DDDD
DDDD
Small print These notes are a digest of much more complete notes by M. S. Joshi and
A. J. Wassermann which are also being issued for this course. I should very much
appreciate being told of any corrections or possible improvements and might even part
with a small reward to the first finder of particular errors. This document is written
in LATEX2e and stored in the file labelled ~twk/IIB/PDE.tex on emu in (I hope) read
permitted form. My e-mail address is twk@dpmms.
Contents
1 Introduction 2
1.1 Generalities . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.2 The symbol . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
3 Distributions 13
3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
3.2 The support of a distribution . . . . . . . . . . . . . . . . . . 17
3.3 Fourier transforms and the Schwartz space . . . . . . . . . . . 18
3.4 Tempered distributions . . . . . . . . . . . . . . . . . . . . . . 20
1
4.5 Existence of the fundamental solution . . . . . . . . . . . . . . 29
5 The Laplacian 31
5.1 A fundamental solution . . . . . . . . . . . . . . . . . . . . . . 31
5.2 Identities and estimates . . . . . . . . . . . . . . . . . . . . . 32
5.3 The dual Dirichlet problem for the unit ball . . . . . . . . . . 35
9 References 47
1 Introduction
1.1 Generalities
When studying ordinary (i.e. not partial) differential equations we start
with linear differential equations with constant coefficients. We then study
linear differential equations and then plunge into a boundless ocean of non-
linear differential equations. The reader will therefore not be surprised if
most of a first course on the potentially much more complicated study of
partial differential equations should limit itself essentially to linear partial
differential equations.
A linear partial differential equation is an equation of the form P u = f
where u and f are suitable functions from Rn to R and
X
P = aα (x)∂ α .
|α|≤k
2
and
α ∂ |α|
∂ u= .
∂xα1 1 ∂xα2 2 . . . ∂xαnn
|α|
Sometimes we write ∂ α u = ∂ ∂xu .
Although the first part of the course will deal with ‘first order’ linear
partial differential equations without restrictions on the coefficients (and in-
deed even with slightly more general partial differential equations) the main
part of the course will deal with linear partial differential equations with
constant coefficients. The main tools used will be Laurent Schwartz’s theory
of distributions and the Fourier transform.
The fact that we do not deal with non-linear equations does not mean that
they are not important. The equations of general relativity are non-linear
and at a more mundane level the Navier-Stokes equation of fluid dynamics
∂u
− ∆u + u.∆u = f − ∆p, ∆.u = 0
∂t
is non-linear. P
We call P = |α|≤k aα (x)∂ α a differential operator (or just an ‘operator’).
We look at three such operators in detail. The first which must be most
studied non-trivial differential operator in mathematics is the Laplacian ∆
known in more old fashioned texts as ∇2 and defined by
n
X ∂ 2u
∆u =
j=1
∂x2j
3
∂
obtained by replacing ∂xj
by iξj so that
X
σ(P ) = p(x, ξ) = aα (x)(iξ)α = e−ix.ξ P (eix.ξ ).
|α|≤k
Note that x and ξ are n dimensional vectors and that we use the convention
n
α
Y
yα = yj j .
j=1
To see where the symbol comes from, observe that taking Fourier transforms
Notice that if P has degree k and Q degree l then the principal symbol of
P Q is given by the simple formula
If the principal symbol is never zero or only vanishes to first order then
the operator is said to be of principal type. In more advanced work it is
shown that when the operator is of principal type the lower order terms have
4
little effect on the qualitative behaviour of the associated partial differential
equation.
We define the characteristic set to be the subset of Rn × Rn where the
principal symbol vanishes.
Traditionally second order operators which behaved like the Laplacian were
called elliptic, those which behaved like the wave operator were called hy-
perbolic and those that behaved like the heat operator were called parabolic.
The distinction is very useful but the reference to conic sections is not.
Example 1.2 (Example 1). (i) We could have considered complex valued u
in place of real valued u. If we do this the operator
∂ ∂
P = +i
∂x ∂y
has principal symbol
σ1 (P )(x, y : ξ, η) = iξ − η
and so is elliptic.
(ii) The Laplacian ∆ is elliptic but the wave operator 2 and the heat
operator J are not.
5
2 Ordinary differential equations
2.1 The contraction mapping theorem
Hadamard introduced the idea of a well posed problem into the study of
partial differential equations. According to Hadamard a well posed problem
must have a solution which exists, is unique and varies continuously on the
given data. Without going too deeply into the matter we may agree that
these are reasonable matters to investigate.
Thus given a partial differential equation with boundary conditions we
shall study the following problems.
Existence Can we show that there is a solution in a neighbourhood of a
given point? Can we show that there exists a solution everywhere?
Uniqueness Is the solution unique?
Continuity Does u depend continuously on the boundary conditions?
Does u depend continuously on on other elements of the problem?
Smoothness How many times is u differentiable? Does u have points of
singularity? Does the solution blow up in some way after a finite time?
We may illustrate these ideas in the case of ordinary differential equa-
tions. Some of the material in this section will be familiar from examination
questions on the first analysis course in 1B but you should consider it all
with the exception of the contraction mapping theorem (Theorem 2.1) itself
to be examinable.
In 1B we proved Banach’s contraction mapping theorem.
Theorem 2.1 (Theorem 1). Let (X, d) be a complete non-empty metric space
and T : X → X a map such that d(T x, T y) ≤ k(x, y) for all x, y ∈ X and
some k with 0 ≤ k < 1. Then there exists a unique x0 ∈ X such that
T x0 = x0 . If x ∈ X then T n x → x0 as n → ∞.
This can be strengthened as follows.
