Grad Anal
Grad Anal
Grad Anal
Richard B. Melrose
Contents
Introduction 6
Chapter 1. Measure and Integration 7
1. Continuous functions 7
2. Measures and σ-algebras 14
3. Measureability of functions 20
4. Integration 22
Chapter 2. Hilbert spaces and operators 35
1. Hilbert space 35
2. Spectral theorem 38
Chapter 3. Distributions 43
1. Test functions 43
2. Tempered distributions 50
3. Convolution and density 55
4. Fourier inversion 65
5. Sobolev embedding 70
6. Differential operators. 74
7. Cone support and wavefront set 89
8. Homogeneous distributions 102
9. Operators and kernels 103
10. Fourier transform 103
11. Schwartz space. 103
12. Tempered distributions. 104
13. Fourier transform 105
14. Sobolev spaces 106
15. Weighted Sobolev spaces. 109
16. Multiplicativity 112
17. Some bounded operators 115
Chapter 4. Elliptic Regularity 117
1. Constant coefficient operators 117
2. Constant coefficient elliptic operators 119
3. Interior elliptic estimates 126
3
4 CONTENTS
Bibliography 243
6 CONTENTS
Introduction
These notes are for the graduate analysis courses (18.155 and 18.156)
at MIT. They are based on various earlier similar courses. In giving
the lectures I usually cut many corners!
To thank:- Austin Frakt, Philip Dorrell, Jacob Bernstein....
CHAPTER 1
1. Continuous functions
A the beginning I want to remind you of things I think you already
know and then go on to show the direction the course will be taking.
Let me first try to set the context.
One basic notion I assume you are reasonably familiar with is that
of a metric space ([6] p.9). This consists of a set, X, and a distance
function
d : X × X = X 2 −→ [0, ∞) ,
satisfying the following three axioms:
i) d(x, y) = 0 ⇔ x = y, (and d(x, y) ≥ 0)
(1.1) ii) d(x, y) = d(y, x) ∀ x, y ∈ X
iii) d(x, y) ≤ d(x, z) + d(z, y) ∀ x, y, z ∈ X.
The basic theory of metric spaces deals with properties of subsets
(open, closed, compact, connected), sequences (convergent, Cauchy)
and maps (continuous) and the relationship between these notions.
Let me just remind you of one such result.
Proposition 1.1. A map f : X → Y between metric spaces is
continuous if and only if one of the three following equivalent conditions
holds
(1) f −1 (O) ⊂ X is open ∀ O ⊂ Y open.
(2) f −1 (C) ⊂ X is closed ∀ C ⊂ Y closed.
(3) limn→∞ f (xn ) = f (x) in Y if xn → x in X.
The basic example of a metric space is Euclidean space. Real n-
dimensional Euclidean space, Rn , is the set of ordered n-tuples of real
numbers
x = (x1 , . . . , xn ) ∈ Rn , xj ∈ R , j = 1, . . . , n .
7
8 1. MEASURE AND INTEGRATION
It is also the basic example of a vector (or linear) space with the oper-
ations
x + y = (x1 + y1 , x2 + y2 , . . . , xn + yn )
cx = (cx1 , . . . , cxn ) .
The metric is usually taken to be given by the Euclidean metric
Xn
2 2 1/2
|x| = (x1 + · · · + xn ) = ( x2j )1/2 ,
j=1
Now, the ‘obvious’ norm on this linear space is the supremum (or ‘uni-
form’) norm
kf k∞ = sup |f (x)| .
x∈X
Here X is an arbitrary metric space. For the moment X is sup-
posed to be a “physical” space, something like Rn . Corresponding to
the finite-dimensionality of Rn we often assume (or demand) that X
is locally compact. This just means that every point has a compact
neighborhood, i.e., is in the interior of a compact set. Whether locally
compact or not we can consider
(1.3) C0 (X) = f ∈ C(X); ∀ > 0 ∃ K b Xs.t. sup |f (x)| ≤ .
x∈K
/
This is also a normed linear space where the linear operations are
One of the basic questions I wish to pursue in the first part of the
course is: What is the dual of C0 (X) for a locally compact metric space
X? The answer is given by Riesz’ representation theorem, in terms of
(Borel) measures.
Let me give you a vague picture of ‘regularity of functions’ which
is what this course is about, even though I have not introduced most
of these spaces yet. Smooth functions (and small spaces) are towards
the top. Duality flips up and down and as we shall see L2 , the space
of Lebesgue square-integrable functions, is generally ‘in the middle’.
What I will discuss first is the right side of the diagramme, where we
have the space of continuous functions on Rn which vanish at infinity
and its dual space, Mfin (Rn ), the space of finite Borel measures. There
are many other spaces that you may encounter, here I only include test
functions, Schwartz functions, Sobolev spaces and their duals; k is a
1. CONTINUOUS FUNCTIONS 11
(1.5) S(R n ) w
_
*
H k (R n
) Cc (R n ) / C0 (Rn )
_ Kk _
y
L2 (R b
) s _
%
H −k
(R n
) M (Rn ) o ? _ Mfin (Rn )
_ Gg
t
0
S (Rn ).
I have set the goal of understanding the dual space Mfin (Rn ) of
C0 (X), where X is a locally compact metric space. This will force me
to go through the elements of measure theory and Lebesgue integration.
It does require a little forcing!
The basic case of interest is Rn . Then an obvious example of a
continuous linear functional on C0 (Rn ) is given by Riemann integration,
for instance over the unit cube [0, 1]n :
Z
u(f ) = f (x) dx .
[0,1]n
Note that (1.9) shows that u+ (f ) = u+ (f1 ) − u+ (f2 ) for any decom-
posiiton of f = f1 − f2 with fi ∈ C0 (X), both positive. [Since f1 + f− =
f2 + f+ so u+ (f1 ) + u+ (f− ) = u+ (f2 ) + u+ (f+ ).] Moreover,
|u+ (f )| ≤ max(u+ (f+ ), u(f− )) ≤ kuk kf k∞
=⇒ ku+ k ≤ kuk .
The functional
u− = u+ − u
is also positive, since u+ (f ) ≥ u(f ) for all 0 ≤ f ∈ C0 (x). Thus we
have proved
Lemma 1.5. Any element u ∈ (C0 (X))0 can be decomposed,
u = u+ − u−
into the difference of positive elements with
ku+ k , ku− k ≤ kuk .
The idea behind the definition of u+ is that u itself is, more or
less, “integration against a function” (even though we do not know
how to interpret this yet). In defining u+ from u we are effectively
throwing away the negative part of that ‘function.’ The next step is
to show that a positive functional corresponds to a ‘measure’ meaning
a function measuring the size of sets. To define this we really want to
evaluate u on the characteristic function of a set
1 if x ∈ E
χE (x) =
0 if x ∈/ E.
The problem is that χE is not continuous. Instead we use an idea
similar to (15.9).
If 0 ≤ u ∈ (C0 (X))0 and U ⊂ X is open, set1
(1.11) µ(U ) = sup {u(f ) ; 0 ≤ f (x) ≤ 1, f ∈ C0 (X) , supp(f ) b U } .
Here the support of f , supp(f ), is the closure of the set of points where
f (x) 6= 0. Thus supp(f ) is always closed, in (15.4) we only admit f if
its support is a compact subset of U. The reason for this is that, only
then do we ‘really know’ that f ∈ C0 (X).
Suppose we try to measure general sets in this way. We can do this
by defining
(1.12) µ∗ (E) = inf {µ(U ) ; U ⊃ E , U open} .
Already with µ it may happen that µ(U ) = ∞, so we think of
(1.13) µ∗ : P(X) → [0, ∞]
1See [6] starting p.42 or [1] starting p.206.
14 1. MEASURE AND INTEGRATION
µ∗ (A ∩ Fn ) = µ∗ (A ∩ Fn ∩ En ) + µ∗ (A ∩ Fn ∩ EnC )
= µ∗ (A ∩ En ) + µ∗ (A ∩ Fn−1 ) .
16 1. MEASURE AND INTEGRATION
µ∗ (A) = µ∗ (A ∩ Fn ) + µ∗ (A ∩ FnC )
Xn
≥ µ∗ (A ∩ Ej ) + µ∗ (A ∩ F C ) .
j=1
≥ µ∗ (A ∩ F ) + µ∗ (A ∩ F C ) ≥ µ∗ (A)
proves that inequalities are equalities, so F is also µ∗ -measurable.
In general, for any countable union of µ∗ -measurable sets,
[∞ ∞
[
Aj = A
ej ,
j=1 j=1
j−1 j−1
!C
[ [
ej = Aj \
A Ai = Aj ∩ Ai
i=1 i=1
Since
A\ supp(f ) ⊃ A ∩ U C , 0 ≤ f + g ≤ 1 , supp(f + g) b A ,
µ(A) ≥ u(f + g) = u(f ) + u(g)
> µ∗ (A ∩ U ) + µ∗ (A ∩ U C ) − 2
≥ µ∗ (A) − 2
using subadditivity of µ∗ . Letting ↓ 0 we conclude that
µ∗ (A) ≤ µ∗ (A ∩ U ) + µ∗ (A ∩ U C ) ≤ µ∗ (A) = µ(A) .
This gives (3.9) when A is open.
In general, if E ⊂ X and µ∗ (E) < ∞ then given > 0 there exists
A ⊂ X open with µ∗ (E) > µ∗ (A) − . Thus,
µ∗ (E) ≥ µ∗ (A ∩ U ) + µ∗ (A ∩ U C ) −
≥ µ∗ (E ∩ U ) + µ∗ (E ∩ U C ) −
≥ µ∗ (E) − .
This shows that (3.9) always holds, so U is µ∗ -measurable if it is open.
We have already observed that µ(U ) = µ∗ (U ) if U is open.
Thus we have shown that the σ-algebra given by Caratheodory’s
theorem contains all open sets. You showed in Problem 3 that the
intersection of any collection of σ-algebras on a given set is a σ-algebra.
Since P(X) is always a σ-algebra it follows that for any collection
E ⊂ P(X) there is always a smallest σ-algebra containing E, namely
\
ME = {M ⊃ E ; M is a σ-algebra , M ⊂ P(X)} .
The elements of the smallest σ-algebra containing the open sets are
called ‘Borel sets’. A measure defined on the σ-algebra of all Borel sets
is called a Borel measure. This we have shown:
Proposition 2.6. The measure defined by (15.4), (15.12) from
0 ≤ u ∈ (C0 (X))0 by Caratheodory’s theorem is a Borel measure.
Proof. This is what Proposition 3.14 says! See how easy proofs
are.
We can even continue in the same vein. A Borel measure is said to
be outer regular on E ⊂ X if
(2.6) µ(E) = inf {µ(U ) ; U ⊃ E , U open} .
Thus the measure constructed in Proposition 3.14 is outer regular on
all Borel sets! A Borel measure is inner regular on E if
(2.7) µ(E) = sup {µ(K) ; K ⊂ E , K compact} .
2. MEASURES AND σ-ALGEBRAS 19
Here we need to know that compact sets are Borel measurable. This
is Problem 5.
Definition 2.7. A Radon measure (on a metric space) is a Borel
measure which is outer regular on all Borel sets, inner regular on open
sets and finite on compact sets.
Proposition 2.8. The measure defined by (15.4), (15.12) from
0 ≤ u ∈ (C0 (X))0 using Caratheodory’s theorem is a Radon measure.
Proof. Suppose K ⊂ X is compact. Let χK be the charac-
teristic function of K , χK = 1 on K , χK = 0 on K C . Suppose
f ∈ C0 (X) , supp(f ) b X and f ≥ χK . Set
U = {x ∈ X ; f (x) > 1 − }
where > 0 is small. Thus U is open, by the continuity of f and
contains K. Moreover, we can choose g ∈ C(X) , supp(g) b U , 0 ≤
g ≤ 1 with g = 1 near3 K. Thus, g ≤ (1 − )−1 f and hence
µ∗ (K) ≤ u(g) = (1 − )−1 u(f ) .
Letting ↓ 0, and using the measurability of K,
µ(K) ≤ u(f )
⇒ µ(K) = inf {u(f ) ; f ∈ C(X) , supp(f ) b X , f ≥ χK } .
In particular this implies that µ(K) < ∞ if K b X, but is also proves
(3.17).
Let me now review a little of what we have done. We used the
positive functional u to define an outer measure µ∗ , hence a measure
µ and then checked the properties of the latter.
This is a pretty nice scheme; getting ahead of myself a little, let me
suggest that we try it on something else.
Let us say that Q ⊂ Rn is ‘rectangular’ if it is a product of finite
intervals (open, closed or half-open)
n
Y
(2.8) Q= (or[ai , bi ]or) ai ≤ bi
i=1
3Meaning in a neighborhood of K.
20 1. MEASURE AND INTEGRATION
3. Measureability of functions
Suppose that M is a σ-algebra on a set X 4 and N is a σ-algebra on
another set Y. A map f : X → Y is said to be measurable with respect
to these given σ-algebras on X and Y if
(3.1) f −1 (E) ∈ M ∀ E ∈ N .
Notice how similar this is to one of the characterizations of continuity
for maps between metric spaces in terms of open sets. Indeed this
analogy yields a useful result.
Lemma 3.1. If G ⊂ N generates N , in the sense that
\
(3.2) N = {N 0 ; N 0 ⊃ G, N 0 a σ-algebra}
then f : X −→ Y is measurable iff f −1 (A) ∈ M for all A ∈ G.
4Then X, or if you want to be pedantic (X, M), is often said to be a measure
space or even a measurable space.
3. MEASUREABILITY OF FUNCTIONS 21
is increasing and has limit f and that this limit is uniform on any
measurable set where f is finite.
4. Integration
The (µ)-integral of a non-negative simple function is by definition
Z X
(4.1) f dµ = ai µ(Y ∩ Ei ) , Y ∈ M .
Y i
Since these sets increase with n, if (4.4) does not hold then one of these
must have positive measure.
R In that case the simple function n−1 χEn
has positive integral so E f dµ > 0.
Notice the fundamental difference in approach here between Rie-
mann and Lebesgue integrals. The Lebesgue integral, (4.3), uses ap-
proximation by functions constant on possibly quite nasty measurable
sets, not just intervals as in the Riemann lower and upper integrals.
Theorem 4.3 (Monotone Convergence). Let fn be an increasing
sequence of non-negative measurable (extended) functions, then f (x) =
limn→∞ fn (x) is measurable and
Z Z
(4.5) f dµ = lim fn dµ
E n→∞ E
Since the sets (a, ∞] generate the Borel σ-algebra this shows that f is
measurable.
So we proceed to prove the main part of the proposition, which
is (4.5). Rudin has quite a nice proof of this, [6] page 21. Here I
paraphrase it. We can easily see from (4.1) that
Z Z Z
α = sup fn dµ = lim fn dµ ≤ f dµ.
E n→∞ E E
We further extend the integral to complex-valued functions, just
saying that
f :X→C
is integrable if its real and imaginary parts are both integrable. Then,
by definition,
Z Z Z
f dµ = Re f dµ + i Im f dµ
E E E
for any E ⊂ X measurable. It follows that if f is integrable then so is
|f |. Furthermore
Z Z
f dµ ≤ |f | dµ .
E E
R
This is obvious if E f dµ = 0, and if not then
Z
f dµ = Reiθ R > 0 , θ ⊂ [0, 2π) .
E
4. INTEGRATION 27
Then
Z Z
−iθ
f dµ = e f dµ
E E
Z
= e−iθ f dµ
ZE
= Re(e−iθ f ) dµ
ZE
≤ Re(e−iθ f ) dµ
ZE Z
−iθ
≤ e f dµ = |f | dµ .
E E
The other important convergence result for integrals is Lebesgue’s
Dominated convergence theorem.
Theorem 4.6. If fn is a sequence of integrable functions, fk → f
a.e.5 and |fn | ≤ g for some integrable g then f is integrable and
Z Z
f dµ = lim fn dµ .
n→∞
Similarly, if Hk = g + fk then
Z Z Z Z
(g + f )dµ = lim inf Hk dµ ≤ g dµ + lim inf fk dµ.
k→∞ k→∞
It follows that
Z Z Z
lim sup fk dµ ≤ f dµ ≤ lim inf fk dµ.
k→∞ k→∞
Thus in fact Z Z
fk dµ → f dµ .
Having proved Lebesgue’s theorem of dominated convergence, let
me use it to show something important. As before, let µ be a positive
measure on X. We have defined L1 (X, µ); let me consider the more
general space Lp (X, µ). A measurable function
f :X→C
is said to be ‘Lp ’, for 1 ≤ p < ∞, if |f |p is integrable6, i.e.,
Z
|f |p dµ < ∞ .
X
As before we consider equivalence classes of such functions under the
equivalence relation
(4.9) f ∼ g ⇔ µ {x; (f − g)(x) 6= 0} = 0 .
p
We denote by L (X, µ) the space of such equivalence classes. It is a
linear space and the function
Z 1/p
p
(4.10) kf kp = |f | dµ
X
1 1
for any 1 < p < ∞, with p
+ q
= 1.
