕ. Then
hϕ, ∆N i = h∆ϕ, N i
Z
= ∆ϕ N dV
Rn Z
= lim ∆ϕ N dV.
ε↓0 Rn \Bε (0)
The last equality follows from the local integrability of N near the origin and
the Lebesgue dominated convergence theorem. Suppose that the support of ϕ
5.4. FUNDAMENTAL SOLUTIONS 143
where ν denotes the outward unit normal to ∂Bε (0). By arguing as in the
proof of Theorem 1.5.5, we see that
Z Z
ϕ ∂ν N dσ → ϕ(0) and that ∂ν ϕ N dσ → 0
∂Bε (0) ∂Bε (0)
for all ϕ, whose support contains 0. This, together with (5.4.1), proves the
required result. 2
Exercise 5.4.5 Prove that the function E(x) := (1/2) |x| is a fundamental
solution of the operator d2 / dx2 (with pole 0).
Exercise 5.4.7 Prove that for every complex number c the function
e−c|x|
uc (x) := ∀x ∈ R3 \ {0}
4π|x|
and Z
lim ∂ν ϕ(Y ) N (x − Y ) dσ(Y ) = 0.
ε↓0 ∂Bε (x)
Therefore
h(∆ϕ) ∗ N, f i = hϕ, f i .
The formula above, together with (5.4.3), shows that ∆(N ∗ f ) = f in the
sense of distributions, as required.
Next we prove (ii). We already know that N ∗ f is a smooth function
on Rn . It remains to prove that
∆(N ∗ f )(x) = f (x) ∀x ∈ Rn .
By (i), ∆(N ∗ f ) = f in the sense of distributions. Since both ∆(N ∗ f )
and f are locally integrable functions (for they are continuous), they must
be equal a.e. Since both ∆(N ∗ f ) and f are continuous, they must agree at
every point, as required. 2
hϕ, u0 i = − hϕ0 , ui
Z
= − ϕ0 (x) u(x) dx
ZI
(because u = v a.e.) = − ϕ0 (x) v(x) dx
Z I
(by parts, for ϕ and v are functions in C 1 (I)) = ϕ(x) v 0 (x) dx
I
= hϕ, v 0 i ,
f (x + h) − f (x)
→ g(x)
h
pointwise a.e.. The following example illustrate this statement.
hϕ, F 0 i = − hϕ0 , F i
Z
= − ϕ0 (x) F (x) dx
ZR Z x
0
=− dx ϕ (x) dµF (y)
R −∞
By Fubini’s theorem,
Z Z ∞
0
hϕ, F i = − dµ (y) F
ϕ0 (x) dx
R y
Z
= ϕ(y) dµF (y)
R
= ϕ, µF .
5.5. SOBOLEV SPACES 149
The following result shows that there are at least five equivalent definitions
of L2 -derivative of a function in L2 (Rn ).
as h tends to 0;
(iv) there exists a sequence {fn } of test functions such that
kf − fn kL2 (Rn ) → 0
and {∂1 fn } is a Cauchy sequence in L2 (Rn );
(v) there exists g in L2 (Rn ) such that
Z Z
− ∂1 ϕ f dV = ϕ g dV ∀ϕ ∈ Cc1 (Ω).
Rn Rn
Proof. We prove only that (i) and (v) are equivalent. The rest of the proof
is left to the interested reader.
Suppose that (i) holds. Then
Z Z
ϕ ∂1 f dV = − ∂1 ϕ f dV ∀ϕ ∈ D. (5.5.1)
Rn Rn
(5.5.2)
=− (∂1 ϕ) ∗ ψε f dV.
Rn
150 CHAPTER 5. DISTRIBUTIONS AND SOBOLEV SPACES
Note that
ϕ ∗ ψε − ϕ 2
→0 and (∂1 ϕ) ∗ ψε − ∂1 ϕ 2
→0
Since this holds for every ϕ ∈ Cc1 (Rn ), (v) is proved, with g = ∂1 f . Con-
versely, if (v) holds, then, in particular,
− h∂1 ϕ, f i = hϕ, gi ∀ϕ ∈ D.
We shall use this equivalence without any further comment in the sequel.
Exercise 5.5.6 Prove directly that the function sgn on the real line does
not admit a weak first derivative.