Theorem 2.2 (Corollary 1). Let (X, d) be a complete non-empty metric
space and T : X → X a map. Suppose further that there exists an integer
N ≥ 1 and a k with 0 ≤ k < 1 such that d(T N x, T N y) ≤ k(x, y) for all
x, y ∈ X. Then there exists a unique x0 ∈ X such that T x0 = x0 . If x ∈ X
then T n x → x0 as n → ∞.
Note that these theorems not only give a fixed point but also give a
method for finding it.
For the rest of section 2.1 f will be a function from R × Rn . Let
E = {t ∈ R : |t − t0 | ≤ a} × {x ∈ Rn : ∥x − x0 ∥ ≤ b}.
6
We assume that f satisfies the Lipschitz condition
on E.
Exercise 2.3. (i) Show that a function that satisfies a Lipschitz condition is
continuous.
(ii) Show that any continuously differentiable function satisfies a Lipschitz
condition.
(iii) Show that a function that satisfies a Lipschitz condition need not be
differentiable everywhere.
(iv) Show that the function f considered above is bounded on E.
We set M = sup(t,x)∈E ∥f (t, x)∥ and h = min(a, bM −1 ).
Theorem 2.4. If f is as above the differential equation
dx
= f (t, x), x(t0 , 0) = x0 ⋆
dt
has a unique solution for |t − t0 | ≤ h.
Example 2.5. (Here x : R → R.)
(i) The differential equation
dx
= 0, x(0) = 0, x(1) = 1
dt
has no solution.
(ii) Show that the differential equation
dx 2
= x 3 , x(0) = 0
dt
has at least two solutions.
From now on we move away from 1B.
Theorem 2.6 (Theorem 3). The solution of ⋆ in Theorem 2.4 depends
continuously on x0 . More formally, if we define T : R → C([t − h, t + h]) by
taking T y to be the solution of
dx
= f (t, x), x(t0 ) = y
dt
and give C([t − h, t + h]) the uniform norm then T is continuous.
7
In the special case of linear ordinary differential equations we can give a
rather strong perturbation result.
Theorem 2.7 (Theorem 4). We use the standard operator norm on the
space L = L(Rn , Rn ) of linear maps. Suppose that A, B : R × Rn → L are
continuous and that M ≥ ∥A(t, x)∥, ∥B(t, x)∥ for all t and x.
dξ
(t, x) = A(t, x)ξ(t, x), ξ(t0 , x) = a(x)
dt
dη
(t, x) = B(t, x)η(t, x), η(t0 , x) = b(x)
dt
then if |t − t0 | ≤ K
f : (−ϵ, ϵ) × U → Rn
α : (−δ, δ) × U0 → U
8
Thus if x is fixed αx (t) = α(t.x) is an integral curve with starting point x.
In some sense a flow is a collection of integral curves.
We note but shall not prove that the smoother the vector field the smoother
the associated flow.
d
α(t, x) = f (t, α(t, x)), α(0, x) = x.
dt
then α ∈ C k [1 ≤ k].
Lemma 2.10. With the notation just established there exists an ϵ > 0 such
that
αt+s (u) = αt (αs (u))
for all |t|, |s|, |u| < ϵ.
αt+s = αt ◦ αs
will hold whenever it is reasonable for it to hold and in particular if the flow
is determined for all time we have a group action. This kind of group action
is called a dynamical system (see Course O6).
9
In fact the method works for the slightly more general semi-linear first order
partial differential equation
n
X ∂u
aj (x) = f (x, u). (*)
j=1
∂xj
we have ( )
n
X
char(L) = (x, ξ) : aj (x)ξj = 0 .
j=1
10
differential operator L which we are considering, S is non-characteristic if at
each x ∈ S the normal ζ to S satisfies
n
X
aj ζj ̸= 0.
j=1
Theorem 2.13 (Theorem 6). Locally, there is a unique solution of (∗) with
u(x, 0) given on a non-characteristic hypersurface S.
Working through a few examples will make the theory much clearer.
∂u ∂u
+ 2x = u2
∂x ∂y
(a1 , a2 , . . . , an , b)
11
on Rn+1 and solve for the integral curves
dxs
= a(xs (t), ys (t)) xs (0) = g(s)
dt
dys
= b(xs (t), ys (t)) ys (0) = ϕ(s).
dt
Our solution is basically y(s, t) but we want it as a function of x not (s, t).
The map
(s, t) 7→ x(s, t)
will have an invertible derivative at t = 0 provided that the vector
(s, t) 7→ x(s, t)
Theorem 2.15 (Theorem 7). Equation (***) has a unique solution locally.
Once again, working through a few examples will make the theory much
clearer.
∂u ∂u
u + =1
∂x ∂y
subject to u = s/2 on x = y = s.
If a semi-linear equation (with real coefficients) has real right hand side
it can be solved by the technique of this subsection but if the right hand side
is complex we must use the previous method.
12
3 Distributions
3.1 Introduction
Distribution theory is a synthesis of ideas coming from partial differential
equations (e.g. weak solutions), physics (the Dirac delta function) and func-
tional analysis but the whole is much more than the sum of its parts.
We shall need the notion of the support of a continuous function.
13
Definition 3.4. Suppose that fj ∈ D for each j and f ∈ D. We say that
fj → f
D
fn → f implies T fn → T f.
D
T f = ⟨t, f ⟩.
Lemma 3.5. The set D′ is a vector space if we use the natural definitions
we have Tg ∈ D′ .
and say that every continuous function is a distribution. (Hence the name
‘generalised function’ sometimes given to distributions.)
The second insight is more in the nature of a recipe than a theorem.