Proof. If kf kp = 0 or kgkq = 0 the result is trivial, as it is if either
is infinite. Thus consider
p q
f (x) g(x)
a= ,b=
kf kp kgkq
and apply (5.16) with γ = p1 . This gives
|f (x)g(x)| |f (x)|p |g(x)|q
≤ + .
kf kp kgkq pkf kpp qkgkqq
Integrating over X we find
Z
1
|f (x)g(x)| dµ
kf kp kgkq X
1 1
≤ + = 1.
p q
R R
Since X
f g dµ ≤ X
|f g| dµ this implies (5.18).
The final inequality we need is Minkowski’s inequality.
Proposition 4.9. If 1 < p < ∞ and f, g ∈ Lp (X, µ) then
(4.14) kf + gkp ≤ kf kp + kgkp .
Proof. The case p = 1 you have already done. It is also obvious
if f + g = 0 a.e.. If not we can write
|f + g|p ≤ (|f | + |g|) |f + g|p−1
and apply Hölder’s inequality, to the right side, expanded out,
Z Z 1/q
p q(p−1)
|f + g| dµ ≤ (kf kp + kgkp ) , |f + g| dµ .
1
Since q(p − 1) = p and 1 − q
= 1/p this is just (5.20).
So, now we know that Lp (X, µ) is a normed space for 1 ≤ p < ∞. In
particular it is a metric space. One important additional property that
a metric space may have is completeness, meaning that every Cauchy
sequence is convergent.
30 1. MEASURE AND INTEGRATION
Next I want to return to our starting point and discuss the Riesz
representation theorem. There are two important results in measure
theory that I have not covered — I will get you to do most of them
in the problems — namely the Hahn decomposition theorem and the
Radon-Nikodym theorem. For the moment we can do without the
latter, but I will use the former.
So, consider a locally compact metric space, X. By a Borel measure
on X, or a signed Borel measure, we shall mean a function on Borel
sets
µ : B(X) → R
which is given as the difference of two finite positive Borel measures
(4.16) µ(E) = µ1 (E) − µ2 (E) .
Similarly we shall say that µ is Radon, or a signed Radon measure, if
it can be written as such a difference, with both µ1 and µ2 finite Radon
measures. See the problems below for a discussion of this point.
Let Mfin (X) denote the set of finite Radon measures on X. This is
a normed space with
(4.17) kµk1 = inf(µ1 (X) + µ2 (X))
with the infimum over all Radon decompositions (4.16). Each signed
Radon measure defines a continuous linear functional on C0 (X):
Z Z
(4.18) · dµ : C0 (X) 3 f 7−→ f · dµ .
X
1. Hilbert space
We have shown that Lp (X, µ) is a Banach space – a complete
normed space. I shall next discuss the class of Hilbert spaces, a spe-
cial class of Banach spaces, of which L2 (X, µ) is a standard example,
in which the norm arises from an inner product, just as it does in
Euclidean space.
An inner product on a vector space V over C (one can do the real
case too, not much changes) is a sesquilinear form
V ×V →C
written (u, v), if u, v ∈ V . The ‘sesqui-’ part is just linearity in the first
variable
(1.1) (a1 u1 + a2 u2 , v) = a1 (u1 , v) + a2 (u2 , v),
anti-linearly in the second
(1.2) (u, a1 v1 + a2 v2 ) = a1 (u, v1 ) + a2 (u, v2 )
and the conjugacy condition
(1.3) (u, v) = (v, u) .
Notice that (1.2) follows from (1.1) and (1.3). If we assume in addition
the positivity condition1
(1.4) (u, u) ≥ 0 , (u, u) = 0 ⇒ u = 0 ,
then
(1.5) kuk = (u, u)1/2
is a norm on V , as we shall see.
Suppose that u, v ∈ V have kuk = kvk = 1. Then (u, v) =
e |(u, v)| for some θ ∈ R. By choice of θ, e−iθ (u, v) = |(u, v)| is
iθ
2. Spectral theorem
For a bounded operator T on a Hilbert space we define the spectrum
as the set
(2.1) spec(T ) = {z ∈ C; T − z Id is not invertible}.
Proposition 2.1. For any bounded linear operator on a Hilbert
space spec(T ) ⊂ C is a compact subset of {|z| ≤ kT k}.
Proof. We show that the set C \ spec(T ) (generally called the
resolvent set of T ) is open and contains the complement of a sufficiently
large ball. This is based on the convergence of the Neumann series.
Namely if T is bounded and kT k < 1 then
X∞
(2.2) (Id −T )−1 = Tj
j=0
The total mass |µφ,ψ | of this measure is the norm of the functional.
Since it is a Borel measure, we can take the integral on −∞, b] for any
b ∈ R ad, with the uniqueness, this shows that we have a continuous
sesquilinear map
(2.21) Z
Pb (φ, ψ) : H×H 3 (φ, ψ) 7−→ dµφ,ψ ∈ R, |Pb (φ, ψ)| ≤ kAkkφkkψk.
[m,b]
which takes values in the projections on H and which allows the func-
tions of A to be written as integrals in the form
Z
(2.25) f (A) = f dµ
[m,M ]
Distributions
1. Test functions
So far we have largely been dealing with integration. One thing we
have seen is that, by considering dual spaces, we can think of functions
as functionals. Let me briefly review this idea.
Consider the unit ball in Rn ,
n
B = {x ∈ Rn ; |x| ≤ 1} .
I take the closed unit ball because I want to deal with a compact metric
space. We have dealt with several Banach spaces of functions on Bn ,
for example
C(Bn ) = u : Bn → C ; u continuous
Z
2 2
L (Bn ) = u : Bn → C; Borel measurable with |u| dx < ∞ .
Definition 1.2. Let C01 (Rn ) be the subspace of C0 (Rn ) = C00 (Rn )
such that each element u ∈ C01 (Rn ) is continuously differentiable and
∂u
∂xj
∈ C0 (Rn ), j = 1, . . . , n.
Proposition 1.3. The function
n
X ∂u
kukC 1 = kuk∞ + k k∞
i=1
∂x1
is a norm on C01 (Rn ) with respect to which it is a Banach space.
Proof. That k kC 1 is a norm follows from the properties of k k∞ .
Namely kukC 1 = 0 certainly implies u = 0, kaukC 1 = |a| kukC 1 and the
triangle inequality follows from the same inequality for k k∞ .
Similarly, the main part of the completeness of C01 (Rn ) follows from
the completeness of C00 (Rn ). If {un } is a Cauchy sequence in C01 (Rn )
then un and the ∂u n
∂xj
are Cauchy in C00 (Rn ). It follows that there are
limits of these sequences,
∂un
un → v , → vj ∈ C00 (Rn ) .
∂xj
However we do have to check that v is continuously differentiable and
∂v
that ∂x j
= vj .
One way to do this is to use the Fundamental Theorem of Calculus
in each variable. Thus
Z t
∂un
un (x + tei ) = (x + sei ) ds + un (x) .
0 ∂xj
As n → ∞ all terms converge and so, by the continuity of the integral,
Z t
u(x + tei ) = vj (x + sei ) ds + u(x) .
0
This shows that the limit in (6.20) exists, so vi (x) is the partial deriva-
tion of u with respect to xi . It remains only to show that u is indeed
differentiable at each point and I leave this to you in Problem 17.
So, almost by definition, we have an example of Lemma 6.17,
C01 (Rn ) ,→ C00 (Rn ).
It is in fact dense but I will not bother showing this (yet). So we know
that
(C00 (Rn ))0 → (C01 (Rn ))0
and we expect it to be injective. Thus there are more functionals on
C01 (Rn ) including things that are ‘more singular than measures’.
46 3. DISTRIBUTIONS
These are all Banach spaces, since if {un } is Cauchy in C0k (Rn ),
it is Cauchy and hence convergent in C0k−1 (Rn ), as is ∂un /∂xj , j =
1, . . . , n − 1. Furthermore the limits of the ∂un /∂xj are the derivatives
of the limits by Proposition 1.3.
This gives us a sequence of spaces getting ‘smoother and smoother’
C00 (Rn ) ⊃ C01 (Rn ) ⊃ · · · ⊃ C0k (Rn ) ⊃ · · · ,
with norms getting larger and larger. The duals can also be expected
to get larger and larger as k increases.
As well as looking at functions getting smoother and smoother, we
need to think about ‘infinity’, since Rn is not compact. Observe that
an element g ∈ L1 (Rn ) (with respect to Lebesgue measure by default)
defines a functional on C00 (Rn ) — and hence all the C0k (Rn )s. However a
function such as the constant function 1 is not integrable on Rn . Since
we certainly want to talk about this, and polynomials, we consider a
second condition of smallness at infinity. Let us set
(1.9) hxi = (1 + |x|2 )1/2
a function which is the size of |x| for |x| large, but has the virtue of
being smooth1
1See Problem 18.
1. TEST FUNCTIONS 47
(1.14) ˜ v) = ku − vk
d(u,
1 + ku − vk
is a metric on any normed space, since then we may sum over k. Thus
we consider
ku − vk kv − wk
+
1 + ku − vk 1 + kv − wk
ku − vk(1 + kv − wk) + kv − wk(1 + ku − vk)
= .
(1 + ku − vk)(1 + kv − wk)
˜ w) we must show that
Comparing this to d(v,
(1 + ku − vk)(1 + kv − wk)ku − wk
≤ (ku − vk(1 + kv − wk) + kv − wk(1 + ku − vk))(1 + ku − wk).
Starting from the LHS and using the triangle inequality,
LHS ≤ ku − wk + (ku − vk + kv − wk + ku − vkkv − wk)ku − wk
≤ (ku − vk + kv − wk + ku − vkkv − wk)(1 + ku − wk)
≤ RHS.
Thus, d is a metric.
Suppose un is a Cauchy sequence. Thus, d(un , um ) → 0 as n, m →
∞. In particular, given
> 0 ∃ N s.t. n, m > N implies
d(un , um ) < 2−k ∀ n, m > N.
The terms in (1.13) are all positive, so this implies
kun − um k(k)
< ∀ n, m > N.
1 + kun − um k(k)
If < 1/2 this in turn implies that
kun − um k(k) < 2,
1. TEST FUNCTIONS 49
so the sequence is Cauchy in hxi−k C0k (Rn ) for each k. From the com-
pleteness of these spaces it follows that un → u in hxi−k C0k (Rn )j for
each k. Given > 0 choose k so large that 2−k < /2. Then ∃ N s.t.
n>N
⇒ ku − un k(j) < /2 n > N, j ≤ k.
Hence
X ku − un k(j)
d(un , u) = 2−j
j≤k
1 + ku − un k(j)
X ku − un k(j)
+ 2−j
j>k
1 + ku − un k(j)
Consider
(1.17)
T = {U ⊂ Cc∞ (Rn ); U ∩ C˙c∞ (B(n)) is open in C˙∞ (B(n)) for each n}.
This is a topology on Cc∞ (Rn ) – contains the empty set and the whole
space and is closed under finite intersections and arbitrary unions –
50 3. DISTRIBUTIONS
simply because the same is true for the open sets in C˙∞ (B(n)) for each
n. This is in fact the inductive limit topology. One obvious question
is:- what does it mean for a linear functional u : Cc∞ (Rn ) −→ C to be
continuous? This just means that u−1 (O) is open for each open set in C.
Directly from the definition this in turn means that u−1 (O)∩ C˙∞ (B(n))
should be open in C˙∞ (B(n)) for each n. This however just means that,
restricted to each of these subspaces u is continuous. If you now go
forwards to Lemma 2.3 you can see what this means; see Problem 74.
Of course there is a lot more to be said about these spaces; you can
find plenty of it in the references.
2. Tempered distributions
A good first reference for distributions is [2], [5] gives a more ex-
haustive treatment.
The complete metric topology on S(Rn ) is described above. Next I
want to try to convice you that elements of its dual space S 0 (Rn ), have
enough of the properties of functions that we can work with them as
‘generalized functions’.
First let me develop some notation. A differentiable function ϕ :
n n
R → C has partial derivatives which we have denoted ∂ϕ/∂x √ j:R →
C. For reasons that will become clear later, we put a −1 into the
definition and write
1 ∂ϕ
(2.1) Dj ϕ = .
i ∂xj
We say ϕ is once continuously differentiable if each of these Dj ϕ is
continuous. Then we defined k times continuous differentiability in-
ductively by saying that ϕ and the Dj ϕ are (k − 1)-times continuously
differentiable. For k = 2 this means that
Dj Dk ϕ are continuous for j, k = 1, · · · , n .
Now, recall that, if continuous, these second derivatives are symmetric:
(2.2) Dj Dk ϕ = Dk Dj ϕ .
This means we can use a compact notation for higher derivatives. Put
N0 = {0, 1, . . .}; we call an element α ∈ Nn0 a ‘multi-index’ and if ϕ is
at least k times continuously differentiable, we set3
1 ∂ α1 ∂ αn
(2.3) Dα ϕ = |α| ··· ϕ whenever |α| = α1 +α2 +· · ·+αn ≤ k.
i ∂x1 ∂xn
3Periodicallythere is the possibility of confusion between the two meanings of
|α| but it seldom arises.
2. TEMPERED DISTRIBUTIONS 51
Thus,
(2.7) ϕ ∈ S(Rn ) ⇔ Dα (hxik ϕ) ∈ C00 (Rn ) ∀ |α| ≤ k and all k .
Lemma 2.1. The condition ϕ ∈ S(Rn ) can be written
hxik Dα ϕ ∈ C00 (Rn ) ∀ |α| ≤ k , ∀ k .
Proof. We first check that
ϕ ∈ C00 (Rn ) , Dj (hxiϕ) ∈ C00 (Rn ) , j = 1, · · · , n
⇔ ϕ ∈ C00 (Rn ) , hxiDj ϕ ∈ C00 (Rn ) , j = 1, · · · , n .
Since
Dj hxiϕ = hxiDj ϕ + (Dj hxi)ϕ
and Dj hxi = 1i xj hxi−1 is a bounded continuous function, this is clear.
Then consider the same thing for a larger k:
(2.8) Dα hxip ϕ ∈ C00 (Rn ) ∀ |α| = p , 0 ≤ p ≤ k
⇔ hxip Dα ϕ ∈ C00 (Rn ) ∀ |α| = p , 0 ≤ p ≤ k .
I leave you to check this as Problem 2.1.
Corollary 2.2. For any k ∈ N the norms
X
khxik ϕkC k and kxα Dxβ ϕk∞
|α|≤k,
|β|≤k
are equivalent.
52 3. DISTRIBUTIONS
X
≤ C0 sup xα Dβ ψ .
Rn
|α|≤k+1, |β|≤k
but we need to know that these maps are injective before we can forget
about them.
We can see this using convolution. This is a sort of ‘product’ of
functions. To begin with, suppose v ∈ C00 (Rn ) and ψ ∈ S(Rn ). We
define a new function by ‘averaging v with respect to ψ:’
Z
(3.3) v ∗ ψ(x) = v(x − y)ψ(y) dy .
Rn
To see that this is small for x0 small, we split the integral into two
pieces. Since ψ is very small near infinity, given > 0 we can choose
R so large that
Z
(3.5) kvk∞ · |ψ(y)| dy ≤ /4 .
|y]|≥R
Here we have used the fact that ϕt ≥ 0 has support in |y| ≤ t and has
integral 1. Thus vt → v uniformly on any set on which v is uniformly
continuous, namel Rn !
Corollary 3.3. C0k (Rn ) is dense in C0p (Rn ) for any k ≥ p.
Proposition 3.4. S(Rn ) is dense in C0k (Rn ) for any k ≥ 0.
Proof. Take k = 0 first. The subspace Cc0 (Rn ) is dense in C00 (Rn ),
by cutting off outside a large ball. If v ∈ Cc0 (Rn ) has support in
{|x| ≤ R} then
v ∗ ϕt ∈ Cc∞ (Rn ) ⊂ S(Rn )
has support in {|x| ≤ R + 1}. Since v ∗ ϕt → v the result follows for
k = 0.
For k ≥ 1 the same argument works, since Dα (v ∗ ϕt ) = (Dα V ) ∗
ϕt .
Corollary 3.5. The map from finite Radon measures
(3.19) Mfin (Rn ) 3 µ 7−→ uµ ∈ S 0 (Rn )
is injective.
Now, we want the same result for L2 (Rn ) (and maybe for Lp (Rn ),
1 ≤ p < ∞). I leave the measure-theoretic part of the argument to
you.
Proposition 3.6. Elements of L2 (Rn ) are “continuous in the mean”
i.e.,
Z
(3.20) lim |u(x + t) − u(x)|2 dx = 0 .
|t|→0 Rn
Note that at the second step here I have used Schwarz’s inequality with
the integrand written as the product
1/2 1/2 1/2 1/2
|u(x − y) − u(x)| ϕt (y)ϕt (y 0 ) · |u(x − y 0 ) − u(x)| ϕt (y)ϕt (y 0 ) .
Thus we now know that
L2 (Rn ) ,→ S 0 (Rn ) is injective.
This means that all our usual spaces of functions ‘sit inside’ S 0 (Rn ).
60 3. DISTRIBUTIONS
Clearly the right side of (3.25) is contained in the left. To see the
converse, suppose first that
(3.26) ϕ ∈ S(Rn ) , ϕ = 0 in |x| < 1 .