5.5. SOBOLEV SPACES 151
Note that this norm is associated to the following natural inner product in
H k (Ω)
XZ
(u, v) := Dα u Dα v dλ.
|α|≤k Ω
More generally, for each p in [1, ∞] we denote by W k,p (Ω) the space of all
complex valued functions u in Lp (Ω) whose distributional derivatives up to
the order k belong to Lp (Ω), endowed with the norm
X Z 1/p
kukW k,p (Ω) = |Dα u|p dλ
|α|≤k Ω
Clearly W k,2 (Ω) = H k (Ω). In the rest of these notes we focus on the case
where p = 2.
To state the second result we need the following definition. Suppose that α
is a number in (0, 1]. We say that a function u ∈ C k (Ω) is in C k,α (Ω) if for
every multiindex β of order k, the derivatives Dβ u are Hölder continuous of
order α on Ω.
(iv) if m > n/2, then H m (Ω) ,→ C k,α (Ω); here k := m − [n/2] − 1, and, if
n is odd, then α := m − n/2 − k, whereas, if n is even, then α is any
number in (0, 1).
Remark 5.5.13 By Theorem 5.5.12 (i), the Sobolev space H m (Ω) may be
equivalently defined as the completion of C ∞ (Ω) with respect to the norm
(5.5.3).
5.5. SOBOLEV SPACES 153
Remark 5.5.14 For a generic domain Ω, the space C ∞ (Ω) is not dense
in H k (Ω), as the following example in one dimension shows. Denote by
Ω the set (−1/2, 0) ∪ (0, 1/2) in R. We argue by contradiction. Suppose
that C ∞ (Ω) is dense in H 1 (Ω). Denote by f the function, which is equal
to −1 on (−1/2, 0) and equal to 1 on (0, 1/2). Clearly f belongs to H 1 (Ω)
and kf kH 1 (Ω) = kf kL2 (Ω) , because the distributional derivative of f vanishes.
Suppose that {ϕn } is a sequence of functions in C ∞ ([−1/2, 1/2]) that is
convergent to f in H 1 (Ω). Then {ϕn } is convergent to f in L2 (Ω), hence
lim kϕn kL2 (Ω) = kf kL2 (Ω) = 1.
n→∞
whence Z 1/2
|ϕn (x)| ≤ |ϕ0n | dλ
−1/2
Exercise 5.5.16 Prove directly that H 1 ((−1, 1)) ⊂ C([−1, 1]). Hint: use
the fundamental theorem of calculus (for the Lebesgue integral).
154 CHAPTER 5. DISTRIBUTIONS AND SOBOLEV SPACES
Chapter 6
Proof. Since H01 (Ω) is the completion of Cc∞ (Ω), it suffices to show that the
required estimate holds for all ϕ ∈ Cc∞ (Ω).
The one dimensional case has already been proved in Proposition 2.2.6.
155
156 CHAPTER 6. DIRICHLET L2
Note that we cannot hope to control the L2 norm of a generic smooth function
with the L2 norm of its gradient (just take a nonzero constant function on an
interval of the real line). Thus, the assumption that ϕ vanish at some point
in Ω is really needed.
√
Observe that D is a norm on H01 (Ω). Indeed, observe that if u ∈ H01 (Ω),
then
D(u) = 0 ⇒ u is constant a.e. on each connected component of Ω
⇒ u = 0 a.e. on Ω,
Proof. The left hand inequality follows directly from the definition of kukH 1 (Ω) ;
the right hand inequality is a consequence of Poincaré’s inequality in the pre-
ceding lemma. 2
6.2. SOLUTION TO THE MINIMIZATION PROBLEM 157
and (
∆u2 = 0 in Ω
u2 |∂Ω = 0.
Clearly, u2 = 0, thereby showing that u is real valued.
Thus, when trying to solve the Dirichlet problem with real data, it suffices
to look for solutions in a real Banach space. Note that a fundamental role in
the reasoning below is played by the fact that ∆ preserves the class of real
valued functions, This may fail for more general operators.
158 CHAPTER 6. DIRICHLET L2
We follow a recent expository article of N.H. Friedel [Fr]. Note that there
are no restrictions on ∂Ω in the next statement. An estimate of the unique
solution to the minimization problem (4.0.6) is obtained by just setting f = 0
in the next statement.
v := w − G.