14
and that At : D → D is a linear map which is continuous in the sense that
fn → f implies At fn → At f. (2)
D D
Suppose in addition
⟨Ag, f ⟩ = ⟨g, At f ⟩ (3)
Then, if S is any distribution we can define a distribution TA S by the equation
⟨TA S, f ⟩ = ⟨S, At f ⟩.
The mapping TA : D′ → D′ is linear and whenever g ∈ D we will have
Ag = TA g.
We usually write TA = A, so that
⟨AT, f ⟩ = ⟨T, At f ⟩.
Exercise 3.8. Although we shall not use this in the course, convergence in
D′ is defined as follows. If Sj [j ≥ 1] are distributions we say that Tj →′ S if
D
⟨Sn , f ⟩ → ⟨S, f ⟩
for all f ∈ D. Show that if A satisfies the conditions of Lemma 3.7, the map
TA : D′ → D′ is continuous in the sense that
Sj →′ S implies TA S →′ TA S.
D D
15
Here is another example along the same lines.
A∗ f (x) = f (Ax).
If A is invertible we can define a transpose map (A∗ )t by (A∗ )t = | det A|−1 (A−1 )∗
i.e. by
((A∗ )t (f ))(x) = | det A|−1 f (T −1 x).
If T is a distribution we can define an associated distribution (i.e. obtain
a change of variables result for distributions) by following Lemma 3.7 to
obtain
⟨A∗ T, f ⟩ = ⟨T, | det A|−1 (A−1 )∗ f ⟩.
[Joshi and Wasserman think in terms of ‘signed areas’ and have det A
where we have | det A| in accordance with the conventions of applied math-
ematics.]
As an indication that we are on the right road, the next example shows
consistency with the Dirac deltas of mathematical methods in 1B.
⟨δa , f ⟩ = f (a)
τb δa = δa+b .
(iii) We have
⟨∂xj δa , f ⟩ = −(∂xj f )(a)
for all f ∈ D.
(iv) If we work on R and define the Heaviside function H : R → R by
H(t) = 1 for t ≥ 0,
H(t) = 0 otherwise
16
3.2 The support of a distribution
We start with a definition.
Definition 3.12. If T ∈ D′ then the support supp T of T is defined as
follows. A point x ∈
/ supp T if we can find an open set U such that x ∈ U
and whenever f ∈ D is such that supp f ⊆ U we have ⟨T, f ⟩ = 0.
Exercise 3.13. (i) Check that if f is a continuous function its support as
a continuous function is the same as its support when it is considered as a
distribution.
(ii) Show that if T is a distribution supp T is closed.
The following theorem is extremely important as is the method (partition
of unity) by which we obtain it.
Theorem 3.14. If f ∈ D and T ∈ D′ then
supp T ∩ supp f = ∅ implies ⟨T, f ⟩ = 0.
In my opinion the partition of unity is more an idea rather than a theorem
but here is one form expressed as a theorem.
Exercise 3.15 (Theorem 12). Let K be aScompact subset of Rn and U1 ,
U2 , . . . Um open sets in Rn such that K ⊆ Uj . Then
P we can find fj ∈ D
with 0 ≤ fi (x) ≤ 1 for all x, supp fj ⊆ Uj and
P fj (x) = 1 for x ∈ K,
fj (x) ≤ 1 everywhere.
When looking at Theorem 3.14 you should keep the following important
example in mind.
Exercise 3.16. We work in R.
(i) Show that there exists an f ∈ D with f (0) = 0, f ′ (0) = 1.
(ii) Observe that f (x) = 0 when x ∈ supp δ0 but ⟨δ0′ , f ⟩ = −1.
(iii) Why does this not contradict Theorem 3.14?
The support tells us where a distribution lives. A related concept is the
singular support.
Definition 3.17. If T ∈ D′ then the singular support singsupp T of T is
defined as follows. A point x ∈
/ singsupp T if we can find an f ∈ D such that
x∈/ supp T − f .
A careful argument using partitions of unity shows that if singsupp T = ∅
then T is in fact a smooth function.
When working with distributions it is helpful to recall the following re-
sults.
17
Lemma 3.18. If T is a distribution and ∂j T = 0 for all j then T = c where
c is a constant.
(We shall not prove this plausible result but like much else it may be
found in Friedlander’s beautiful little book, see section 2.1 [2].)
Lemma 3.19 (Theorem 28, A version of Taylor’s theorem). If f ∈ C ∞ (R)
and f (0) = f ′ (0) = · · · = f m−1 (0) = 0 then we can find g ∈ C ∞ (R) such that
f (x) = xm g(x)
for all x.
There is no difficulty (apart from notation) in extending Lemma 3.19 to
higher dimensions. In the next section we shall make use of the following
special case.
Lemma 3.20. If f ∈ C ∞ (Rn ) and f (0) = 0 then we can find gj ∈ C ∞ (Rn )
such that X
f (x) = gj xj
for all x.
As an example of the use of Lemma 3.19 we do the following exercise.
Exercise 3.21. Show that the general solution of
xT = 0
is T = cδ0 .
S(Rn ) = {f ∈ C ∞ (Rn ) : sup(1 + ∥x∥)m |∂ α f (x)| < ∞ for all m ≥ 0 and α}.
x
18
If f ∈ S(Rn ) then the Fourier transform
n/2 Z
ˆ 1
f (ξ) = e−ix.ξ f (x) dx
2π Rn
x 7→ xj f (x).
It is clear that
Dj : S(Rn ) → S(Rn )
xj : S(Rn ) → S(Rn )
D
d ˆ
j f = ξj f
x
d j f = −Dj f
ˆ
Lemma 3.22 can be used to give a neat proof of the Fourier inversion
formula for S. We use the following lemma.