Then define
0 |x| < 1
ψ= 2
ϕ/ |x| |x| ≥ 1 .
this shows that ϕ of the form (3.26) is in the right side of (3.25). In
general suppose ϕ ∈ S(Rn ). Then
Z t
d
ϕ(x) − ϕ(0) = ϕ(tx) dt
0 dt
(3.27) n Z t
X ∂ϕ
= xj (tx) dt .
j=1 0 ∂xj
Certainly these integrals are C ∞ , but they may not decay rapidly at
infinity. However, choose µ ∈ Cc∞ (Rn ) with µ = 1 in |x| ≤ 1. Then
(3.27) becomes, if ϕ(0) = 0,
ϕ = µϕ + (1 − µ)ϕ
n Z t
X ∂ϕ
= xj ψj + (1 − µ)ϕ , ψj = µ (tx) dt ∈ S(Rn ) .
j=1 0 ∂xj
→ 0 as ξ 0 → ξ .
In fact
Proposition 3.11. Fourier transformation, (3.28), defines a con-
tinuous linear map
(3.29) F : S(Rn ) → S(Rn ) , Fϕ = ϕ̂ .
Proof. Differentiating under the integral6 sign shows that
Z
∂ξj ϕ̂(ξ) = −i e−ix·ξ xj ϕ(x) dx .
Since the integral on the right is absolutely convergent that shows that
(remember the i’s)
(3.30) xj ϕ , ∀ ϕ ∈ S(Rn ) .
Dξj ϕ̂ = −d
Similarly, if we multiply by ξj and observe that ξj e−ix·ξ = i ∂x∂ j e−ix·ξ
then integration by parts shows
Z
∂ −ix·ξ
(3.31) ξj ϕ̂ = i ( e )ϕ(x) dx
∂xj
Z
∂ϕ
= −i e−ix·ξ dx
∂xj
n
D j ϕ = ξj ϕ̂ , ∀ ϕ ∈ S(R ) .
d
Since xj ϕ, Dj ϕ ∈ S(Rn ) these results can be iterated, showing that
ξ α Dξβ ϕ̂ = F (−1)|β| Dα x xβ ϕ .
(3.32)
Since this is true for ϕ = exp(− |x|2 /2). The identity allows us to
invert the Fourier transform.
4. Fourier inversion
It is shown above that the Fourier transform satisfies the identity
Z
−n
(4.1) ϕ(0) = (2π) ϕ̂(ξ) dξ ∀ ϕ ∈ S(Rn ) .
Rn
= eiy·ξ ϕ̂(ξ) .
Applied to ψ the inversion formula (4.1) becomes
Z
−n
(4.2) ϕ(y) = ψ(0) = (2π) ψ̂(ξ) dξ
Z
−n
= (2π) eiy·ξ ϕ̂(ξ) dξ .
Rn
D
d α u(ϕ) = D α u(ϕ̂) = u((−1)|α| D α ϕ̂)
Thus hξim û = α m α m
P
|α|≤−m ξ hξi v̂α . Since all the factors ξ hξi are
2 n m 2 n
bounded, each term here is in L (R ), so hξi û ∈ L (R ) which is the
definition, u ∈ hξi−m L2 (Rn ).
Conversely, suppose u ∈ H m (Rn ), i.e., hξim û ∈ L2 (Rn ). The func-
tion
X
|ξ α | · hξim ∈ L2 (Rn ) (m < 0)
|α|≤−m
Each of the functions v̂α = sgn(ξ α )v̂ ∈ L2 (Rn ) so the identity (4.16),
and hence (4.15), follows with these choices.
4. FOURIER INVERSION 69
pairing
(4.18) H m (Rn ) × H −m (Rn ) 3 (u, u0 ) 7−→
Z
0
((u, u )) = û0 (ξ)û0 (·ξ) dξ ∈ C
Rn
n 0
gives an identification (H m (R )) = H −m (Rn ).
Proof. The Hilbert space property follows essentially directly from
the definition (4.14) since hξi−m L2 (Rn ) is a Hilbert space with the norm
(4.17). Similarly the density of S in H m (Rn ) follows, since S(Rn ) dense
in L2 (Rn ) (Problem L11.P3) implies hξi−m S(Rn ) = S(Rn ) is dense in
hξi−m L2 (Rn ) and so, since F is an isomorphism in S(Rn ), S(Rn ) is
dense in H m (Rn ).
Finally observe that the pairing in (4.18) makes sense, since hξi−m û(ξ),
hξi û (ξ) ∈ L2 (Rn ) implies
m 0
5. Sobolev embedding
The properties of Sobolev spaces are briefly discussed above. If
m is a positive integer then u ∈ H m (Rn ) ‘means’ that u has up to
m derivatives in L2 (Rn ). The question naturally arises as to the sense
in which these ‘weak’ derivatives correspond to old-fashioned ‘strong’
derivatives. Of course when m is not an integer it is a little harder
to imagine what these ‘fractional derivatives’ are. However the main
result is:
Theorem 5.1 (Sobolev embedding). If u ∈ H m (Rn ) where m >
n/2 then u ∈ C00 (Rn ), i.e.,
(5.1) H m (Rn ) ⊂ C00 (Rn ) , m > n/2 .
Proof. By definition, u ∈ H m (Rn ) means v ∈ S 0 (Rn ) and hξim û(ξ) ∈
L2 (Rn ). Suppose first that u ∈ S(Rn ). The Fourier inversion formula
shows that
Z
n
(2π) |u(x)| = eix·ξ û(ξ) dξ
Z 1/2 !1/2
X
≤ hξi2m |û(ξ)|2 dξ · hξi−2m dξ .
Rn Rn
Now, if m > n/2 then the second integral is finite. Since the first
integral is the norm on H m (Rn ) we see that
(5.2) sup |u(x)| = kukL∞ ≤ (2π)−n kukH m , m > n/2 .
Rn
5. SOBOLEV EMBEDDING 71
Proof. This follows directly from (5.5) since the left side is con-
tained in \
hxi−k C0k−n (Rn ) ⊂ S(Rn ).
k
Theorem 5.5 (Schwartz representation). Any tempered distribu-
tion can be written in the form of a finite sum
X
(5.8) u= xα Dxβ uαβ , uαβ ∈ C00 (Rn ).
|α|≤m
|β|≤m
or in the form
X
(5.9) u= Dxβ (xα vαβ ), vαβ ∈ C00 (Rn ).
|α|≤m
|β|≤m
Thus,8
X
(5.10) u = hxik Dαγ vγ , vγ ∈ C00 (Rn ) .
|γ|≤M
To get (5.9) we ‘commute’ the factor hxik to the inside; since I have
not done such an argument carefully so far, let me do it as a lemma.
Lemma 5.6. For any γ ∈ Nn0 there are polynomials pα,γ (x) of de-
grees at most |γ − α| such that
X
hxik Dγ v = Dγ−α pα,γ hxik−2|γ−α| v .
α≤γ
6. Differential operators.
In the last third of the course we will apply what we have learned
about distributions, and a little more, to understand properties of dif-
ferential operators with constant coefficients. Before I start talking
about these, I want to prove another density result.
So far we have not defined a topology on S 0 (Rn ) – I will leave
this as an optional exercise.9 However we shall consider a notion of
convergence. Suppose uj ∈ S 0 (Rn ) is a sequence in S 0 (Rn ). It is said
to converge weakly to u ∈ S 0 (Rn ) if
(6.1) uj (ϕ) → u(ϕ) ∀ ϕ ∈ S(Rn ) .
There is no ‘uniformity’ assumed here, it is rather like pointwise con-
vergence (except the linearity of the functions makes it seem stronger).
Proposition 6.1. The subspace S(Rn ) ⊂ S 0 (Rn ) is weakly dense,
i.e., each u ∈ S 0 (Rn ) is the weak limit of a subspace uj ∈ S(Rn ).
Proof. We can use Schwartz representation theorem to write, for
some m depending on u,
X
u = hxim Dα uα , uα ∈ L2 (Rn ) .
|α|≤m
Here I have cut the space {|x| ≤ , |y| ≤ } out of the integral and used
the local integrability in taking the limit as ↓ 0. Integrating by parts
in x we find
Z Z
−1
− (x + iy) ∂x ϕ dx dy = (∂x (x + iy)−1 )ϕ dx dy
|x|≥ |x|≥
|y|≥ |y|≥
Z Z
−1
+ (x + iy) ϕ(x, y) dy − (x + iy)−1 ϕ(x, y) dy .
|y|≤ |y|≤
x= x=−
(6.13) 2π∂E(ϕ) =
Z
lim [( + iy)−1 ϕ(, y) − (− + iy)−1 ϕ(−, y)] dy
↓0 |y|≤
Z
+ i lim [(x + i)−1 ϕ(x, ) − (x − i)−1 ϕ(x, )] dx ,
↓0 |x|≤
Thus we get the same result in (6.13) by replacing ϕ(x, y) by ϕ(0, 0).
Then 2π∂E(ϕ) = cϕ(0),
Z Z
dy dy
c = lim 2 2 2
= lim < 2
= 2π .
↓0 |y|≤ + y ↓0 |y|≤1 1 + y
Let me remind you that we have already discussed the convolution
of functions
Z
u ∗ v(x) = u(x − y)v(y) dy = v ∗ u(x) .
so it follows that
(6.15) |u ∗ ϕ(x)| = |hu, ϕ(x − ·)i| ≤ C(1 + |x|2 )k/2 .
The argument above shows that x 7→ ϕ(x − ·) is a continuous func-
tion of x ∈ Rn with values in S(Rn ), so u ∗ ϕ is continuous and satisfies
(6.15). It is therefore an element of S 0 (Rn ).
Differentiability follows in the same way since for each j, with ej
the jth unit vector
ϕ(x + sej − y) − ϕ(x − y)
∈ S(Rn )
s
is continuous in x ∈ Rn , s ∈ R. Thus, u ∗ ϕ has continuous partial
derivatives and
Dj u ∗ ϕ = u ∗ Dj ϕ .
The same argument then shows that u∗ϕ ∈ C ∞ (Rn ). That Dj (u∗ϕ) =
Dj u ∗ ϕ follows from the definition of derivative of distributions
Dj (u ∗ ϕ(x)) = (u ∗ Dj ϕ)(x)
= hu, Dxj ϕ(x − y)i = −hu(y), Dyj ϕ(x − y)iy
= (Dj u) ∗ ϕ .
Finally consider the support property. Here we are assuming that
supp(ϕ) is compact; we also know that supp(u) is a closed set. We
have to show that
(6.16) x∈
/ supp(u) + supp(ϕ)
implies u ∗ ϕ(x ) = 0 for x0 near x. Now (6.16) just means that
0
≤ CN (1 + |x|2 )−N/2 .
Thus (1 + |x|2 )N/2 |v ∗ ϕ| is bounded for each N . The same argument
applies to the derivative using Theorem 6.6, so
v ∗ ϕ ∈ S(Rn ) .
80 3. DISTRIBUTIONS
In fact we get a little more, since we see that for each k there exists
k 0 and C (depending on k and v) such that
kv ∗ ϕk(k) ≤ Ckϕk(k0 ) .
This means that
v∗ : S(Rn ) → S(Rn )
is a continuous linear map.
Now (6.19) allows us to define u∗v when u ∈ S 0 (Rn ) and v ∈ S 0 (Rn )
has compact support by
u ∗ v(ϕ) = u ∗ (v ∗ ϕ̌)(0) .
Using the continuity above, I ask you to check that u ∗ v ∈ S 0 (Rn ) in
Problem 36. For the moment let me assume that this convolution has
the same properties as before – I ask you to check the main parts of
this in Problem 37.
Recall that E ∈ S 0 (Rn ) is a fundamental situation for P (D), a
constant coefficient differential operator, if P (D)E = δ. We also use a
weaker notion.
Definition 6.8. A parametrix for a constant coefficient differential
operator P (D) is a distribution F ∈ S 0 (Rn ) such that
(6.20) P (D)F = δ + ψ , ψ ∈ C ∞ (Rn ) .
An operator P (D) is said to be hypoelliptic if it has a parametrix sat-
isfying
(6.21) sing supp(F ) ⊂ {0} ,
where for any u ∈ S 0 (Rn )
The second condition here means that if ϕ ∈ Cc∞ (Rn ) and ϕ(x) = 1
in |x| < for some > 0 then (1 − ϕ)F ∈ C ∞ (Rn ). Since P (D)((1 −
ϕ)F ) ∈ C ∞ (Rn ) we conclude that
P (D)(ϕF ) − δ ∈ Cc∞ (Rn )
and we may well suppose that F , replaced now by ϕF , has compact
support. Last time I showed that
If P (D) is hypoelliptic and u ∈ S 0 (Rn ) then
sing supp(u) = sing supp(P (D)u) .
I will remind you of the proof later.
First however I want to discuss the important notion of ellipticity.
Remember that P (D) is ‘really’ just a polynomial, called the charac-
teristic polynomial
X
P (ξ) = Cα ξ α .
|α|≤m
However even the Laplacian, ∆ = nj=1 Dj2 , does not satisfy this rather
P
stringent condition.
It is reasonable to expect the top order derivatives to be the most
important. We therefore consider
X
Pm (ξ) = Cα ξ α
|α|=m
Now xd α F = (−1)|α| D α F
ξ , so what we need to cinsider is the behaviour
b
of the derivatives of Fb, which is just Q(ξ) in (6.32).
Lemma 6.14. Let P (ξ) be a polynomial of degree m satisfying
(6.35) |P (ξ)| ≥ C |ξ|m in |ξ| > 1/C for some C > 0 ,
then for some constants Cα
1
(6.36) Dα ≤ Cα |ξ|−m−|α| in |ξ| > 1/C .
P (ξ)
Proof. The estimate in (6.36) for α = 0 is just (6.35). To prove
the higher estimates that for each α there is a polynomial of degree at
most (m − 1) |α| such that
1 Lα (ξ)
(6.37) Dα = .
P (ξ) (P (ξ))1+|α|
Once we know (6.37) we get (6.36) straight away since
1 Cα0 |ξ|(m−1)|α|
D α
≤ m(1+|α|)
≤ Cα |ξ|−m−|α| .
P (ξ) C 1+|α| |ξ|
We can prove (6.37) by induction, since it is certainly true for α = 0.
Suppose it is true for |α| ≤ k. To get the same identity for each β with
|β| = k +1 it is enough to differentiate one of the identities with |α| = k
once. Thus
1 1 Dj Lα (ξ) (1 + |α|)Lα Dj P (ξ)
Dβ = Dj Dα = 1+|α|
− .
P (ξ) P (ξ) P (ξ) (P (ξ))2+|α|
Since Lβ (ξ) = P (ξ)Dj Lα (ξ) − (1 + |α|)Lα (ξ)Dj P (ξ) is a polynomial of
degree at most (m − 1) |α| + m − 1 = (m − 1) |β| this proves the lemma.
6. DIFFERENTIAL OPERATORS. 85
1−ϕ
Going backwards, observe that Q(ξ) = Pm (ξ)
is smooth in |ξ| ≤ 1/C,
so (6.36) implies that
(6.38) |Dα Q(ξ)| ≤ Cα (1 + |ξ|)−m−|α|
n
⇒ hξi` Dα Q ∈ L2 (Rn ) if ` − m − |α| < − ,
2
which certainly holds if ` = |α| + m − n − 1, giving (6.34). Now, by
Sobolev’s embedding theorem
n
xα F ∈ C k if |α| > n + 1 − m + k + .
2
∞ n
In particular this means that if we choose µ ∈ Cc (R ) with 0 ∈
/ supp(µ)
2k
then for every k, µ/ |x| is smooth and
µ
µF = 2k |x|2k F ∈ C 2`−2n , ` > n .
|x|
Thus µF ∈ Cc∞ (Rn ) and this is what we wanted to show, sing supp(F ) ⊂
{0}.
So now we have actually proved that Pm (D) is hypoelliptic if it is
elliptic. Rather than go through the proof again to make sure, let me
go on to the general case and in doing so review it.
Proof. Proof of theorem. We need to show that if P (ξ) is elliptic
then P (D) has a parametrix F as in (6.27). From the discussion above
the ellipticity of P (ξ) implies (and is equivalent to)
|Pm (ξ)| ≥ c |ξ|m , c > 0 .
On the other hand
X
P (ξ) − Pm (ξ) = Cα ξ α
|α|<m
(0, ∞)} for 0 6= x̄ ∈ Rn . Since two points give the same half-line if and
only if they are positive multiples of each other, this means we think
of the sphere as the quotient
(7.1) Sn−1 = (Rn \ {0})/R+ .
Of course if we have a metric on Rn , for instance the usual Euclidean
metric, then we can identify Sn−1 with the unit sphere. However (7.1)
does not require a choice of metric.