Observe that
Z
1 1
J(v + G) = D(v) + Re ∇v · ∇G dV + D(G) − (v, f )L2 − (G, f )L2
2 Ω 2
= J(v)
e + const
where Z
e := 1 D(v) +
J(v) ∇v · ∇G dV − (v, f )L2 , (6.2.3)
2 Ω
and const = 21 D(G) − (f, G)L2 . Thus, w solves the above minimization
problem if and only if v minimizes the functional J,
e subject to the condition
v ∈ Re H01 (Ω).
we have used Poincaré’s inequality for the domain Ω in the second inequal-
ity above. Then, by Riesz’s representation theorem, there exists a unique
element vΛf,G ∈ H01 (Ω), with |||Λf,G |||H01 = kvΛf,G kH01 , that represents the
functional Λf,G , i.e.
Since Λf,G is real on Re H01 (Ω), the element vΛf,G must be real valued, i.e.,
vΛf,G is in Re H01 (Ω).
Now, the functional Je on Re H01 (Ω) may be rewritten as follows
which is obviously equal to −vΛf,G . Thus, the minimizer of the original func-
tional J is u := G − vΛf,G . To conclude the proof of the theorem, it remains
to estimate the norm of u in terms of the data. The triangle inequality and
(6.2.5) imply that
kukH01 ≤ kGkH01 + kvΛf,G kH01
≤ 2kGkH01 + δΩ kf kL2 ,
as required. 2
which associates to the pair (f, G) the unique solution S(f, G) of the mini-
mization problem with data (f, G). The proof of Theorem 6.2.1 above shows
that
S(f, G) = G − vΛf,G , (6.2.6)
160 CHAPTER 6. DIRICHLET L2
where vΛf,G is the unique function in Re H01 (Ω) that represents the continuous
linear functional
Z
Λf,G (v) = ∇v · ∇G dV − (v, f )L2 .
Ω
G − H − vΛ0,G−H = G − H − (G − H) = 0,
as required.
Thus, if fact, the solution map (f, G) 7→ S(f, G) does not depend on the
element G in Re H 1 (Ω), but on the coset G + Re H01 (Ω) in the coset space
Re H 1 (Ω)/ Re H01 (Ω). Therefore there exists a unique map Se : Re L2 (Ω) ×
Re H 1 (Ω)/ Re H01 (Ω) → Re H 1 (Ω) such that the following diagram
Re L2 (Ω) × Re H 1 (Ω)
Q
Q
Q
Q
I ×π Q
Q
S
Q
Q
? Q
s
Q
2 1
Re L (Ω) × Re H (Ω)/ Re H01 (Ω) - Re H 1 (Ω)
Se
Now, we know that the minimization problem (6.2.1) subject to the con-
dition (6.2.2) has a unique solution. We would like to relate the minimizimg
6.2. SOLUTION TO THE MINIMIZATION PROBLEM 161
This result suggest to study the Banach dual of H01 (Ω) in some detail. This
will be done in Section 6.3.
We conclude this section will the following result, which goes in the di-
rection of interpreting the minimizing function of the problem (6.2.8) as a
solution of a differential problem associated to the Laplace operator.
(ii) for each ϕ ∈ Re H01 (Ω), the function v satisfies the equation
Z
∇ϕ · ∇v dV = F (ϕ). (6.2.9)
Ω
as required.
6.3. THE DUAL OF H01 (Ω) 163
(ii) for every bounded linear functional Λ on H01 (Ω) there exist f and gj in
L2 (Ω), j ∈ {1, . . . , n}, such that
Z Z
Λ(u) = u f dV + u div G dV ∀u ∈ H01 (Ω).
Ω Ω
Observe that
D n
X E n
X
ϕ, f + ∂j gj ≤ |hϕ, f i| + h∂j ϕ, gj i
j=0 j=0
n
X
≤ kϕkL2 kf kL2 + k∂j ϕkL2 kgj kL2
j=0
h n
X i
≤ C kϕkH01 kf kL2 + kgj kL2 ∀ϕ ∈ D(Ω).
j=0
as required.
164 CHAPTER 6. DIRICHLET L2
Observe that
(ϕ, vΛ )H01 = (∇ϕ, ∇vΛ )L2
= − hϕ, div ∇vΛ i
= − hϕ, ∆vΛ i ∀ϕ ∈ D(Ω).
We have proved that
Definition 6.3.2 The dual of H01 (Ω), endowed with the operator norm,
will be denoted by H −1 (Ω).
Note that the Dirac delta at 0 is in H −1 ((−1, 1)), for it is the distributional
derivative of the Heaviside function.
Exercise 6.3.4 Prove that ∇1B1 (0) is in H −1 (B1 (0)) for all n. Which is
the support of ∇1B1 (0) ?