Lemma 3.23 (Lemma 4). If T : S(Rn ) → S(Rn ) is a linear map which
commutes with xj and Dj for all j then, writing I for the identity map,
T = cI
for all x.
19
We can now prove our inversion theorem.
Theorem 3.25. If R : S(Rn ) → S(Rn ) is the map given by Rf (x) = f (−x)
then
F 2 = R.
Observe that this means that F 4 = I so F is invertible. Stating our result
in a more traditional but equivalent manner we have the following theorem.
Theorem 3.26 (Theorem 9). The Fourier transform
f 7→ fˆ
is an isomorphism of S onto itself with inverse given by
n/2 Z
1
f (x) = eix.ξ fˆ(ξ) dx.
2π Rn
20
We can now say that T ∈ S ′ to be a linear map T : D → C which is
continuous in the sense that
fn → f impliesT fn → T f.
S
T f = ⟨T, f ⟩.
and that
⟨Fg, f ⟩ = ⟨g, Ff ⟩ (3)
for any f, g ∈ S. (Thus, in the language of Lemma 3.7, F t = F.) Thus,
applying the result corresponding to Lemma 3.7 if T ∈ S ′ we can define FT
by the equation
⟨FT, f ⟩ = ⟨T, Ff ⟩,
or in the obvious corresponding notation
F 2 = R.
D
dj T = ξj T̂
j T = −Dj T̂
xd
21
However we do not have to develop the theory of S ′ and D′ separately
since D sits nicely in S and so S ′ sits nicely in D′ .
Theorem 3.30. (i) We have D ⊆ S. Moreover if that fj ∈ D for each j
and f ∈ D then
fj → f implies fj → f
D S
′
(ii) We may consider any element T ∈ S as an element of D in a natural
manner.
Theorem 3.31 (Theorem 13). (i) Given any f ∈ S we may find a sequence
fj ∈ D such that
fj → f
S
as j → ∞.
(ii) If T ∈ F and ⟨T, f ⟩ = 0 whenever f ∈ D then ⟨T, f ⟩ = 0 whenever
f ∈ S.
(iii) Let T, S ∈ F. If T and S are unequal when considered as members
of F they remain unequal when considered as members D.
Thus we may, and shall consider S ′ as a subspace of D′ . When we talk
about distributions we shall mean members of D′ . When we talk about
tempered distributions we shall mean members of S ′ .
Example 3.32 (Example 6). The delta function δ0 is a tempered distribu-
tion. Its Fourier transform δˆ0 is the constant function (2π)−n/2 .
The constant function 1 is a tempered distribution. Its Fourier transform
1̂ is (2π)n/2 δ0 .
The way we have defined T̂ by Lemma 3.7 ensures that
i.e. n/2 Z
1
fˆ(ξ) = e−ix.ξ f (x) dx.
2π Rn
We shall also need the following simple result which the reader may al-
ready have met.
22
|f (x)| dx < ∞ then fˆ is continuous
R
Lemma 3.34. If f is continuous and
and bounded.
for some ϵ > 0 then the Fourier transform and Fourier inverse transform of
q is C l .
4.1 Convolution
Unfortunately the notion of convolution of distributions is not as ‘clean’ as
the rest of the theory but it is essential.
Recall that if f and g are ‘nice functions’ (e.g. if they are both Schwartz
functions) we define the convolution f ∗ g by
Z
f ∗ g(x) = f (x − t)g(t) dt.
Rn
∗ 1 = 1̂1̂ = (2π)n δ0 δ0
1d
an ‘equation’ in which neither side makes sense (at least within distribution
theory as developed here). In general we can not multiply two distributions
and we can not convolve two distributions.
Having said this, there are many circumstances in which it is possible
to convolve two distributions. To see this observe that under favourable
23
circumstances, if f , g and h are all nice functions
Z Z
⟨f ∗ g, h⟩ = f (s − y)g(y) dy h(s) ds
Rn
Z Z
= f (s − y)g(y)h(y) dy ds
Z Z
= f (s − y)g(y)h(s) ds dy
Z Z
= f (x)g(y)h(x + y) dx, dy
f ⊗ g(x, y) = f (x)g(y).
Lemma 4.1. With the notation and hypotheses of the previous paragraph,
S ⊗ T is a well defined member of D(Rm+n ).
Exercise 4.2. Prove the two statements made in the last sentence.
24
So far we have only allowed elements of D′ to operate on elements of D
so we appear to be stuck.
However, under certain circumstances, elements of D′ will operate on
more general C ∞ functions in a natural manner.
Lemma 4.3. (i) Suppose T ∈ D′ (Rn ) and f ∈ C ∞ (Rn ). If supp f ∩ supp T
is compact then there is a natural definition of ⟨T, f ⟩.
(ii) Suppose T ∈ D′ (Rn ) and fj ∈ C ∞ (Rn ). Suppose further that there
exists a compact set K such that K ⊇ supp fj ∩ supp T for all j. Then if
∂ α fj → 0 uniformly on K for each α it follows that ⟨T, fj ⟩ → 0 as j → ∞.
We have arrived at our goal.
Theorem 4.4. If S, T ∈ D′ (Rn ) and supp S or supp T is compact then
we may define S ∗ T as follows. If h ∈ D(Rn ) define h̃ : R2n → C by
h̃(x, y) = h(x + y), the expression ⟨S ⊗ T, h̃⟩ is well defined in the sense of
Lemma 4.3 so we set
⟨S ∗ T, h⟩ = ⟨S ⊗ T, h̃⟩
The map S ∗ T so defined is a distribution.
Lemma 4.5. With the notation and hypotheses of Theorem 4.4
δ0 ∗ T = T.