Now, suppose we consider functions on Rn \ {0} which are (posi-
tively) homogeneous of degree 0. That is f (ax̄) = f (x̄), for all a > 0,
and they are just functions on Sn−1 . Smooth functions on Sn−1 cor-
respond (if you like by definition) with smooth functions on Rn \ {0}
which are homogeneous of degree 0. Let us take such a function ψ ∈
C ∞ (Rn \{0}), ψ(ax) = ψ(x) for all a > 0. Now, to make this smooth on
Rn we need to cut it off near 0. So choose a cutoff function χ ∈ Cc∞ (Rn ),
with χ(x) = 1 in |x| < 1. Then
(7.2) ψR (x) = ψ(x)(1 − χ(x/R)) ∈ C ∞ (Rn ),
for any R > 0. This function is supported in |x| ≥ R. Now, if ψ has
support near some point ω ∈ Sn−1 then for R large the corresponding
function ψR will ‘localize near ω as a point at infinity of Rn .’ Rather
than try to understand this directly, let us consider a corresponding
analytic construction.
First of all, a function of the form ψR is a multiplier on S(Rn ). That
is,
(7.3) ψR · : S(Rn ) −→ S(Rn ).
To see this, the main problem is to estimate the derivatives at infinity,
since the product of smooth functions is smooth. This in turn amounts
to estimating the deriviatives of ψ in |x| ≥ 1. This we can do using the
homogeneity.
Lemma 7.1. If ψ ∈ C ∞ (Rn \ {0}) is homogeneous of degree 0 then
(7.4) |Dα ψ| ≤ Cα |x|−|α| .
Proof. I should not have even called this a lemma. By the chain
rule, the derivative of order α is a homogeneous function of degree −|α|
from which (7.4) follows.
for some seminorm on S(Rn ). Thus the map (7.3) is actually continu-
ous. This continuity means that ψR is a multiplier on S 0 (Rn ), defined
as usual by duality:
(7.7) ψR u(f ) = u(ψR f ) ∀ f ∈ S(Rn ).
Definition 7.2. The cone-support and cone-singular-support of a
tempered distribution are the subsets Csp(u) ⊂ Rn ∪ Sn−1 and Css(u) ⊂
Rn ∪ Sn−1 defined by the conditions
(7.8)
Csp(u) ∩ Rn = supp(u)
(Csp(u)){ ∩ Sn−1 ={ω ∈ Sn−1 ;
∃ R > 0, ψ ∈ C ∞ (Sn−1 ), ψ(ω) 6= 0, ψR u = 0},
Css(u) ∩ Rn = singsupp(u)
(Css(u)){ ∩ Sn−1 ={ω ∈ Sn−1 ;
∃ R > 0, ψ ∈ C ∞ (Sn−1 ), ψ(ω) 6= 0, ψR u ∈ S(Rn )}.
That is, on the Rn part these are the same sets as before but ‘at
infinity’ they are defined by conic localization on Sn−1 .
In considering Csp(u) and Css(u) it is convenient to combine Rn
and Sn−1 into a compactification of Rn . To do so (topologically) let
us identify Rn with the interior of the unit ball with respect to the
Euclidean metric using the map
x
(7.9) Rn 3 x 7−→ ∈ {y ∈ Rn ; |y| ≤ 1} = Bn .
hxi
Clearly |x| < hxi and for 0 ≤ a < 1, |x| = ahxi has only the solution
1
|x| = a/(1 − a2 ) 2 . Thus if we combine (7.9) with the identification of
Sn with the unit sphere we get an identification
(7.10) Rn ∪ Sn−1 ' Bn .
Using this identification we can, and will, regard Csp(u) and Css(u) as
subsets of Bn .12
12In fact while the topology here is correct the smooth structure on Bn is not
the right one – see Problem?? For our purposes here this issue is irrelevant.
92 3. DISTRIBUTIONS
Lemma 7.3. For any u ∈ S 0 (Rn ), Csp(u) and Css(u) are closed
subsets of Bn and if ψ̃ ∈ C ∞ (Sn ) has supp(ψ̃) ∩ Css(u) = ∅ then for R
sufficiently large ψ̃R u ∈ S(Rn ).
and
(7.12) Csp(c1 u1 + c2 u2 ) ⊂ Csp(u1 ) ∪ Csp(u2 ),
Css(c1 u1 + c2 u2 ) ⊂ Css(u1 ) ∪ Css(u2 )
∀ u1 , u2 ∈ S 0 (Rn ), c1 , c2 ∈ C.
One useful consequence of having the cone support at our disposal
is that we can discuss sufficient conditions to allow us to multiply dis-
tributions; we will get better conditions below using the same idea but
applied to the wavefront set but this preliminary discussion is used
there. In general the product of two distributions is not defined, and
indeed not definable, as a distribution. However, we can always multi-
ply an element of S 0 (Rn ) and an element of S(Rn ).
To try to understand multiplication look at the question of pairing
between two distributions.
Lemma 7.5. If Ki ⊂ Bn , i = 1, 2, are two disjoint closed (hence
compact) subsets then we can define an unambiguous pairing
(7.13)
{u ∈ S 0 (Rn ); Css(u) ⊂ K1 } × {u ∈ S 0 (Rn ); Css(u) ⊂ K2 } 3 (u1 , u2 )
−→ u1 (u2 ) ∈ C.
Proof. To define the pairing, choose a function ψ ∈ C ∞ (Sn−1 )
which is identically equal to 1 in a neighbourhood of K1 ∩Sn−1 and with
support disjoint from K2 ∩ Sn−1 . Then extend it to be homogeneous, as
above, and cut off to get ψR . If R is large enough Csp(ψR ) is disjoint
from K2 . Then ψR + (1 − ψ)R = 1 + ν where ν ∈ Cc∞ (Rn ). We can
find another function µ ∈ Cc∞ (Rn ) such that ψ1 = ψR + µ = 1 in
a neighbourhood of K1 and with Csp(ψ1 ) disjoint from K2 . Once we
have this, for u1 and u2 as in (7.13),
(7.14) ψ1 u2 ∈ S(Rn ) and (1 − ψ1 )u1 ∈ S(Rn )
since in both cases Css is empty from the definition. Thus we can define
the desired pairing between u1 and u2 by
(7.15) u1 (u2 ) = u1 (ψ1 u2 ) + u2 ((1 − ψ1 )u1 ).
Of course we should check that this definition is independent of the
cut-off function used in it. However, if we go through the definition and
choose a different function ψ 0 to start with, extend it homogeneoulsy
and cut off (probably at a different R) and then find a correction term
µ0 then the 1-parameter linear homotopy between them
(7.16) ψ1 (t) = tψ1 + (1 − t)ψ10 , t ∈ [0, 1]
94 3. DISTRIBUTIONS
(7.19) u1 ((ψ1 − ψ10 )u2 ) = u1 (χ2 (ψ1 − ψ10 )u2 ) = (χu1 )((ψ1 − ψ10 )χu2 )
= (χu2 )(ψ1 − ψ10 )χu1 ) = u2 (ψ1 − ψ10 )χu1 ).
Here the second equality is just the identity for χ as a (multiplica-
tive) linear map on S(Rn ) and hence S 0 (Rn ) and the operation to give
the crucial, third, equality is permissible because both elements are in
S(Rn ).
So, the definition is really always the same. To show that (p, q) ∈
/
WFsc (u) we need to find ‘a cutoff Φ near p’ – depending on whether
p ∈ Rn or p ∈ Sn−1 this is either Φ = φ ∈ Cc∞ (Rn ) with F = φ(p) 6= 0
or a ψR where ψ ∈ C ∞ (Sn−1 ) has ψ(p) 6= 0 – such that q ∈/ Css(Φu).
c
One crucial property is
Lemma 7.13. If (p, q) ∈ / WFsc (u) then if p ∈ Rn there exists a
neighbourhood U ⊂ Rn of p and a neighbourhood U ⊂ Bn of q such
that for all φ ∈ Cc∞ (Rn ) with support in U, U 0 ∩ Css(φu)
c = ∅; similarly
n−1
if p ∈ S then there exists a neigbourhood Ũ ⊂ Bn of p such that
U 0 ∩ Css(ψdR u) = ∅ if Csp(ωR ) ⊂ Ũ .
7. CONE SUPPORT AND WAVEFRONT SET 99
(7.45) (p, q) ∈
/ WFsc (u) ⇐⇒
u = u1 + u2 , u1 , u2 ∈ S 0 (Rn ), p ∈
/ Css(u1 ), q ∈
/ Css(ub2 ).
Proof. For given (p, q) ∈ / WFsc (u), take Φ = φ ∈ Cc∞ (Rn ) with
φ ≡ 1 near p, if p ∈ Rn or Φ = ψR with ψ ∈ C ∞ (Sn−1 ) and ψ ≡ 1
near p, if p ∈ Sn−1 . In either case p ∈
/ Css(u1 ) if u1 = (1 − Φ)u directly
from the definition. So u2 = u − u1 = Φu. If the support of Φ is small
enough it follows as in the discussion in the proof of Proposition 7.14
that
(7.46) q∈
/ Css(ub2 ).
Thus we have (7.45) in the forward direction.
7. CONE SUPPORT AND WAVEFRONT SET 101
8. Homogeneous distributions
Next time I will talk about homogeneous distributions. On R the
functions
s
s x x>0
xt =
0 x<0
where S ∈ R, is locally integrable (and hence a tempered distribution)
precisely when S > −1. As a function it is homogeneous of degree s.
Thus if a > 0 then
(ax)st = as xst .
11. SCHWARTZ SPACE. 103
−n iz·Ξ
= (2π) dΞe φ̂(Ξ)eiz·η û(η)dη
Rn
n
= (2π) φ(z)u(z).
14. SOBOLEV SPACES 107
First, take s = 0 and prove this way the, rather obvious, fact that
S is a space of multipliers on L2 . Writing out the square of the abso-
lute value of the integral as the product with the complex conjugate,
estimating by the absolute value and then using the Cauchy-Schwarz
inequality gives what we want
Z Z
| | ψ(ξ − η)û(η)dη|2 dξ
Z Z
≤ |ψ(ξ − η1 )||û(η1 )||ψ(ξ − η2 )||û(η2 )|dη1 dη2 dξ
(14.5) Z Z
≤ |ψ(ξ − η1 )||ψ(ξ − η2 )||û(η2 )|2 dη1 dη2
Z
≤ ( |ψ|)2 kuk2L2 .
1
Here, we have decomposed the integral as the product of |ψ(ξ−η1 )| 2 |û(η1 )||ψ(ξ−
1
η2 )| 2 and the same term with the η variables exchanged. The two re-
sulting factors are then the same after changing variable so there is no
square-root in the integral.
Note that what we have really shown here is the well-known result:-
Lemma 14.2. Convolution gives is a continous bilinear map
(14.6)
L1 (Rn ) × L2 (Rn ) 3 (u, v) 7−→ u ∗ v ∈ L2 (Rn ), ku ∗ vkL2 ≤ kukL1 kvkL2 .
Now, to do the general case we need to take care of the weights in
the integral for the Sobolev norm
Z
(14.7) kφukH s = (1 + |ξ|2 )s |φu(ξ)|
2 c 2
dξ.
Exchange the order of integration and note that in region II the two
variables η2 and ξ are each bounded relative to the other. Thus the
quotient of the weights is bounded above so the same argument applies
to estimate the integral by
Z 2
(14.12) C dΞ|ψ(Ξ)| kuk2H s
as desired.
The Sobolev spaces are Hilbert spaces, so their duals are (conjugate)
isomorphic to themselves. However, in view of our inclusion L2 (Rn ) ,→
S 0 (Rn ), we habitually identify
in fact the intersection here is quite a lot smaller, but nowhere near as
small as S(Rn ). To discuss decay at infinity, as will definitely want to
do, we may use weighted Sobolev spaces.
The ordinary Sobolev spaces do not effectively define decay (or
growth) at infinity. We will therefore also set
H m,l (Rn ) = {u ∈ S 0 (Rn ); hzi` u ∈ H m (Rn )}, m, ` ∈ R,
= hzi−` H m (Rn ) ,
where the second notation is supported to indicate that u ∈ H m,l (Rn )
may be written as a product hzi−` v with v ∈ H m (Rn ). Thus
0 0
H m,` (Rn ) ⊂ H m ,` (Rn ) if m ≥ m0 and ` ≥ `0 ,
so the spaces are decreasing in each index. As consequences of the
Schwartz structure theorem
[
S 0 (Rn ) = H m,` (Rn )
m,`
(15.2) \
n
S(R ) = H m,` (Rn ).
m,`
16. Multiplicativity
Of primary importance later in our treatment of non-linear prob-
lems is some version of the multliplicative property
(
H s (Rn ) ∩ L∞ (Rn ) s ≤ n2
(16.1) As (Rn ) = is a C ∞ algebra.
H s (Rn ) s > n2
Here, a C ∞ algebra is an algebra with an additional closure property.
Namely if F : RN −→ C is a C ∞ function vanishing at the origin and
u1 , . . . , uN ∈ As are real-valued then
F (u1 , . . . , un ) ∈ As .
I will only consider the case of real interest here, where s is an
integer and s > n2 . The obvious place to start is
n
Lemma 16.1. If s > 2
then
(16.2) u, v ∈ H s (Rn ) =⇒ uv ∈ H s (Rn ).
Proof. We will prove this directly in terms of convolution. Thus,
in terms of weighted Sobolev spaces u ∈ H s (Rn ) = H s,0 (Rn ) is equiva-
lent to û ∈ H 0,s (Rn ). So (16.2) is equivalent to
(16.3) u, v ∈ H 0,s (Rn ) =⇒ u ∗ v ∈ H 0,s (Rn ).
Using the density of S(Rn ) it suffices to prove the estimate
n
(16.4) ku ∗ vkH 0,s ≤ Cs kukH 0,s kvkH 0,s for s > .
2
−s 0
Now, we can write u(ζ) = hζi u etc and convert (16.4) to an estimate
on the L2 norm of
Z
−s
(16.5) hζi hξi−s u0 (ξ)hζ − ξi−s v 0 (ζ − ξ)dξ
then
(16.9)
N
Y N
Y
s n αi s−T n
ui ∈ H (R ) =⇒ U = D ui ∈ H (R ), kU kH s−T ≤ CN kui kH s .
i=1 i=1
v1 (ξ1 ) · · · hξN i−s+aN vN (ξN )hη1 i−s+a1 v̄1 (η1 ) · · · hηN i−s+aN v̄N (ηN ).
This is really an integral over R2N −1 with respect to Lebesgue measure.
Applying Cauchy-Schwarz inequality the absolute value is estimated by
Z YN X N
Y
(16.14) 2
|vi (ξi )| h ηl i2s−2T
hηi i−2s+2ai
{(ξ,η)∈R2N ;
P P
ξi = ηi } i=1 l i=1
i i
P P
The domain of integration, given by ηi = ξi , is covered by the
i i
finite number of subsets Γj on which in addition |ηj | ≥ |ηi |, for all i.
On this set we may take P the variables of integration to be ηi for i 6= j
and the ξl . Then |ηi | ≥ | ηl |/N so the second part of the integrand
l
in (16.14) is estimated by
(16.15) X Y Y Y
hηj i−2s+2aj h ηl i2s−2T hηi i−2s+2ai ≤ CN hηj i−2T +2aj hηi i−2s+2ai ≤ CN0 hηi i−2s
l i6=j i6=j i6=j
Thus the integral in (16.14) is finite and the desired estimate follows.
Proposition 16.3. If F ∈ C ∞ (Rn × R) and u ∈ H s (Rn ) for s > n
2
an integer then
s
(16.16) F (z, u(z)) ∈ Hloc (Rn ).
Proof. Since the result is local on Rn we may multiply by a com-
pactly supported function of z. In fact since u ∈ H s (Rn ) is bounded we
17. SOME BOUNDED OPERATORS 115
Elliptic Regularity
1 ∂
Dα = D1α1 . . . Dnαn , Dj = ,
i ∂zj
but it also acts on various other spaces. So, really it is just a polynomial
P (ζ) in n variables. This ‘characteristic polynomial’ has the property
that
(1.2) F(P (D)u)(ζ) = P (ζ)Fu(ζ),
which you may think of as a little square
P (D)
(1.3) S(Rn ) / S(Rn )
O O
F F
S(Rn ) / S(Rn )
P×
and this is why the Fourier tranform is especially useful. However, this
still does not solve the important questions directly.
Question 1.1. P (D) is always injective as a map (1.1) but is usu-
ally not surjective. When is it surjective? If Ω ⊂ Rn is a non-empty
open set then
(1.4) P (D) : C ∞ (Ω) −→ C ∞ (Ω)
is never injective (unless P (ζ) is constnat), for which polynomials is it
surjective?
117
118 4. ELLIPTIC REGULARITY
The first three points are relatively easy. As a map (1.1) P (D)
is injective since if P (D)u = 0 then by (1.2), P (ζ)Fu(ζ) = 0 on Rn .