We conclude this section with the following result, which we shall use later.
Proof. First we show that ∆ is a continuous map from H01 (Ω) to H −1 (Ω).
This is implicit in the proof of Theorem 6.3.1 (ii). However, for the sake of
clarity, we repeat the argument here. Since H01 (Ω) is included in D0 (Ω), ∆u
is a distribution on Ω. For every ϕ in D(Ω)
Here the space H01 (Ω) is endowed with the norm kukH01 (Ω) := k|∇u|kL2 (Ω) ,
which is equivalent to the H 1 (Ω)-norm, as already mentioned. We now take
the supremum of both sides with respect to all test functions ϕ such that
kϕkH01 (Ω) ≤ 1, and obtain
Hence ∆ : H01 (Ω) → H0−1 (Ω) with operator norm at most one.
We claim that ∆ has closed range. Indeed, suppose that {gk } is a Cauchy
sequence in Ran(∆). Denote by uk the unique (for ∆ is injective) function
in H01 (Ω) such that ∆uk = gk . By (6.3.4) we have that
for all j and k. Hence {uk } is a Cauchy sequence in H01 (Ω). Since H01 (Ω)
is complete, there exists u in H01 (Ω) such that kuk − ukH01 (Ω) tends to 0 as
j and k tend to infinity. Since ∆ is continuous from H01 (Ω) to H −1 (Ω), it
follows that
∆u = lim gk ,
k→∞
∆u = f.
Furthermore,
kukH01 (Ω) = kf kH −1 (Ω) .
Proof. Denote by ∆−1 : H −1 (Ω) → H01 (Ω) its inverse (which, by Lemma 6.3.5,
is a continuous linear map). Then, given f in H −1 (Ω), then u := ∆−1 f in
H01 (Ω) and ∆u = f . Thus, u is the unique solution in H01 (Ω) of the Dirichlet
problem associated to the Poisson equation.
The required norm estimate is a consequence of the fact, established
in Lemma 6.3.5, that ∆ is an isometric isomorphism between H01 (Ω) and
H −1 (Ω). 2
∆u = f (6.4.1)
6.4. WEAK SOLUTIONS AND REGULARITY 167
Notice that both integrals above are absolutely convergent, because ϕ van-
ishes near the boundary of Ω. We can integrate by parts in the left hand side
and get Z Z
− ∇ϕ(x) · ∇u(x) dx = ϕ(x) f (x) dx. (6.4.2)
Ω Ω
The key observation here is that we do not need to assume that u is in C 2 (Ω)
to give meaning to the left hand side of (6.4.2). For instance, it suffices to
assume that u and its distributional derivatives ∂j u are in L1loc (Ω). Obviously
this is a much weaker requirement than u belongs to C 2 (Ω).
Proof. We prove this result under the additional assumption that the distri-
butional solution u be locally integrable. For the proof of the general case,
see [Me, Thm 3.3.14 and Corollary 3.3.15].
168 CHAPTER 6. DIRICHLET L2
Therefore
sup u ∗ ϕε (x) − u ∗ ϕε0 (x)
x∈Ω0
Z
n
≤ n sup (u ∗ ϕε )(x + y) − (u ∗ ϕε0 )(x + y) dV (y)
r ωn x∈Ω0 Br (0)
n
≤ n sup (u ∗ ϕε )(x + ·) − (u ∗ ϕε0 )(x + ·) L1 (Br (0))
r ωn x∈Ω0
n
≤ n u ∗ ϕε − u ∗ ϕε0 L1 (Ω0 +Br (0))
r ωn
→0
as ε and ε0 tend to 0. Thus, {u ∗ ϕε } is convergent uniformly in Ω0 as ε tends
to 0. By (6.4.4), its uniform limit must be u. Hence u is continuous. We
then deduce from (6.4.3) that for r small enough
Z
n
u(x) = n u(x + y) dV (y) ∀x ∈ Ω0 .
r ωn Br (0)
Therefore u is harmonic in Ω0 by Theorem 1.3.9. 2
Bibliography
[AF] R.A. Adams and J.J.F. Fournier, Sobolev spaces, Pure and Ap-
plied Mathematics Series, vol. 140, Academic Press, IInd edition.
[Fo1] G.B. Folland, Real analysis. Modern techniques and their appli-
cations, Second edition, J. Wiley & Sons, 1999.
169
170 BIBLIOGRAPHY