25
(iv) If T and S are distributions at least one of which has compact support
then
∂ α (T ∗ S) = (∂ α T ) ∗ S = T ∗ (∂ α S).
Note that this gives ∂ α u = ∂ α δ0 ∗ u, so applying a differential operator is
equivalent to convolving with some distribution.
One of the key facts about convolution is that it smooths.
Lemma 4.7. If T is a distribution and f an m times continuously dif-
ferentiable function and at least one of the two has compact support then
⟨Ty , f (x − y)⟩ is a function and, as distributions,
∗ T (ξ) = Tb(ξ)S(ξ).
S[ b
P (D)T = δ0 .
26
The interest of fundamental solutions is shown by the next lemma.
Lemma 4.11. We use the notation of Definition 4.9
(i) If T is a distribution of compact support and E is a fundamental
solution of p(D) then S = E ∗ T is a solution of P (D)S = T .
(ii) If f is an m times continuously differentiable function of compact
support and E is a fundamental solution of P (D) then S = E ∗ f is an m
times continuously differentiable solution of P (D)S = T .
Example 4.12. (i) We continue with Example 4.10. Suppose f is a contin-
uous function of compact support and we wish to solve
u′ (x) = f (x).
Applying Lemma 4.11 we obtain
Z ∞ Z x
u(x) = c f (t) dt + f (t) dt.
−∞ −∞
(ii) Particular interest attaches to the two cases c = 0 when the solution
is Z x
u(x) = f (t) dt (A)
−∞
and c = −1 when the solution is
Z ∞
u(x) = − f (t) dt (B)
x
since equation (A) produces a solution for u′ (x) = f (x) valid whenever
f (x) = 0 for x large and positive and equation (B) produces a solution for
u′ (x) = f (x) valid whenever f (x) = 0 for x large and negative.
One of the points to note about Example 4.12 (ii) is that it shows the
advantage of extending the notion of convolution as far as as we can.
Example 4.13. If S and T are distributions on R such that we can find a
real number R such that
supp S, supp T ⊆ (−R, ∞).
Explain in as much detail as you consider desirable why we can define S ∗ T .
It also shows that it may be desirable to consider different fundamental
solutions for different kinds of problems. We can think of fundamental so-
lutions as inverses on different classes of functions in accordance with the
formula
E ∗ P (D) = P (D) ∗ E = δ0 .
27
4.3 Our first fundamental solution
If E is a fundamental solution of P (D) then by definition
P (D)E = δ0
−u′′ + u = δ0
and verify by direct calculation that you have indeed got a solution.
A large part of the theory of partial differential equations consists of
different ways of deal with the zeros of P (ξ).
28
We have
P (D)E = δ0 + f
where f ∈ S(Rn ). Also
singsupp E ⊆ {0}.
Lemma 4.17. If P (D) is elliptic and U a neighbourhood of 0 then we can
find a tempered distribution E with supp E ⊆ U , singsupp E ⊆ {0}, and
P (D)E = δ0 + f
where f ∈ S(Rn ).
We call a distribution like E in Theorem 4.16 or Lemma 4.17 a parametrix.
It will, no doubt, have occured to the reader that if we solve a partial
differential equation
P (D)u = f
in the distributional sense what we get is a distribution u and not necessarily
a classical function even if f is a classical function. The next theorem removes
this problem in certain cases.
Theorem 4.18 (Theorem 15, special case of Weyl’s lemma). If u ∈ D′ and
P (D) is an elliptic differential operator then u and P (D)u have the same
singular support. In particular if P (D)u is smooth so is u.
29
We prove this result by contour pushing so we need a result on analyticity
and growthR
With the proof of Theorem 4.20 we come to the end of our general account
of linear constant coefficient partial differential equations. For the rest of the
course we shall study the Laplacian, the heat and the wave operator. We
close this section by looking at the heat and wave operator in the context of
Theorem 4.20.
∂ X ∂2
P (D) = −
∂t ∂x2j
|P (τ + ia, ξ)| ≥ −a
|P (τ + ia, ξ)| = 0.
∂2 X ∂2
P (D) = −
∂t2 ∂x2j
for τ ∈ R, ξ ∈ Rn .
Note that our shifting contour argument gives one fundamental solution
for the heat operator but two for the wave equation.
30
5 The Laplacian
5.1 A fundamental solution
The reader can probably guess a fundamental solution for the Laplacian.
We shall proceed more slowly in order to introduce some useful ideas. Ob-
serve first that by Theorem 4.18 any fundamental solution E will have
singsupp E ⊆ {0}. Since the Laplacian is rotationally invariant an averaging
argument shows (as one might fairly confidently guess) that there must be a
rotationally invariant fundamental solution.
Exercise 5.1. We could now proceed as follows. Suppose f (x) = g(r) with
r = ∥x∥. Show that
1 d n−1
(∆f )(x) = r f g(r).
rn−1
dr
If f is smooth and ∆f = 0 on Rn \ {0} find f .
Instead of following the path set out in Exercise 5.1 we use another nice
idea — that of homogeneity. The following remark is due to Euler.
Lemma 5.2. Let f : Rn \ {0} → R be continuously differentiable and m be
an integer. The following statements are equivalent.
(i) f (λx) = λm f (x) for all λ > 0. (In other words, f is homogeneous of
degree m.) !
n
X ∂
(ii) xj − m f (x) = 0.
j=1
∂x j
31
Lemma 5.5. Suppose n ̸= 2.
(i) En is a well defined distribution according to our usual formula
Z
⟨En , f ⟩ = En (x)f (x) dx.
Rn
32
LemmaZ 5.8 (Proposition 5.1,
Z Green’s identities).