However, a zero set, in Rn , of a non-trivial polynomial alwasys has
empty interior, i.e. the set where it is non-zero is dense, so Fu(ζ) = 0
on Rn (by continuity) and hence u = 0 by the invertibility of the
Fourier transform. So (1.1) is injective (of course excepting the case
that P is the zero polynomial). When is it surjective? That is, when
can every f ∈ S(Rn ) be written as P (D)u with u ∈ S(Rn )? Taking
the Fourier transform again, this is the same as asking when every
g ∈ S(Rn ) can be written in the form P (ζ)v(ζ) with v ∈ S(Rn ). If
P (ζ) has a zero in Rn then this is not possible, since P (ζ)v(ζ) always
vanishes at such a point. It is a little trickier to see the converse, that
P (ζ) 6= 0 on Rn implies that P (D) in (1.1) is surjective. Why is this
not obvious? Because we need to show that v(ζ) = g(ζ)/P (ζ) ∈ S(Rn )
whenever g ∈ S(Rn ). Certainly, v ∈ C ∞ (Rn ) but we need to show that
the derivatives decay rapidly at infinity. To do this we need to get an
estimate on the rate of decay of a non-vanishing polynomial
Lemma 1.1. If P is a polynomial such that P (ζ) 6= 0 for all ζ ∈ Rn
then there exists C > 0 and a ∈ R such that
(1.5) |P (ζ)| ≥ C(1 + |ζ|)a .
Proof. This is a form of the Tarski-Seidenberg Lemma. Stated
loosely, a semi-algebraic function has power-law bounds. Thus
(1.6) F (R) = inf{|P (ζ)|; |ζ| ≤ R}
is semi-algebraic and non-vanishing so must satisfy F (R) ≥ c(1 + R)a
for some c > 0 and a (possibly negative). This gives the desired bound.
Is there an elementary proof?
Thirdly the non-injectivity in (1.4) is obvious for the opposite rea-
son. Namely for any non-constant polynomial there exists ζ ∈ Cn such
that P (ζ) = 0. Since
(1.7) P (D)eiζ·z = P (ζ)eiζ·z
such a zero gives rise to a non-trivial element of the null space of
(1.4). You can find an extensive discussion of the density of these
sort of ‘exponential’ solutions (with polynomial factors) in all solutions
in Hörmander’s book [4].
What about the surjectivity of (1.4)? It is not always surjective
unless Ω is convex but there are decent answers, to find them you
should look under P -convexity in [4]. If P (ζ) is elliptic then (1.4) is
surjective for every open Ω; maybe I will prove this later although it is
not a result of great utility.
2. CONSTANT COEFFICIENT ELLIPTIC OPERATORS 119
With the final φu replaced by u this is just the effect of expanding out
the derivatives on the product. Namely, the ψP (D)u term is when no
122 4. ELLIPTIC REGULARITY
derivative hits ψ and the other terms come from at least one derivative
hitting ψ. Since φ = 1 on the support of ψ we may then insert φ
without changing the result. Thus the first term on the right in (2.16)
is in H s (Rn ) and all terms in the sum are in H t (Rn ) (since at most
m − 1 derivatives are involved and φu ∈ H m+t−1 (Rn ) be definition
of t). Applying the simple case discussed above it follows that ψu ∈
H m+r (Rn ) with r the minimum of s and t. This would result in the
estimate
(2.17) kψuks+m ≤ CkψP (D)uks + C 0 kφuks+m−1
provided we knew that φu ∈ H s+m−1 (since then t = s). Thus, initially
we only have this estimate with s replaced by T where T = min(s, t).
However, the only obstruction to getting the correct estimate is know-
ing that ψu ∈ H s+m−1 (Rn ).
To see this we can use a bootstrap argument. Observe that ψ can
be taken to be any smooth function with support in the interior of the
set where φ = 1. We can therefore insert a chain of functions, of any
finite (integer) length k ≥ s − t, between then, with each supported in
the region where the previous one is equal to 1 :
(2.18)
supp(ψ) ⊂ {φk = 1}◦ ⊂ supp(φk ) ⊂ · · · ⊂ supp(φ1 ) ⊂ {φ = 1}◦ ⊂ supp(φ)
where ψ and φ were our initial choices above. Then we can apply the
argument above with ψ = φ1 , then ψ = φ2 with φ replaced by φ1 and
so on. The initial regularity of φu ∈ H t+m−1 (Rn ) for some t therefore
allows us to deduce that
(2.19) φj u ∈ H m+Tj (Rn ), Tj = min(s, t + j − 1).
If k is large enough then min(s, t + k) = s so we conclude that ψu ∈
H s+m (Rn ) for any such ψ and that (2.17) holds.
Although this is a perfectly adequate proof, I will now discuss the
second method to get elliptic regularity; the main difference is that
we think more in terms of operators and avoid the explicit iteration
technique, by doing it all at once – but at the expense of a little more
thought. Namely, going back to the easy case of a tempered distibution
on Rn give the map a name:-
(2.20)
1 − χ(ζ)
Q(D) : f ∈ S 0 (Rn ) 7−→ F −1 q̂(ζ)fˆ(ζ) ∈ S 0 (Rn ), q̂(ζ) = .
P (ζ)
Here χ ∈ Cc∞ (Rn ) is chosen to be equal to one on the set |ζ| ≤ 1c + 1
corresponding to the ellipticity estimate (2.2). Thus q̂(ζ) ∈ C ∞ (Rn ) is
2. CONSTANT COEFFICIENT ELLIPTIC OPERATORS 123
linear maps
(3.3)
s+m
P (z, D) : Hloc s
(Ω) −→ Hloc (Ω), ∀ s ∈ R, P (z, D) : C ∞ (Ω) −→ C ∞ (Ω).
Now, we arrived at the estimate (2.12) in the constant coefficient
case by iteration from the case M = s + m − 1 (by nesting cutoff
functions). Pick a point z̄ ∈ Ω. In a small ball around z̄ the coefficients
are ‘almost constant’. In fact, by Taylor’s theorem,
X
(3.4) P (z, ζ) = P (z̄, ζ) + Q(z, ζ), Q(z, ζ) = (z − z̄)j Pj (z, z̄, ζ)
j
≤ δCkψukm + C 0 kφukm−1 .
What we would like to say next is that we can choose δ so small that
δC < 21 and so inserting (3.6) into (3.5) we would get
There are three main things to do. First we need to get the a priori
estimate first for general s, rather than s = 0, and then for general ψ
(since up to this point it is only for ψ with sufficiently small support).
One problem here is that in the estimates in (3.6) the L2 norm of a
product is estimated by the L∞ norm of one factor and the L2 norm
of the other. For general Sobolev norms such an estimate does not
hold, but something similar does; see Lemma 15.2. The proof of this
theorem occupies the rest of this Chapter.
Proposition 3.2. Under the hypotheses of Theorem 3.1 if in ad-
dition u ∈ C ∞ (Ω) then (3.8) follows.
Proof of Proposition 3.2. First we can generalize (3.5), now
using Lemma 15.2. Thus, if ψ has support near the point z̄
(3.9) kψuks+m ≤ CkψP (z̄, D)uks + kφQ(z, D)ψuks + C 0 kφuks+m−1
≤ CkψP (z̄, D)uks + δCkψuks+m + C 0 kφuks+m−1 .
This gives the extension of (3.7) to general s (where we are assuming
that u is indeed smooth):
(3.10) kψuks+m ≤ Cs kψP (z, D)uks + C 0 kφuks+m−1 .
Now, given a general element ψ ∈ Cc∞ (Ω) and φ ∈ Cc∞ (Ω) with φ = 1
in a neighbourhood of supp(ψ) we may choose a partition of unity ψj
with respect to supp(ψ) for each element of which (3.10) holds for some
φj ∈ Cc∞ (Ω) where in addition φ = 1 in a neighbourhood of supp(φj ).
Then, with various constants
(3.11)
X X X
kψuks+m ≤ kψj uks+m ≤ Cs kψj φP (z, D)uks +C 0 kφj φuks+m−1
j j j
00
≤ Cs (K)kφP (z, D)uks + C kφuks+m−1 ,
where K is the support of ψ and Lemma 15.2 has been used again.
This removes the restriction on supports.
Now, to get the full (a priori) estimate (3.8), where the error term
on the right has been replaced by one with arbitrarily negative Sobolev
order, it is only necessary to iterate (3.11) on a nested sequence of cutoff
functions as we did earlier in the constant coefficient case.
This completes the proof of Proposition 3.2.
So, this proves a priori estimates for solutions of the elliptic op-
erator in terms of Sobolev norms. To use these we need to show the
regularity of solutions and I will do this by constructing parametrices
in a manner very similar to the constant coefficient case.
3. INTERIOR ELLIPTIC ESTIMATES 129
is the sum of the terms which arise from at least one derivative in the
‘parameter variable’ z in q0 (which is to say ultimately the coefficients
of P (z, ζ)). We need to examine this in detail. First however notice
that we may rewrite (3.20) as
we see that
X α
(3.24) e1 (z, Z) = pα (z) Dzα−γ DZγ q0 (z, Z).
γ
|α|≤m, |γ|<m
but will be chosen to be large. Then from our identity P (z, D)Q(k) =
Id +R(k) it follows that
N
(3.42) P (z, D)uj = fj + gj , gj = R(k) fj → R(k) f in Hloc (Ω)
for k large enough depending on s and N. Thus, for k large, the right
s S
side converges in Hloc (Ω) and by (3.41), uj → u in some Hloc (Ω). But
∞
now we can use the a priori estimates (3.8) on uj ∈ C (Ω) to conclude
that
(3.43) kψuj ks+m ≤ Ckψ(fj + gj )ks + C 00 kφuj kS
to see that ψuj is bounded in H s+m (Rn ) for any ψ ∈ Cc∞ (Ω). In fact,
applied to the difference uj − ul it shows the sequence to be Cauchy.
s+m
Hence in fact u ∈ Hloc (Ω) and the estimates (3.8) hold for this u.
That is, Q(k) has the mapping property (3.39) for large k.
In fact the continuity property (3.39) holds for all Op(a) where a ∈
S m (Ω × Rn ), not just those which are parametrices for elliptic differ-
ential operators. I will comment on this below – it is one of the basic
results on pseudodifferential operators.
There is also the question of the second identity in (3.13), at least
in the same finite-order-error sense. To solve this we may use the
transpose identity. Thus taking formal transposes this second identity
should be equivalent to
(3.44) P t Qt = Id −RLt .
The transpose of P (z, D) is the differential operotor
X
(3.45) P t (z, D) = (−D)αz pα (z).
|α|≤m
1. Local diffeomorphisms
Let Ωi ⊂ Rn be open and f : Ω1 −→ Ω2 be a diffeomorphism, so it
is a C ∞ map, which is equivalent to the condition
(1.1) f ∗ u ∈ C ∞ (Ω1 ) ∀ u ∈ C ∞ (Ω2 ), f ∗ u = u ◦ f, f ∗ u(z) = u(f (z)),
and has a C ∞ inverse f −1 : Ω2 −→ Ω1 . Such a map induces an isomor-
phisms f ∗ : Cc∞ (Ω2 ) −→ Cc∞ (Ω1 ) and f ∗ : C ∞ (Ω2 ) −→ C ∞ (Ω1 ) with
inverse (f −1 )∗ = (f ∗ )−1 .
Recall also that, as a homeomorphism, f ∗ identifies the (Borel)
measurable functions on Ω2 with those on Ω1 . Since it is continuously
differentiable it also identifies L1loc (Ω2 ) with L1loc (Ω1 ) and
(1.2) Z Z
∗ ∂fi (z)
1
u ∈ Lc (Ω2 ) =⇒ f u(z)|Jf (z)|dz = u(z 0 )dz 0 , Jf (z) = det .
Ω1 Ω1 ∂zj
The absolute value appears because the definition of the Lebesgue in-
tegral is through the Lebesgue measure.
It follows that f ∗ : L2loc (Ω2 ) −→ L2loc (Ω1 ) is also an isomorphism. If
u ∈ L2 (Ω2 ) has support in some compact subset K b Ω2 then f ∗ u has
support in the compact subset f −1 (K) b Ω1 and
(1.3) Z Z
kf ∗ uk2L2 = |f ∗ u|2 dz ≤ C(K) |f ∗ u|2 |Jf (z)|dz = C(K)kuk2L2 .
Ω1 Ω1
However, the formula for the coefficients, i.e. the explicit formula for
Pf , is rather complicated:-
X X
(1.16) P = =⇒ Pf = pα (g(z 0 ))(Jf (z 0 )Dz0 )α
|α|≤m |α|≤m
Thus
(1.21) σm (Pf )(ψ 0 , ζ 0 )) = g ∗ σm (P )(f ∗ ψ 0 , z).
This allows us to ‘geometrize’ the transformation law for the leading
part (called the principal symbol) of the differential operator P. To do
2. MANIFOLDS 141
Lemma 1.2. The transformation law (1.21) shows that for any el-
ement P ∈ Diff m (Ω) the principal symbol is well-defined as an element
(1.26) σ(P ) ∈ C ∞ (T ∗ Ω)
Proof. The formula (1.19) is consistent with (1.23) and hence with
(1.21) in showing that σm (P ) is a well-defined function on T ∗ Ω.
2. Manifolds
I will only give a rather cursory discussion of manifolds here. The
main cases we are interested in are practical ones, the spheres Sn and
the balls Bn . Still, it is obviously worth thinking about the general case,
since it is the standard setting for much of modern mathematics. There
are in fact several different, but equivalent, definitions of a manifold.
142 5. COORDINATE INVARIANCE AND MANIFOLDS
Note that there are some abuses of notation here. In the first part
of (2.1) we use the fact that (Fa−1 )∗ (ρa u), defined really on Ω0a (the
image of the coordinate patch Fa : Ωa → Ω0a ∈ Rn ), vanishes outside a
compact subset and so can be unambiguously extended as zero outside
Ω0a to give a function on Rn . The second form of (2.1) is better, but
there is an equivalence relation, of equality off sets of measure zero,
which is being ignored. The definition doesn’t work well for s < 0
because u might then not be representable by a function so we don’t
know what u0 is to start with.
The most sysetematic approach is to define distributions on M first,
so we know what we are dealing with. However, there is a problem
here too, because of the transformation law (1.5) that was forced on
us by the local identification C ∞ (Ω) ⊂ C −∞ (Ω). Namely, we really
need densities on M before we can define distributions. I will discuss
densities properly later; for the moment let me use a little ruse, sticking
for simplicity to the compact case.
Definition 2.3. If M is a compact C ∞ manifold then C 0 (M ) is a
Banach space with the supremum norm and a continuous linear func-
tional
(2.2) µ : C 0 (M ) −→ R
is said to be a positive smooth measure if for every coordinate patch on
M, F : Ω −→ Ω0 there exists µF ∈ C ∞ (Ω0 ), µF > 0, such that
Z
(2.3) µ(f ) = (F −1 )∗ f µF dz ∀ f ∈ C 0 (M ) with supp(f ) ⊂ Ω.
Ω0
using Lebesgue measure in each Ω0a . The fact that this satisfies (2.3) is
similar to the exercise above.
Now, for a compact manifold, we can define a smooth positive den-
sity µ0 ∈ C ∞ (M ; Ω) as a continuous linear functional of the form
(2.5) µ0 : C 0 (M ) −→ C, µ0 (f ) = µ(ϕf ) for some ϕ ∈ C ∞ (M )
where ϕ is allowed to be complex-valued. For the moment the notation,
C ∞ (M ; Ω), is not explained. However, the choice of a fixed positive C ∞
measure allows us to identify
C ∞ (M ; Ω) 3 µ0 −→ ϕ ∈ C ∞ (M ),
meaning that this map is an isomorphism.
Lemma 2.5. For a compact manifold, M, C ∞ (M ; Ω) is a complete
metric space with the norms and distance function
kµ0 k(k) = sup |V1α1 · · · Vpα1 ϕ|
|α|≤k
∞
X kµ0 k(k)
d(µ01 , µ02 ) = 2−k
k=0
1 + kµ0 k(k)
just the coordinate vector fields cut off in Ωa . Clearly, taken together,
these span the tangent space at each point of M, i.e., the local coor-
dinate vector fields are really linear combinations of the Vi given by
renumbering the Va` . It follows that
Here the condition can be the requirement for all coordinate sys-
tems or for a covering by coordinate systems in view of the coordinate
independence of the local Sobolev spaces on Rn , that is the weaker
condition implies the stronger.
3. VECTOR BUNDLES 147
Next we will use the local elliptic estimates obtained earlier on open
sets in Rn to analyse the global invertibility properties of elliptic oper-
ators on compact manifolds. This includes at least a brief discussion
of spectral theory in the self-adjoint case.
where ϕ0a = (Fa−1 )∗ ϕa and similarly for ψa0 (Fa−1 )∗ ϕa ∈ Cc∞ (Ω0a ), are
the local coordinate representations. We know that (1.3) holds for
every v ∈ C −∞ (Ω0a ) such that Pa v ∈ Hloc M
(Ω0a ). Applying (1.3) to
−1 ∗
(Fa ) u = va , for u ∈ H (M ), it follows that P u ∈ H s (M ) implies
M
M
Pa va ∈ Hloc (Ω0a ), by coordinate-invariance of the Sobolev spaces and
then conversely
s+m
va ∈ Hloc (Ω0a ) ∀ a =⇒ u ∈ H s+m (M ).
The norm on H s (M ) can be taken to be
!1/2
X
kuks = k(Fa−1 )∗ (ϕa u)k2s
a
P = .. .. .
. .
Pn1 (z, Dz ) · · · Pn` (z, Dz )
The (usual) order of P is the maximum of the orders of the Pij (z, D3 )
and the symbol is just the corresponding matrix of symbols
σm (P11 )(z, ζ) · · · σm (P1` )(z, ζ)
(1.4) σm (P )(z, ζ) = .. .. .
. .
σm (Pn1 )(z, ζ) · · · σm (Pn` )(z, ζ)
Such a P is said to be elliptic at z if this matrix is invariable for all
ζ 6= 0, ζ ∈ Rn . Of course this implies that the matrix is square, so the
two vector bundles have the same rank, `. As a differential operator,
P ∈ Diff m (M, E), E = E1 , E2 , is elliptic if it is elliptic at each point.
1. GLOBAL ELLIPTIC ESTIMATES 151
Once we have this local result we easily deduce the global result.
Proposition 2.2. On a compact manifold the inclusion H s (M ) ,→
H t (M ), for any s > t, is compact.
Proof. If φi ∈ Cc∞ (Ui ) is a partition of unity subordinate to an
open cover of M by coordinate patches gi : Ui −→ Ui0 ⊂ Rn , then
(2.6) u ∈ H s (M ) =⇒ (gi−1 )∗ φi u ∈ H s (Rn ), supp((gi−1 )∗ φi u) b Ui0 .
Thus if un is a bounded sequence in H s (M ) then the (gi−1 )∗ φi un form
a bounded sequence in H s (Rn ) with fixed compact supports. It follows
from Lemma 2.1 that we may choose a subsequence so that each φi unj
converges in H t (Rn ). Hence the subsequence unj converges in H t (M ).
4. Generalized inverses
Written, at least in part, by Chris Kottke.
As discussed above, a bounded operator between Hilbert spaces,
T : H1 −→ H2
is Fredholm if and only if it has a parametrix up to compact errors,
that is, there exists an operator
S : H2 −→ H1
such that
T S − Id2 = R2 , ST − Id1 = R1
are both compact on the respective Hilbert spaces H1 and H2 . In this
case of Hilbert spaces there is a “preferred” parametrix or generalized
inverse.
Recall that the adjoint
T ∗ : H2 −→ H1
of any bounded operator is defined using the Riesz Representation The-
orem. Thus, by the continuity of T , for any u ∈ H2 ,
H1 3 φ −→ hT φ, ui ∈ C
is continuous and so there exists a unique v ∈ H1 such that
hT φ, ui2 = hφ, vi1 , ∀ φ ∈ H1 .
Thus v is determined by u and the resulting map
H2 3 u 7→ v = T ∗ u ∈ H1
is easily seen to be continuous giving the adjoint identity
(4.1) hT φ, ui = hφ, T ∗ ui, ∀ φ ∈ H1 , u ∈ H2
In particular it is always the case that
(4.2) Nul(T ∗ ) = (Ran(T ))⊥
158 6. INVERTIBILITY OF ELLIPTIC OPERATORS
so that T̃ is invertible.
Lemma 4.3. Suppose T ∈ Frk (H1 , H2 ) has kernel N1 ⊂ H1 and
M1 ⊃ N1 is a finite dimensional subspace of H1 then defining T 0 = T
on M1⊥ and T 0 = 0 on M1 gives an element T 0 ∈ Frk .
Proof. Since N1 ⊂ M1 , T 0 is obtained from (4.6) by replacing
T̃ by T̃ 0 which is defined in essentially the same way as T 0 , that is
T̃ 0 = 0 on M1 /N1 , and T̃ 0 = T̃ on the orthocomplement. Thus the
range of T̃ 0 in Ran(T ) has complement T̃ (M1 /N1 ) which has the same
dimension as M1 /N1 . Thus T 0 has null space M1 and has range in H2
with complement of dimension that of M1 /N1 + N2 , and hence has
index k.
Lemma 4.4. If A is a finite rank operator A : H1 −→ H2 such that
Ran A ∩ Ran T = {0}, then T + A ∈ Frk .
Proof. First note that Nul(T + A) = Nul T ∩ Nul A since
x ∈ Nul(T +A) ⇔ T x = −Ax ∈ Ran T ∩Ran A = {0} ⇔ x ∈ Nul T ∩Nul A.
Similarly the range of T +A restricted to Nul T meets the range of T +A
restricted to (null T )⊥ only in 0 so the codimension of the Ran(T + A)
is the codimension of Ran AN where AN is A as a map from Nul T to
H2 / Ran T. So, the equality of row and column rank for matrices,
codim Ran(T +A) = codim Ran T −dim Nul(AN ) = dim Nul(T )−k−dim Nul(AN ) = dim Nul(T +
Thus T + A ∈ Frk .
Proposition 4.5. If A : H1 −→ H2 is any finite rank operator,
then T + A ∈ Frk .
Proof. Let E2 = Ran A ∩ Ran T , which is finite dimensional, then
E1 = T̃ −1 (E2 ) has the same dimension. Put M1 = E1 ⊕ N1 and apply
Lemma 4.3 to get T 0 ∈ Frk with kernel M1 . Then
T + A = T 0 + A0 + A
where A0 = T on E1 and A0 = 0 on E1⊥ . Then A0 + A is a finite rank
operator and Ran(A0 + A) ∩ Ran T 0 = {0} and Lemma 4.4 applies.
Thus
T + A = T 0 + (A0 + A) ∈ Frk (H1 , H2 ).
Proposition 4.6. If B : H1 −→ H2 is compact then T + B ∈ Frk .
160 6. INVERTIBILITY OF ELLIPTIC OPERATORS
The matrix Pij (ζ) is invertible if and only if k = l and the polyno-
mial of order mk, det P (ζ) 6= 0. Such a matrix is said to be elliptic if
det P (ζ) is elliptic. The cofactor matrix defines a matrix P 0 of differ-
ential operators of order (k − 1)m and we may construct a parametrix
for P (assuming it to be elliptic) from a parametrix for det P :
(5.14) QP = Qdet P P 0 (D).
It is then easy to see that it has the same mapping properties as in the
case of a single operator (although notice that the product is no longer
commutative because of the non-commutativity of matrix multiplica-
tion)
(5.15) QP P = Id −RL , P QP = Id −RR
164 6. INVERTIBILITY OF ELLIPTIC OPERATORS
since the sum is finite on each bounded set. Moreover F (t) ≥ 1 and is
itself invariant under translation by any integer; set µ(t) = µ1 (t)/F (t).
Then µ generates a ti-partition of unity.
Using such a function we can easily decompose L2 (R). Thus, setting
τk (t) = t − k,
(1.12)
XZ
2 ∗ 2
f ∈ L (R) ⇐⇒ (τk f )µ ∈ Lloc (R) ∀ k ∈ Z and |τk∗ f µ|2 dt < ∞.
k∈Z
Now, Dt τk∗ f ∗
= τk (Dt f ), so we can use (1.12) torewrite the definition
of thespaces Htik (R × M ) in a form that extendsto all orders. Namely
(1.13) X
u ∈ Htis (R × M ) ⇐⇒ (τk∗ u)µ ∈ Hcs (R × M ) and kτk∗ ukH s < ∞
k
Lemma 1.2. With Diff m ti (R×M ) defined by (1.10) and the translation-
invariant Sobolev spaces by (1.13),
P ∈ Diff m
ti (R × M ) =⇒
(1.14)
P :Htis+m (R × M ) −→ Htis (R × M ) ∀ s ∈ R.
(1.15) u ∈ Htis+m (R × M ) =⇒
0 0
X
P (τk∗ uµ) = τk∗ (P u) + τk∗ (Pm0 u)Dtm−m µ, Pm0 ∈ Diff m
ti (R × M ).
m0 <m
The left side is in Htis (R × M ), with the sum over k of the squares
of the norms bounded, by the regularity of u. The same is easily seen
to be true for the sum on the right by the inductive hypothesis, and
hence for the first term on the right. This proves the mapping property
(1.14) and continuity follows by the same argument or the closed graph
theorem.
where on the right we are using the Sobolev norms on Rn . Thus, ap-
plying the estimates for Euclidean space to each term on the right we
get the same estimate on any compact manifold.
Corollary 1.4. If u ∈ H̃tis (R × M ), for s > 0, then for any
0<t<s
Z
(1.21) hτ i2t kû(τ, ·)k2H s−t (M ) dτ < ∞
R
(1.22)
|u(t, z) − u(t0 , z)|2 |u(t, z 0 ) − u(t, z)|2
Z Z
0
0 2s+1
dtdt ν+ s+ n dtν(z)ν(z 0 ) < ∞,
R2 ×M |t − t | R×M 2 ρ(z, z ) 0 2
n = dim M,
where 0 ≤ ρ ∈ C ∞ (M 2 ) vanishes exactly quadratically at Diag ⊂ M 2 .
Proof. This follows as in the cases of Rn and a compact manifold
discussed earlier since the second term in (1.22) gives (with the L2
norm) a norm on L2 (R; H s (M )) and the first term gives a norm on
L2 (M ; H s (R)).
Using these results we can see directly that the Sobolev spaces in
(1.16) have the following ‘obvious’ property as in the cases of Rn and
M.
Lemma 1.7. Schwartz space S(R × M ) = C ∞ (M ; S(R)) is dense in
each Htis (R × M ) and the L2 pairing extends by continuity to a jointly
continuous non-degenerate pairing
(1.23) Htis (R × M ) × Hti−s (R × M ) −→ C
which identifies Hti−s (R×M ) with the dual of Htis (R×M ) for any s ∈ R.
Proof. I leave the density as an exercise – use convolution in R
and the density of C ∞ (M ) in H s (M ) (explicity, using a partition of
unity on M and convolution on Rn to get density in each coordinate
patch).
Then the existence and continuity of the pairing follows from the
definitions and the corresponding pairings on R and M. We can assume
that s > 0 in (1.23) (otherwise reverse the factors). Then if u ∈
Htis (R × M ) and v = v1 + v2 ∈ Hti−s (R × M ) as in (1.17),
Z Z
(1.24) (u, v) = (u(t, ·), u1 (t, ·)) dt + (u(·, z), v2 (·, z)) νz
R M
174 7. SUSPENDED FAMILIES AND THE RESOLVENT
2. Translation-invariant Operators
Some corrections from Fang Wang added, 25 July, 2007.
Next I will characterize those operators P ∈ Diff m
ti (R×M ; E) which
give invertible maps (1.14), or rather in the case of a pair of vector
bundles E = (E1 , E2 ) over M :
(2.1) P : Htis+m (R×M ; E1 ) −→ Htis (R×M ; E2 ), P ∈ Diff m
ti (R×M ; E).
then
P : S(R × M ; E1 ) −→ S(R × M ; E2 )
and
m
X
(2.3) Pcu(τ, ·) = τ i Pi u
b(τ, ·)
i=0
(2.10) Pa = .. ..
. .
P1l (z, Dt , Dz ) · · · Pll (z, Dt , Dz )
in which the coefficients do not depend on t. As discussed in Sections 2
and 3 above, we can construct a local parametrix in Ω0a using a properly
supported cutoff χ. In the t variable the parametrix is global anyway,
so we use a fixed cutoff χ̃ ∈ Cc∞ (R), χ̃ = 1 in |t| < 1, and so construct
a parametrix
Z
(2.11) Qa f (t, z) = q(t − t0 , z, z 0 )χ̃(t − t0 )χ(z, z 0 )f (t0 , z 0 ) dt0 dz 0 .
Ω0a
This satisfies
(2.12) Pa Qa = Id −Ra , Qa Pa = Id −Ra0
where Ra and Ra0 are smoothing operators on Ω0a with kernels of the
form
Z
Ra f (t, z) = Ra (t − t0 , z, z 0 )f (t0 , z 0 ) dt0 dz 0
(2.13) Ω0a
Ra ∈C (R × Ω02
∞ 0
a ), Ra (t, z, z ) = 0 if |t| ≥ 2
Then
X
τk∗ Q(µτ−k
∗
(2.20) Qf = f) .
k∈Z
from which all the statements follow just as in the standard case when
R ∈ Cc∞ (R) has support in [−A, A].
Proposition 2.5. If R is as in Lemma−1 2.4 then there exists a
discrete subset D ⊂ C such that Id −R(τ
b ) exists for all τ ∈ C \ D
and
−1
(2.25) Id −R(τ )
b = Id −S(τ
b )
(2.28) kR(τ
b )kL2 ≤ 1/2.
Thus, by Neumann series,
∞
X k
(2.29) S(τ
b )= R(τ )
b
k=1
where Rb1 (τ ), R
b2 (τ ) are families of smoothing operators as in Proposi-
tion 2.5. Applying that result to the first equation gives a new mero-
morphic right inverse
Q(τ
b )=Q b0 (τ )(Id −R
b2 (τ ))−1 = Q
b0 (τ ) − Q
b0 (τ )M (τ )
where the first term is entire and the second is a meromorphic family
of smoothing operators with finite rank residues. The same argument
on the second term gives a left inverse, but his shows that Q(τb ) must
be a two-sided inverse.
This we have proved everything except the locations of the poles of
Q(τ
b ) — which are only constrained by (2.26) instead of (2.6). However,
we can apply the same argument to Pθ (z, Dt , Dz ) = P (z, eiθ Dt , Dz ) for
180 7. SUSPENDED FAMILIES AND THE RESOLVENT
|θ| < δ, δ > 0 small, since Pθ stays elliptic. This shows that the poles
of Q(τ
b ) lie in a set of the form (2.6).
3. Invertibility
We are now in a position to characterize those t-translation-invariant
differential operators which give isomorphisms on the translation-invariant
Sobolev spaces.
Theorem 3.1. An element P ∈ Diff m ti (R × M ; E) gives an isomor-
phism (2.1) (or equivalently is Fredholm) if and only if it is elliptic and
D ∩ R = ∅, i.e. P̂ (τ ) is invertible for all τ ∈ R.
Proof. We have already done most of the work for the important
direction for applications, which is that the ellipticity of P and the
invertibility at P̂ (τ ) for all τ ∈ R together imply that (2.1) is an
isomorphism for any s ∈ R.
Recall that the ellipticity of P leads to a parameterix Q which is
translation-invariant and has the mapping property we want, namely
(2.17).
To prove the same estimate for the true inverse (and its existence)
consider the difference
(3.1) P̂ (τ )−1 − Q̂(τ ) = R̂(τ ), τ ∈ R.
Since P̂ (τ ) ∈ Diff m (M ; E) depends smoothly on τ ∈ R and Q̂(τ ) is a
paramaterix for it, we know that
(3.2) R̂(τ ) ∈ C ∞ (R; Ψ−∞ (M ; E))
is a smoothing operator on M which depends smoothly on τ ∈ R as a
parameter. On the other hand, from (2.32) we also know that for large
real τ,
P̂ (τ )−1 − Q̂(τ ) = Q̂(τ )M (τ )
where M (τ ) satisfies the estimates (2.27). It follows that Q̂(τ )M (τ )
also satisfies these estimates and (3.2) can be strengthened to
(3.3) sup kτ k R̂(τ, ·, ·)kC p < ∞ ∀ p, k.
τ ∈R
It certainly suffices to show this for s < 0 and then we know that the
Fourier transform gives a map
(3.6) F : Htis (R × M ; E2 ) −→ hτ i|s| L2 (R; H −|s| (M ; E2 )).
Since the kernel R̂(τ ) is rapidly decreasing in τ, as well as being smooth,
for every N > 0,
(3.7) R̂(τ ) : hτ i|s| L2 (R; H −|s| M ; E2 ) −→ hτ i−N L2 (R; H N (M ; E2 ))
and inverse Fourier transform maps
F −1 : hτ i−N H N (M ; E2 ) −→ HtiN (R × M ; E2 )
which gives (3.5).
Thus Q + R has the same property as Q in (2.17). So it only
remains to check that Q + R is the two-sided version of P and it is
enough to do this on S(R × M ; Ei ) since these subspaces are dense
in the Sobolev spaces. This in turn follows from (3.1) by taking the
Fourier transform. Thus we have shown that the invertibility of P
follows from its ellipticity and the invertibility of P̂ (τ ) for τ ∈ R.
The converse statement is less important but certainly worth know-
ing! If P is an isomorphism as in (2.1), even for one value of s, then
it must be elliptic — this follows as in the compact case since it is
everywhere a local statement. Then if P̂ (τ ) is not invertible for some
τ ∈ R we know, by ellipticity, that it is Fredholm and, by the stability
of the index, of index zero (since P̂ (τ ) is invertible for a dense set of
τ ∈ C). There is therefore some τ0 ∈ R and f0 ∈ C ∞ (M ; E2 ), f0 6= 0,
such that
(3.8) P̂ (τ0 )∗ f0 = 0.
It follows that f0 is not in the range of P̂ (τ0 ). Then, choose a cut off
function, ρ ∈ Cc∞ (R) with ρ(τ0 ) = 1 (and supported sufficiently close
to τ0 ) and define f ∈ S(R × M ; E2 ) by
(3.9) fˆ(τ, ·) = ρ(τ )f0 (·).