Z If u, v ∈ C 2 (B(0, r))
(a) v∆u dV = − ▽v.▽u dV + v▽u.nx dS.
Z ∥x∥<r ∥x∥<r Z ∥x∥=r
Much of the early work on the Laplacian used the following observation.
Lemma 5.9 (Proposition 5.2, energy estimate). If u ∈ C 2 (B(0, r))
Z Z Z
2
∥▽u∥ dV = u▽u.nx dS − u∆u dV.
∥x∥<r ∥x∥=r ∥x∥<r
Theorem 5.10. (i) The Dirichlet problem:- find u ∈ C 2 (B(0, r)) subject to
the conditions
∆u = f on B(0, r)
u=g on {x : ∥x∥ = r}
∆u = f on B(0, r)
▽u.nx = h on {x : ∥x∥ = r}
Theorem 5.12 (Uniqueness for the Dirichlet problem). The Dirichlet problem:-
find u ∈ C(Ω) with u is twice continuously differentiable in Ω subject to the
conditions
∆u = f on Ω
u=g on {x : ∥x∥ = r}
33
The point here is not so much the generality of Ω as that we only need u
differentiable on Ω.
If u is twice continuously differentiable on some open set Ω with ∆u =
we say that u is a harmonic function. The reader is already aware of the
connection with analytic functions. As an example we give another proof of
a result from 1B.
Lemma 5.14 (Gauss mean value theorem). (a) The value of a harmonic
function at a point is equal to its average over any sphere (boundary of a
ball) centred at that point.
(b) The value of a harmonic function at a point is equal to its average
over any ball centred at that point.
We shall see in Lemma 6.6 that condition (a) actually characterises har-
monic functions.
We shall probably not have time to develop the matter much further but
the associated definition (Definition 5.16) is important in later work.
for all ρ sufficiently small (depending on ξ) and all ξ ∈ Ω then we say that
u is subharmonic.
34
5.3 The dual Dirichlet problem for the unit ball
Let Ω be a bounded open set with smooth boundary ∂Ω.
(a) The Dirichlet problem asks for a solution of ∆f = 0 in Ω with f = h
on ∂Ω.
(b) The dual Dirichlet problem asks for a solution of ∆f = 0 in Ω with
f = g on ∂Ω.
We have been deliberately vague about the exact nature of f , h and g.
Maintaining the same vagueness we see that solving one problem is ‘more or
less’ equivalent to solving the other.
Let us try to solve the dual Dirichlet problem for the unit ball B.
∆f = g for ∥x∥ < 1
f =0 for ∥x∥ = 1
To make life easier for ourselves we shall initially assume g ∈ C(B) and
g infinitely differentiable on B. (In Lemma 6.3 we shall see that that the
formula obtained can be extended to a much wider class of functions.) We
shall work in Rn with n ≥ 3.
We start by defining
g̃ = g for ∥x∥ ≤ 1
g̃ = 0 for ∥x∥ > 1.
If E is the fundamental solution obtained earlier we start by looking at
f1 = E ∗ g̃. We know that, in a distributional sense,
∆E ∗ g̃ = g̃
but we wish to have a classical solution.
Lemma 5.18 (Lemma 5.1). (i) f1 ∈ C 1 (Rn ) and ∂j f1 = ∂j E ∗ g̃.
(ii) There is a constant such that |f1 (x)| ≤ A∥x∥2−n , |∂j f1 (x)| ≤ A∥x∥1−n
for ∥x∥ large.
Lemma 5.19 (Lemma 5.2). We have f1 infinitely differentiable on Rn \ {x :
∥x∥ = 1} and
∆f1 = g for ∥x∥ < 1
∆f1 = 0 for ∥x∥ > 1.
We thus have ∆f1 = g in B but the boundary conditions are wrong.
To get round this we use Kelvin’s method of reflections. (If you have done
electrostatics you will recognise this for dimension n = 2.) We put
2−n x
(Kf )(x) = ∥x∥ f .
∥x∥2
35
Lemma 5.20. If we set f2 = Kf1 and f = f1 − f2 then
(i) ∆f2 = 0 in B.
(ii) f1 = f2 on ∂B.
(iii) f solves our Dirichlet problem.
The proof of Lemma 5.20 (i) involves substantial calculation.
Lemma 5.21 (Lemma 5.3). (i) If f is twice differentiable
∂ ∂
r = (−n + 2)Kf − K r .
∂r ∂r
Lemma 5.21 shows us that f2 is harmonic in B \ {0} but a little more
work is needed to establish the full result.
Lemma 5.22 (Lemma 4). The function f2 is harmonic on B.
Doing the appropriate simple substitutions we obtain our final result.
Theorem 5.23 (Theorem 5.4). We work in Rn with n ≥ 3. If we set
∥x − y∥2−n ∥x∥2−n ∥∥x∥−2 x − y∥2
G(x, y) = −
ωn−1 (n − 2) ωn−1 (n − 2)
then Z
f (x) = G(x, y)g(y) dy
∥y∥<1
36
6 Dirichlet’s problem for the ball and Pois-
son’s formula
We noted earlier that a solution of the dual Dirichlet problem will give us a
solution for Dirichlet’s problem.
Lemma 6.1. Consider the Dirichlet problem for the unit ball B.
d
where P (x, y) = G(x, y) the directional derivative along the outward nor-
dn
mal.
Lemma 6.2. Consider Let the Dirichlet problem for the unit ball B.
where
1 1 − ∥x∥2
P (x, y) =
ωn−1 ∥x − y∥n
We call P the Poisson kernel. We can improve Lemma 6.2 and provide
a proof of the improvement which is essentially independent of the work
already done.
Lemma 6.3. Consider the Dirichlet problem for the unit ball B.