Then f ∈ / P · Htis (R × M ; E1 ) for any s ∈ R. To see this, suppose
s
u ∈ Hti (R × M ; E1 ) has
(3.10) P u = f ⇒ P̂ (τ )û(τ ) = fˆ(τ )
where û(τ ) ∈ hτ i|s| L2 (R; H −|s| (M ; E1 )). The invertibility of P (τ ) for
τ 6= τ0 on supp(ρ) (chosen with support close enough to τ0 ) shows that
û(τ ) = P̂ (τ )−1 fˆ(τ ) ∈ C ∞ ((R\{τ0 }) × M ; E1 ).
182 7. SUSPENDED FAMILIES AND THE RESOLVENT
whereas if a + b ≤ 0 then
s,a,b
(3.21) Hti,exp (R × M ; E) = eat Htis (R × M ; E) ∩ e−bt Htis (R × M ; E).
(2) Show that any P ∈ Diff m
ti (R×M ; E) defines a continuous linear
map for any s, a, b ∈ R
s+m,a,b s,a,b
(3.22) P : Hti- exp (R × M ; E1 ) −→ Hti- exp (R × M ; E2 ).
(3) Show that the standard L2 pairing, with respect to dt, a smooth
positive density on M and an inner product on E extends to
a non-degenerate bilinear pairing
s,a,b −s,−a,−b
(3.23) Hti,exp (R × M ; E) × Hti,exp (R × M ; E) −→ C
for any s, a and b. Show that the adjoint of P with respect to
this pairing is P ∗ on the ‘negative’ spaces – you can use this
to halve the work below.
(4) Show that if P is elliptic then (3.22) is Fredholm precisely
when
(3.24) a∈
/ − Im(D) and b ∈
/ Im(D).
Hint:- Assume for instance that a+b ≥ 0 and use (3.20). Given
(3.24) a parametrix for P can be constructed by combining the
inverses on the single exponential spaces
−1
(3.25) Qa,b = χ0 Pa−1 χ + (1 − χ00 )P−b (1 − χ)
where χ is as in (3.19) and χ0 and χ00 are similar but such that
χ0 χ = 1, (1 − χ00 )(1 − χ) = 1 − χ.
(5) Show that P is an isomorphism if and only if
a+b ≤ 0 and [a, −b]∩− Im(D) = ∅ or a+b ≥ 0 and [−b, a]∩− Im(D) = ∅.
(6) Show that if a + b ≤ 0 and (3.24) holds then
X
ind(P ) = dim null(P ) = Mult(P, τi )
τi ∈D∩(R×[b,−a])
4. Resolvent operator
Addenda to Chapter 7
More?
• Why – manifold with boundary later for Euclidean space, but
also resolvent (Photo-C5-01)
• Hölder type estimates – Photo-C5-03. Gives interpolation.
As already noted even a result such as Proposition 3.3 and the
results in the exercises above by no means exhausts the possibile real-
izations of an element P ∈ Diff mti (R × M ; E) as a Fredholm operator.
Necessarily these other realization cannot simply be between spaces
like those in (3.19). To see what else one can do, suppose that the
condition in Theorem 3.1 is violated, so
(4.1) D(P ) ∩ R = {τ1 , . . . , τN } =
6 ∅.
To get a Fredholm operator we need to change either the domain or the
range space. Suppose we want the range to be L2 (R×M ; E2 ). Now, the
condition (3.24) guarantees that P is Fredholm as an operator (3.22).
So in particular
m,, 0,,
(4.2) P : Hti−exp (R × M ; E1 ) −→ Hti- exp (R × M ; E2 )
1. Compactifications of R.
As I will try to show by example later in the course, there are
I believe considerable advantages to looking at compactifications of
non-compact spaces. These advantages show up last in geometric and
analytic considerations. Let me start with the simplest possible case,
namely the real line. There are two standard compactifications which
one can think of as ‘exponential’ and ‘projective’. Since there is only
one connected compact manifold with boundary compactification cor-
responds to the choice of a diffeomorphism onto the interior of [0, 1]:
which both have images (0, x± ) and as diffeomorphism other than signs.
In fact if we want the two ends to be the ‘same’ then we can take
γ− (t) = γ+ (−t). I leave it as an exercise to show that γ then exists
with
(
γ(t) = γ+ (t) t0
(1.3)
γ(t) = 1 − γ− (t) t 0.
are
x = e−t exponential compactification
(CR.4)
x = 1/t projective compactification .
Note that these are alternatives!
Rather than just consider R, I want to consider R × M, with M
compact, as discussed above.
Lemma 1.1. If R : H −→ H is a compact operator on a Hilbert
space then Id −R is Fredholm.
Proof. A compact operator is one which maps the unit ball (and
hence any bounded subset) of H onto a precompact set, a set with
compact closure. The unit ball in the null space of Id −R is
{u ∈ H; kuk = 1 , u = Ru} ⊂ R{u ∈ H; kuk = 1}
and is therefore precompact. Since it is closed, it is compact and any
Hilbert space with a compact unit ball is finite dimensional. Thus the
null space of (Id −R) is finite dimensional.
Consider a sequence un = vn − Rvn in the range of Id −R and
suppose un → u in H. We may assume u 6= 0, since 0 is in the range,
and by passing to a subsequence suppose that of γ on ?? fields. Clearly
γ(t) = e−t ⇒ γ∗ (∂t ) = −x(∂x )
(CR.5)
γ̃(t) = 1/t ⇒ γ̃∗ (∂t ) = −s2 ∂s
where I use ‘s’ for the variable in the second case to try to reduce
confusion, it is just a variable in [0, 1]. Dually
∗ dx
γ = −dt
x
(CR.6)
∗ ds
γ̃ = −dt
s2
in the two cases. The minus signs just come from the fact that both
γ’s reverse orientation.
Proposition 1.2. Under exponential compactification the translation-
invariant Sobolev spaces on R × M are identified with
(1.4)
k 2 dx
Hb ([0, 1] × M ) = u ∈ L [0, 1] × M ; VM ; ∀ `, p ≤ k
x
p ` 2 dx
Pp ∈ Diff (M ) , (xDx ) Pp u ∈ L [0, 1] × M ; VM
x
1. COMPACTIFICATIONS OF R. 189
X
(1.6) Hbs ([0, 1] × M ) = u = (XdJX )pP uj,p ,
0≤j+p≤k
Note that there are other properties I have not translated into this
new setting. There is one additional fact which it is easy to check.
Namely C ∞ ([0, 1] × M ) acts as multipliers on all the spaces Hbs ([0, 1] ×
M ). This follows directly from Proposition 1.2;
(CR.12)
C ∞ ([0, 1] × M ) × Hbs ([0, 1] × M ) 3 (ϕ, u) 7→ ϕu ∈ Hbs ([0, 1] × M ) .
What about the ‘b’ notation? Notice that (1−x)x∂x and the smooth
vector fields on M span, over C ∞ (X), for X = [0, 1] × M , all the vector
fields tangent to {x = 0|u|x = 1}. Thus we can define the ‘boundary
differential operators’ as
(CR.13)
(
X
Diff m E
b ([0, 1] × Mi ) = P = aj,p (xj )((1 − x)xDx )j Pp ,
0≤j+p≤m
Pp ∈ Diff p (Mi )E
190 8. MANIFOLDS WITH BOUNDARY
Namely
(CR.19) kukm+s ≤ Cs kP± uks
where these norms, as in (CR.17) are in the Hbs spaces. Note that
near x = 0 or x = 1, P± are obtained by substituting Dt 7→ xDx or
(1 − x)Dx in (CR.17). Thus
(CR.20) P − P± ∈ (x − x± ) Diff m
b (X) , x± = 0, 1
have coefficients which vanish at the appropriate boundary. This is
precisely how (CR.16) is derived from (CR.13). Now choose ϕ ∈
2. BASIC PROPERTIES 191
Notice that we do know that the coefficients are smooth in 0 < x < δ,
since we are applying a diffeomorphism there. Moreover, the operators
P`,i,j are uniquely determined by (5.7).
So we can exploit the assumed homogeneity of Pij . This means that
for any t > 0, the transformation z 7→ tz gives
(5.8) Pij f (tz) = tm (Pij f )(tz) .
Since |tz| = t|z|, this means that the transformed operator must satisfy
(5.9)
X X
Dx` P`,i,j (x, θ, Dθ )f (x/t, θ) = tm ( D` P`,i,j (·, θ, Dθ )f (·, θ))(x/t) .
` `
6. Scattering structure
Let me briefly review how the main result of Section 5 was arrived
at. To deal with a constant coefficient Dirac operator we first radially
compactified Rn to a ball, then peeled off a multiplicative factor Z0
from the operator showed that the remaining operator was Fredholm by
identifing a neighbourhood of the boundary with part of R×Sn−1 using
the exponential map to exploit the results of Section 1 near infinity.
Here we will use a similar, but different, procedure to treat a different
class of operators which are Fredholm on the standard Sobolev spaces.
Although we will only apply this in the case of a ball, coming from
Rn , I cannot resist carrying out the discussed for a general compact
manifolds — since I think the generality clarifies what is going on.
Starting from a compact manifold with boundary, M, the first step is
essentially the reverse of the radial compactification of Rn .
Near any point on the boundary, p ∈ ∂M, we can introduce ‘ad-
missible’ coordinates, x, y1 , . . . , yn−1 where {x = 0k is the local form of
the boundary and y1 , . . . , yn−1 are tangential coordinates; we normalize
y1 = · · · = yn−1 = 0 at p. By reversing the radial compactification of
Rn I mean we can introduce a diffeomorphism of a neighbourhood of p
to a conic set in Rn :
(6.1) zn = 1/x , zj = yj /x , j = 1, . . . , n − 1 .
Clearly the ‘square’ |y| < , 0 < x < is mapped onto the truncated
conic set
(6.2) zn ≥ 1/ , |z 0 | < |zn | , z 0 = (z1 , . . . , zn−1 ) .
196 8. MANIFOLDS WITH BOUNDARY
s
Definition 6.1. We define spaces Hsc (M ) for any compact mani-
fold with boundary M by the requirements
(6.3) s
u ∈ Hsc s
(M ) ⇐⇒ u ∈ Hloc (M \ ∂M ) and Rj∗ (ϕj u) ∈ H s (Rn )
for ϕj ∈ C ∞ (M ), 0 ≤ ϕi ≤ 1,
P
ϕi = 1 in a neighbourhood of the
boundary and where each ϕj is supported in a coordinate patch (??),
(6.2) with R given by (6.1).
Of course such a definition would not make much sense if it de-
pended on the choice of the partition of unity near the boundary {ϕi k
or the choice of coordinate. So really (6.1) should be preceded by such
an invariance statement. The key to this is the following observation.
Proposition 6.2. If we set Vsc (M ) = xVb (M ) for any compact
manifold with boundary then for any ψ ∈ C ∞ (M ) supported in a coor-
dinate patch (??), and any C ∞ vector field V on M
Xn
(6.4) ψV ∈ Vsc (M ) ⇐⇒ ψV = µj (R−1 )∗ (Dzj ) , µj ∈ C ∞ (M ) .
j=1
Lemma 6.4. If (6.3) holds for s = k ∈ N for any one such partition
0
of unity then u ∈ Hsc (M ) in the sense of (6.6) and
0
(6.7) V1 . . . Vj u ∈ Hsc (M ) ∀ Vi ∈ Vsc (M ) if j ≤ k ,
and conversely.
Proof. For clarity we can proceed by induction on k and re-
k−1 k−1
place (6.7) by the statements that u ∈ Hsc (M ) and V u ∈ Hsc (M )
∀V ∈ Vsc (M ). In the interior this is clear and follows immediately from
Proposition 6.2 provided we carry along the inductive statement that
(6.8) C ∞ (M ) acts by multiplication on Hsc
k
(M ) .
As usual we can pass to general s ∈ R by treating the cases 0 <
s < 1 first and then using the action of the vector fields.
Proposition 6.5. For 0 < s < 1 the condition (6.3) (for any one
0
partition of unity) is equivalent to requiring u ∈ Hsc (M ) and
|u(p) − u(p0 )|2 νM 0
ZZ
νM
(6.9) <∞
M ×M ρn+2s
sc xn+1 (x0 )n+1
where ρsc (p, p0 ) = χχ0 p(p, p0 ) + j ϕj ϕ0j hz − z 0 i.
P
Note that this is not in general homogeneous since the lower order terms
are retained. Despite this one gets essentially the same polynomial at
each point, independent of the admissible coordinates chosen, as will
be shown below. Let’s just assume this for the moment so that the
condition in the following result makes sense.
Theorem 6.7. If P ∈ Diff m sc (M ; E) acts between vector bundles
over M, is elliptic in the interior and each of the polynomials (matrices)
(6.14) is elliptic and has no real zeros then
s+m s
(6.15) P : Hsc (M, E1 ) −→ Hsc (M ; E2 ) is Fredholm
for each s ∈ R and conversely.
Last time at the end I gave the following definition and theorem.
Definition 6.8. We define weighted (non-standard) Sobolev spaces
for (m, w) ∈ R2 on Rn by
(6.16)
H̃ m,w (Rn ) = {u ∈ Mloc
m
(Rn ); F ∗ (1 − χ)r−w u ∈ Htim (R × Sn−1 )}
n
P
Theorem 6.9. If P = Γi Di , Γi ∈ M (N, C), is an elliptic, con-
i=1
stant coefficient, homogeneous differential operator of first order then
(6.18) P : H̃ m,w (Rn ) −→ H̃ m−1,w+1 (Rn ) ∀ (m, w) ∈ R2
is continuous and is Fredholm for w ∈ R \ D̃ where D̃ is discrete.
If P is a Dirac operators, which is to say explicitly here that the
coefficients are ‘Pauli matrices’ in the sense that
(6.19) Γ∗i = Γi , Γ2i = IdN ×N , ∀ i, Γi Γj + Γj Γi = 0, i 6= j,
then
(6.20) D̃ = −N0 ∪ (n − 2 + N0 )
and if n > 2 then for w ∈ (0, n − 2) the operator P in (6.18) is an
isomorphism.
I also proved the following result from which this is derived
Lemma 6.10. In polar coordinates on Rn in which Rn \ {0} '
(0, ∞) × Sn−1 , y = rθ,
(6.21) Dyj =
CHAPTER 9
Electromagnetism
1. Maxwell’s equations
Maxwell’s equations in a vacuum take the standard form
div E = ρ div B = 0
(1.1) ∂B ∂E
curl E = − curl B = +J
∂t ∂t
where E is the electric and B the magnetic field strength, both are
3-vectors depending on position z ∈ R3 and time t ∈ R. The external
quantities are ρ, the charge density which is a scalar, and J, the current
density which is a vector.
We will be interested here in stationary solutions for which E and
B are independent of time and with J = 0, since this also represents
motion in the standard description. Thus we arrive at
div E = ρ div B = 0
(1.2)
curl E = 0 curl B = 0.
The simplest interesting solutions represent charged particles, say
with the charge at the origin, ρ = cδ0 (z), and with no magnetic field,
B = 0. By identifying E with a 1-form, instead of a vector field on R3 ,
(1.3) E = (E1 , E2 , E3 ) =⇒ e = E1 dz1 + E2 dz2 + E3 dz3
we may identify curl E with the 2-form de,
(1.4) de =
∂E2 ∂E1 ∂E3 ∂E2 ∂E1 ∂E3
− dz1 ∧dz2 + − dz2 ∧dz3 + − dz3 ∧dz1 .
∂z1 ∂z2 ∂z2 ∂z3 ∂z3 ∂z1
Thus (1.2) implies that e is a closed 1-form, satisfying
∂E1 ∂E2 ∂E3
(1.5) + + = cδ0 (z).
∂z1 ∂z2 ∂z3
By the Poincaré Lemma, a closed 1-form on R3 is exact, e = dp,
with p determined up to an additive constant. If e is smooth (which it
201
202 9. ELECTROMAGNETISM
Proof. The only solutions are of the form (1.19) where q ∈ S 0 (R3 )
is harmonic. Thus qb ∈ S 0 (R3 ) satisfies |ξ|2 qb = 0, which implies that q
is a polynomial.
2. Hodge Theory
The Hodge ∗ operator discussed briefly above in the case of R3 (and
Minkowski 4-space) makes sense in any oriented real vector space, V,
with a Euclidean inner product—that is, on a finite dimensional real
Hilbert space. Namely, if e1 , . . . , en is an oriented orthonormal basis
then
(2.1) ∗(ei1 ∧ · · · ∧ eik ) = sgn(i∗ )eik+1 ∧ · · · ein
extends by linearity to
Vk Vn−k
(2.2) ∗: V −→ V.
Proposition 2.1. The linear map (2.2) is independent of the ori-
ented orthonormal basis used to define it and so depends only on the
choice of inner product and orientation of V. Moreover,
∗2 = (−1)k(n−k) , on k V .
V
(2.3)
Proof. Note that sgn(i∗ ), the sign of the permutation defined by
{i1 , . . . , in } is fixed by
(2.4) ei1 ∧ · · · ∧ ein = sgn(i∗ )e1 ∧ · · · ∧ en .
Thus, on the basis ei1 ∧ . . . ∧ ein of k V given by strictly increasing
V
sequences i1 < i2 < · · · < ik in {1, . . . , n},
(2.5) e∗ ∧ ∗e∗ = sgn(i∗ )2 e1 ∧ · · · ∧ en = e1 ∧ · · · ∧ en .
The standard inner product on k V is chosen so that this basis is
V
orthonormal. Then (2.5) can be rewritten
(2.6) eI ∧ ∗eJ = heI , eJ ie1 ∧ · · · ∧ en .
This in turn fixes ∗ uniquely since the pairing given by
Vk
V × k−1 V 3 (u, v) 7→ (u ∧ v)/e1 ∧···∧en
V
(2.7)
is non-degenerate, as can be checked on these bases.
Thus it follows from (2.6) that ∗ depends only on the choice of inner
product andV orientation as claimed, provided it is shown that the inner
product on k V only depends on that of V. This is a standard fact
following from the embedding
Vk
(2.8) V ,→ V ⊗k
2. HODGE THEORY 205
as the totally antisymmetric part, the fact thatVV ⊗k has a natural inner
product and the fact that this induces one on k V after normalization
(depending on the convention used in (2.8). These details are omitted.
Since ∗ is uniquely determined in this way, it necessarily depends
smoothly on the data, in particular the inner product. On an ori-
ented Riemannian manifold the induced inner product on Tp∗ M varies
smoothly with p (by assumption) so
∗ : kp M −→ n−k M , kp M = kp (Tp∗ M )
V V V V
(2.9) p
and that the null space in (2.23) is simply the direct sum of these spaces
over k. Indeed, from (2.23) the integration by parts in
Z Z
0 = hdu, (d + δ)ui ν = kdukL2 + hu, δ 2 ui ν = kduk2L2
2
is justified.
Thus we can consider d + δ as a Fredholm operator in three forms
d + δ :C −∞ (M, ∗ M ) −→ C −∞ (M, ∗ M ),
V V
d + δ :H 1 (M, ∗ M ) −→ H 1 (M, ∗ M ),
V V
(2.25)
d + δ :C ∞ (M, ∗ M ) −→ C ∞ (M, ∗ M )
V V
L2 (M, ∗ M ) = HHo∗
⊕ (d + δ)L2 (M, ∗ M ),
V V
(2.26)
C ∞ (M, ∗ M ) = HHo∗
⊕ (d + δ)C ∞ (M, ∗ M ).
V V
where the two de Rham spaces are those in (2.16), not yet shown to be
equal.
We proceed to show that the maps in (2.27) are isomorphisms. First
k
to show injectivity, suppose u ∈ HHo (M ) is mapped to zero in either
space. This means u = dv where v Vis either C ∞ or distributional, so
it suffices to suppose v ∈ C −∞ (M, k−1 M ). Since u is smooth the
integration by parts in the distributional pairing
Z Z
2
kukL2 = hu, dvi ν = hδu, vi ν = 0
M M
is justified, so u = 0 and the maps are injective.
To see surjectivity, use the Hodge decomposition (2.26). If u0 ∈
C −∞ (M, k M ) or C ∞ (M, k M ), we find
V V
(2.28) u0 = u0 + (d + δ)v
where correspondingly, v ∈ C −∞ (M, ∗ M ) or C ∞ (M, ∗ M ) and u0 ∈
V V
k
HHo (M ). If u0 is closed, du0 = 0, then dδv = 0 follows from applying
208 9. ELECTROMAGNETISM
∗
= 0, since δ 2 = 0. Thus δv ∈ HHo
d to (2.28) and hence (d + δ)δv V (M )
∞ ∗
and in particular, δv ∈ C (M, M ). Then the integration by parts
in Z Z
2
kδvkL2 = hδv, δvi ν = hv, (d + δ)δvi ν = 0
is justified, so δv = 0. Then (2.28) shows that any closed form, smooth
k
or distributional, is cohomologous in the same sense to u0 ∈ HHo (M ).
Thus the natural maps (2.27) are isomorphisms and the Theorem is
proved.
Thus, on a compact Riemannian manifold (whether orientable or
not), each de Rham class has a unique harmonic representative.
3. Coulomb potential
4. Dirac strings
Addenda to Chapter 9
CHAPTER 10
Monopoles
1. Gauge theory
2. Bogomolny equations
(1) Compact operators, spectral theorem
(2) Families of Fredholm operators(*)
(3) Non-compact self-adjoint operators, spectral theorem
(4) Spectral theory of the Laplacian on a compact manifold
(5) Pseudodifferential operators(*)
(6) Invertibility of the Laplacian on Euclidean space
(7) Lie groups( ), bundles and gauge invariance
(8) Bogomolny equations on R3
(9) Gauge fixing
(10) Charge and monopoles
(11) Monopole moduli spaces
* I will drop these if it looks as though time will become an issue.
, I will provide a brief and elementary discussion of manifolds and Lie
groups if that is found to be necessary.
3. Problems
Problem 1. Prove that u+ , defined by (15.10) is linear.
is compact.
(c) Show that d(x, K), for K compact, is continuous.
(d) Given > 0 show that there is a continuous function
g : R −→ [0, 1] such that g (t) = 1 for t ≤ /2 and
g (t) = 0 for t > 3/4.
(e) Show that f = g ◦d(·, K) satisfies the conditions for n = 1
if > 0 is small enough.
(3) Prove the general case by induction over n.
(a) In the general case, set K 0 = K ∩ U1{ and show that the
inductive hypothesis applies to K 0 and the Uj for j > 1; let
fj0 , j = 2, . . . , n be the functions supplied by the inductive
assumption and put f 0 = j≥2 fj0 .
P
with the series on the right always absolutely convergenct (i.e., this is
part of the requirement on µ). Define
∞
X
(3.2) |µ| (E) = sup |µ(Ei )|
i=1
for E S∈ M, with the supremum over all measurable decompositions
E = ∞ i=1 Ei with the Ei disjoint. Show that |µ| is a finite, positive
measure.
Hint 1. You must show that |µ| (E) = ∞
P S
i=1
S |µ| (A i ) if i Ai = E,
Ai ∈ M being disjoint. Observe that if Aj = l Ajl is a measurable
decomposition of ASj then together the Ajl give a decomposition of E.
Similarly, if E = j Ej is any such decomposition of E then Ajl =
Aj ∩ El gives such a decomposition of Aj .
Hint 2. See [6] p. 117!
Problem 13. (Hahn Decomposition)
With assumptions as in Problem 12:
212 10. MONOPOLES
will need the paralellogram law to verify the additivity of the RHS.
Note the polarization identity is a bit more transparent for real vector
spaces. There we have
(x, y) = 1/2(kx + yk2 − kx − yk2 )
both are easy to prove using kak2 = (a, a).
Problem 17. Show (Rudin does it) that if u : Rn → C has con-
tinuous partial derivatives then it is differentiable at each point in the
sense of (6.19).
Problem 18. Consider the function f (x) = hxi−1 = (1 + |x|2 )−1/2 .
Show that
∂f
= lj (x) · hxi−3
∂xj
with lj (x) a linear function. Conclude by induction that hxi−1 ∈
C0k (Rn ) for all k.
Problem 19. Show that exp(− |x|2 ) ∈ S(Rn ).
Problem 20. Prove (2.8), probably by induction over k.
Problem 21. Prove Lemma 2.4.
Hint. Show that a set U 3 0 in S(Rn ) is a neighbourhood of 0 if
and only if for some k and > 0 it contains a set of the form
X
n α β
ϕ ∈ S(R ) ; sup x D ϕ < .
|α|≤k,
|β|≤k
I suggest induction!
Problem 26. Prove the generalization of Proposition 3.10 that
u ∈ S 0 (Rn ), supp(w) ⊂ {0} implies there are constants cα , |α| ≤ m,
for some m, such that X
u= cα D α δ .
|α|≤m
Hint This is not so easy! I would be happy if you can show that
u ∈ M (Rn ), supp u ⊂ {0} implies u = cδ. To see this, you can show
that
ϕ ∈ S(Rn ), ϕ(0) = 0
⇒ ∃ϕj ∈ S(Rn ) , ϕj (x) = 0 in |x| ≤ j > 0(↓ 0) ,
sup |ϕj − ϕ| → 0 as j → ∞ .
To prove the general case you need something similar — that given m,
if ϕ ∈ S(Rn ) and Dα x ϕ(0) = 0 for |α| ≤ m then ∃ ϕj ∈ S(Rn ), ϕj = 0
in |x| ≤ j , j ↓ 0 such that ϕj → ϕ in the C m norm.
Problem 27. If m ∈ N, m0 > 0 show that u ∈ H m (Rn ) and
0 0
D u ∈ H m (Rn ) for all |α| ≤ m implies u ∈ H m+m (Rn ). Is the
α
converse true?
Problem 28. Show that every element u ∈ L2 (Rn ) can be written
as a sum
n
X
u = u0 + Dj uj , uj ∈ H 1 (Rn ) , j = 0, . . . , n .
j=1
Let M (Bn ) be the linear space of such measures. The space M (Sn−1 )
of measures on the sphere is defined similarly. Describe an injective
map
M (Sn−1 ) −→ M (Bn ).
Can you define another space so that this can be extended to a short
exact sequence?
Problem 43. Show that the Riemann integral defines a measure
Z
n
(3.3) C(B ) 3 f 7−→ f (x)dx.
Bn
u x>0
= c+ xz− x>0
and u x<0
= c− xz− x<0
3. PROBLEMS 219
viii) Use the Fourier transform on S 0 (Rn ) (and the fact that it is
an isomorphism on L2 (Rn )) to show that any tempered distri-
bution can be written in the form
u = (1 + |x|2 )N (1 + |D|2 )N w for some N and some w ∈ L2 (Rn ).
ix) Show that any tempered distribution can be written in the
form
u = (1+|x|2 )N (1+|D|2 )N +n+1 w̃ for some N and some w̃ ∈ H 2(n+1) (Rn ).
x) Conclude that any tempered distribution can be written in the
form
u = (1 + |x|2 )N (1 + |D|2 )M U for some N, M
and a bounded continuous function U
Problem 59. Distributions of compact support.
i) Recall the definition of the support of a distribution, defined
in terms of its complement
Rn \supp(u) = p ∈ Rn ; ∃ U ⊂ Rn , open, with p ∈ U such that u U = 0
is a non-increasing function of t.
Hint: Multiply the equation by u and integrate over a slab
[t1 , t2 ] × Rn .
x) Show that c in (3.15) is non-zero by arriving at a contradiction
from the assumption that it is zero. Namely, show that if c = 0
then u in viii) satisfies the conditions of ix) and also vanishes
in t < T for some T (depending on ψ). Conclude that u = 0 for
all ψ. Using properties of convolution show that this in turn
implies that F = 0 which is a contradiction.
xi) So, finally, we know that E = 1c F is a fundamental solution of
the heat operator which vanishes in t < 0. Explain why this
allows us to show that for any ψ ∈ Cc∞ (R × Rn ) there is a
solution of
n
X
(3.17) (∂t − ∂x2j )u = ψ, u = 0 in t < T for some T.
j=1
where the vj form a basis of Rn and using the dual basis wj (so wj · vi =
δij is 0 or 1 as i 6= j or i = j) set
( n
)
X
L◦ = w = 2π kj wj , kj ∈ Z .
j=1
converges in H.
v) Show that if ej is a complete orthonormal basis in a separable
Hilbert space then, for each u ∈ H,
∞
X
u= hu, ej iej .
j=1
such that
X
(3.23) |hu, ej i|2 ≤ ∀ u ∈ K.
j≥N
Hint: Problem 70 shows that (3.26) holds for each u ∈ L2 (Rn ); check
that (3.27) also holds for each function. Then use a covering argument
to prove that both these conditions must hold for a compact subset
of L2 (R) and hence for a precompact set. One method to prove the
converse is to show that if (3.26) and (3.27) hold then B is bounded
and to use this to extract a weakly convergent sequence from any given
sequence in B. Next show that (3.26) is equivalent to (3.27) for the
3. PROBLEMS 233
set F(B), the image of B under the Fourier transform. Show, possi-
bly using Problem 71, that if χR is cut-off to a ball of radius R then
χR G(χR ûn ) converges strongly if un converges weakly. Deduce from
this that the weakly convergent subsequence in fact converges strongly
so B̄ is sequently compact, and hence is compact.
Problem 73. Consider the space Cc (Rn ) of all continuous functions
on Rn with compact support. Thus each element vanishes in |x| > R
for some R, depending on the function. We want to give this a toplogy
in terms of which is complete. We will use the inductive limit topology.
Thus the whole space can be written as a countable union
(3.28) [
Cc (Rn ) = {u : Rn ; u is continuous and u(x) = 0 for |x| > R}.
n
Each of the space on the right is a Banach space for the supremum
norm.
(1) Show that the supreumum norm is not complete on the whole
of this space.
(2) Define a subset U ⊂ Cc (Rn ) to be open if its intersection with
each of the subspaces on the right in (3.28) is open w.r.t. the
supremum norm.
(3) Show that this definition does yield a topology.
(4) Show that any sequence {fn } which is ‘Cauchy’ in the sense
that for any open neighbourhood U of 0 there exists N such
that fn − fm ∈ U for all n, m ≥ N, is convergent (in the
corresponding sense that there exists f in the space such that
f − fn ∈ U eventually).
(5) If you are determined, discuss the corresponding issue for nets.
Problem 74. Show that the continuity of a linear functional u :
Cc∞ (Rn ) −→ C with respect to the inductive limit topology defined in
(1.17) means precisely that for each n ∈ N there exists k = k(n) and
C = Cn such that
(3.29) |u(ϕ)| ≤ CkϕkC k , ∀ ϕ ∈ C˙∞ (B(n)).
The point of course is that the ‘order’ k and the constnat C can both
increase as n, measuring the size of the support, increases.
Problem 75. [Restriction from Sobolev spaces] The Sobolev em-
bedding theorem shows that a function in H m (Rn ), for m > n/2 is
continuous – and hence can be restricted to a subspace of Rn . In fact
this works more generally. Show that there is a well defined restriction
234 10. MONOPOLES
map
1 1
(3.30) H m (Rn ) −→ H m− 2 (Rn ) if m >
2
with the following properties:
(1) On S(Rn ) it is given by u 7−→ u(0, x0 ), x0 ∈ Rn−1 .
(2) It is continuous and linear.
Hint: Use the usual method of finding a weak version of the map on
smooth Schwartz functions; namely show that in terms of the Fourier
transforms on Rn and Rn−1
Z
0 −1
(3.31) \
u(0, ·)(ξ ) = (2π) û(ξ1 , ξ 0 )dξ1 , ∀ ξ 0 ∈ Rn−1 .
R
Use Cauchy’s inequality to show that this is continuous as a map on
Sobolev spaces as indicated and then the density of S(Rn ) in H m (Rn )
to conclude that the map is well-defined and unique.
Problem 76. [Restriction by WF] From class we know that the
product of two distributions, one with compact support, is defined
provided they have no ‘opposite’ directions in their wavefront set:
(3.32) / WF(v) then uv ∈ Cc−∞ (Rn ).
(x, ω) ∈ WF(u) =⇒ (x, −ω) ∈
Show that this product has the property that f (uv) = (f u)v = u(f v)
if f ∈ C ∞ (Rn ). Use this to define a restriction map to x1 = 0 for
distributions of compact support satisfying ((0, x0 ), (ω1 , 0)) ∈
/ WF(u)
as the product
(3.33) u0 = uδ(x1 ).
[Show that u0 (f ), f ∈ C ∞ (Rn ) only depends on f (0, ·) ∈ C ∞ (Rn−1 ).
Problem 77. [Stone’s theorem] For a bounded self-adjoint opera-
tor A show that the spectral measure can be obtained from the resolvent
in the sense that for φ, ψ ∈ H
1
(3.34) lim h[(A − t − i)−1 − (A + t + i)−1 ]φ, ψi −→ µφ,ψ
↓0 2πi
X l! ∂lu
wx(l) (t) = l1 l2 li
(tx)xl11 xl22 · · · xlii
l1 !l2 ! · · · li ! ∂x1 ∂x2 · · · ∂xi
l1 +l2 +...+l =l
i
238 10. MONOPOLES
m0 m0
and since hξi ξ α u
b = hξi D d α u is in L2 (Rn ) (note that u ∈ H m (Rn )
0
follows from Dα u ∈ H m (Rn ), |α| ≤ m). The converse is also true since
Cα in the formula above are strictly positive.
Solution 4.7. Take v ∈ L2 (Rn ), and define subsets of Rn by
E0 = {x : |x| ≤ 1},
Ei = {x : |x| ≥ 1, |xi | = max |xj |}.
j
uj ∈ L2 (Rn ). Hence
where hξib
n
X
u = u0 + Dj uj
j=1
where uj ∈ H 1 (Rn ).
Solution 4.8. Since
Z ∞ Z ∞
0
Dx H(ϕ) = H(−Dx ϕ) = i H(x)ϕ (x) dx = i ϕ0 (x) dx = i(0−ϕ(0)) = −iδ(ϕ),
−∞ 0
243