37
with h continuous.
If n ≥ 2 the solution is given by
Z
f (x) = P (x, y)h(y) dy
∥y∥<1
Lemma 6.4. (i) If y is fixed with ∥y∥ < 1 then P (x, y) is harmonic in x for
∥x∥ < 1.Z
(ii) P (x, y) dS(y) = 1
∥y∥=1
Lemma 6.5. The Dirichlet problem for any ball with continuous data has a
unique solution.
Lemma 6.5 is a very useful tool as the proof of the converse of Lemma 5.14 (a)
shows.
38
7 The wave equation
Recall that the wave operator 2 given by
∂ 2u
2u(t, x) = (t, x) − ∆x u(t, x)
∂t
where t ∈ R, x ∈ Rn and u : R × Rn → R is a suitable function.
Guided by our knowledge of physics we concentrate on solving two prob-
lems. The first is the forcing problem.
2u = f subject to u(t, x) = 0 for t ≤ 0
where f (t, x) = 0 for t ≤ 0. (Later we shall replace t ≤ 0 by the slightly
more general t large and negative.) The second is the Cauchy problem
∂u
2u = f subject to u(x, 0) = u0 (x) and (x, 0) = u1 (x).
∂t
The two problems are, as one might expect, closely related and we shall solve
the second problem by reducing it to the first.
Looking at Theorem 4.20 and Example 4.23 we see that we know two
fundamental solutions E+ and E− are given by
(n+1)/2 Z Z
1 ϕ̂(−(τ ± iα), −ξ)
⟨E± , ϕ⟩ = 2 2
dτ dξ,
2π Rn R ∥ξ∥ − (τ ± iα)
39
Putting everything together we obtain our solution.
Lemma 7.3. If n = 3 the the forward fundamental solution E+ is given by
Z ∞ Z
1
⟨E+ , f ⟩ = f (t, x) dS(x) dt.
0 4πt ∥x∥=t
The formula just given can be rewritten (we write E3,+ to emphasise that
n = 3) in various ways such as
Z ∞
f (y, ∥y∥)
Z Z
t 1
⟨E3,+ , ϕ⟩ = f (xt, t) dS(x) dt = dy.
0 4π ∥x∥=1 4π R3 ∥y∥
The main point at issue is whether this actually defines a distribution since
even when f ∈ D we will not (unless f = 0) have f˜ ∈ D. However the
argument of Lemma 4.3 applies since supp E3,+ ⊆ {(t, x) : ∥x∥ ≤ t} so ϵ2,+
is well defined. Now
⟨2ϵ2,+ , f ⟩ = ⟨ϵ2,+ , 2f ⟩
= ⟨E3,+ , 2f
f⟩
= ⟨E3,+ , 2f˜⟩
= ⟨2E3,+ , f˜⟩
= ⟨δ0 , f˜⟩
= ⟨δ0 , f ⟩
40
When we throw a pebble into a pond we do not get a single ripple but a
train of ripples!
It may be shown (see [1]) that in all dimensions the forward fundamental
solution has support in the ‘light cone’ {(t, x) : ∥x∥ ≤ t}, and that in odd
dimensions n > 1 the support lies on the surface {(t, x) : ∥x∥ ≤ t} (but
if n ≥ 5 the form of En,+ is more complicated than that of E3,+ , involving
analogues of δ0′ , δ0′′ and so on). Notice the contrast with ‘elliptic regularity’
as described in Theorem Weyl. If n is even the support of En,+ is spread out
over the whole light cone.
If n = 1 the matter follows an idea of D’Alembert familiar from 1B
mathematical methods.
Exercise 7.6. Let n = 1.
(i) Show that if we make the change of coordinates w = t + x, y = t − x
∂2
the wave operator becomes 2 .
∂w∂y
(ii) Show that the fundamental solution with support in {(w, y) : w ≥
1
0, y ≥ 0} is H(w)H(y).
2
A little thought shows that we have in fact a very general solution to our
forcing problem.
Lemma 7.7. Let A be a real number. If T is a distribution with supp T ⊆
{(t, x) ∈ R × Rn : t ≥ A} then there exists a unique distribution S with
supp S ⊆ {(t, x) ∈ R × Rn : t ≥ A} and
2S = T.
We have S = E+ ∗ T . If T is a k times differentiable function then so is S.
Reversing the sign of t we obtain E− the backwards fundamental solution.
We have the analogue of Lemma 7.7.
Lemma 7.8. Let A be a real number. If T is a distribution with supp T ⊆
{(t, x) ∈ R × Rn : t ≤ A} then there exists a unique distribution S with
suppS ⊆ {(t, x) ∈ R × Rn : t ≤ A} and
2S = T.
We have S = E− ∗ T . If T is a k times differentiable function then so is S.
If T ∈ D′ we can write T = T1 + T2 with
supp T1 ⊆ {(t, x) ∈ R×Rn : t ≥ A} and supp T2 ⊆ {(t, x) ∈ R×Rn : t ≤ B}
for some A and B.
41
Lemma 7.9. If T is a distribution then we can find a distribution S with
2S = T.
2u = f
with
∂u
u(0, x) = u0 (x), (0, x) = u1 (x).
∂t
Clearly it is enough to prove a simpler version.
Theorem 7.12. Suppose that u0 , u1 : Rn → R are smooth. Then there is a
unique solution of
2u = 0
with
∂u
u(0, x) = u0 (x), (0, x) = u1 (x).
∂t
It should be noted that if
2u = 0
with
∂u
u(0, x) = u0 (x), (0, x) = u1 (x),
∂t
then by the definition of 2
∂ 2u
= ∆u
∂t2
42
so in particular
∂ 2u
(0, x) = ∆u0 (x).
∂ 2t
Again
∂ 3u ∂u
3
= ∆u
∂t ∂t
so
∂ 3u
(0, x) = ∆u1 (x)
∂ 3t
∂j u
and so on. Thus j (0, x) is specified to all orders. We shall need a converse
∂ t
observation.
∂j u
(0, x) = uj (x)
∂j t
for all j ≥ 0 then
∂ku
2u (0, x) = 0
∂kt
for all x.
∂j u
(0, x) = uj (x)
∂j t
for all x and all j ≥ 0.
Borel’s lemma is plausible but not trivial. (The obvious Taylor series
argument fails because Taylor series need not converge except at the origin.)
The proof of Theorem 7.12 is now easy. By Theorem 7.14 and Lemma 7.13
we can find a smooth function v : R × Rn → R such that
k
∂ v
2u (0, x) = 0
∂kt
and
∂V
v(0, x) = u0 (x), (0, x) = u1 (x).
∂t
43
If we set f = 2u then by our choice of u all the partial derivatives of f with
respect to t vanish when t = 0 and the functions f1 , f2 : R × Rn → R defined
by
f1 (t, x) = H(t)f (t, x), f2 (t, x) = (1 − H(t))F (t, x)
are smooth. Setting
u = v − E+ ∗ f1 − E− ∗ f2
we have a solution of the required form. (Since E+ ∗ f1 is infinitely differen-
tiable with support in t ≥ 0 all its derivatives must vanish at t = 0.)
To prove uniqueness, observe that by linearity we need only prove that if
2u = 0 and u and its first partial derivative in t vanish at t = 0 then u = 0.
But if 2u = 0 and u and its first partial derivative in t vanish at t = 0 then
all the partial derivatives of u with respect to t vanish when t = 0. We can
thus write
u = u+ + u0
with u± smooth, supported in ±t ≥ 0 and satisfying 2u± = 0. By Lemma 7.7
u+ = 0. Similarly u− = 0 and we are done.
where u : R × Rn → R.
44
Lemma 8.2 (Proposition 7.1). If u ∈ D′ and J is the heat operator then u
and Ju have the same singular support. In particular if Ju is smooth so is
u.
This property is sometimes called hypo-ellipticity — saying that although
the operator is not elliptic it behaves like an elliptic operator it behaves
similarly in this respect.
We prove the next result by direct verification.
Theorem 8.3. If u0 : Rn → R and g : R × Rn → R are continuous functions
then the system
∂
− ∆ u(t, x) = g(t, x) for t > 0
∂t
subject to
lim u(t, x) = u0 (x)
t→0+
has a solution Z t
u(t, x) = Kt ∗ u0 + Kt−s ∗ gs ds
0
where Kt (x) = K(t, x) and gs (x) = g(s, x).
If g = 0 we observe that u(t, x) is a smooth function of x for fixed t > 0
whatever the choice of u0 this is another indication of the arrow of time.
8.2 Uniqueness
We did not claim uniqueness in Theorem 8.3 because the infinite propagation
speed of heat exhibited by our solution allows other (non-physical) solutions.
Exercise 8.4. (This is included for interest, it is non-examinable.) We work
with n = 1 for simplicity.
(i) Show formally (without worrying about rigour) that if g is smooth
∞
X g k (t)x2k
u(t, x) =
k=0
(2k)!
satisfies Ju = 0.
(ii) If g k (0) = 0 for all k but g is not identically zero then u is a solution
of Ju = 0 with u(t, x) → 0 as t → 0+ but u is not the zero function.
(iii) We now need to choose g and make the argument rigorous. It is
up to the reader to choose whether to try this. My own preference would be
to manufacture g by hand but g(t) = exp(−1/t2 ) [t ̸= 0], g(0) = 0 will do
(see [3], Section 7.1).
45
Because parabolic equations lie ‘on the boundary of elliptic equations’ we
try and use ideas about maxima of solutions similar to those of Theorems 5.11
and 5.12.
{(t, x) : ∥x − y∥ ≤ r, t = 0} ∪ {(t, x) : ∥x − y∥ = r, 0 ≤ t ≤ T }.
then
u(t, x) ≤ sup u(0, y)
y∈Rn
subject to
lim u(t, x) = u0 (x)
t→0+
46
Exercise 8.8. Suppose u1 , u2 : Rn → R and g : R × Rn → R are bounded
continuous functions. If ũj is the bounded solution of the system
∂
− ∆ ũj (t, x) = g(t, x) for t > 0
∂t
subject to
lim ũj (t, x) = uj (x)
t→0+
for j = 1, 2. Then
sup ∥ũ2 (t, x) − ũ1 (t, x)∥ = sup ∥u2 (x) − u1 (t, x)∥.
t≥0, x∈Rn x∈Rn
9 References
The reader will be in doubt of my admiration for Friedlander’s little book [2].
The additions to the second edition do not concern this course but some nig-
gling misprints have been removed and you should tell your college library to
buy the second edition even it already has the first. Friedlander’s book mainly
concerns the ‘distributional’ side of the course. Specific partial differential
equations are dealt with along with much else in Folland’s Introduction to
Partial Differential Equations [1]. The two books [2] and [1] are also excellent
further reading. The book [3] is a classic by a major worker in the field. 1+
(N12 − N13 ) (CC1 N12 − CCC1 N13 ) √ (N12 −N1
N12 N13 (N12 +N13 )+N12 N13 CC1 +N12 N13 CCC1 + √
N12 N13 (N12 +N13 )+
References
[1] G. B. Folland Introduction to Partial Differential Equations 2nd Edition,
Princeton University Press, 1995. QA324
47