Lecture Notes
Lecture Notes
RICHARD B. MELROSE
Contents
Introduction 1
1. Continuous functions 2
2. Measures and σalgebras 10
3. Measureability of functions 16
4. Integration 19
5. Hilbert space 30
6. Test functions 34
7. Tempered distributions 42
8. Convolution and density 47
9. Fourier inversion 58
10. Sobolev embedding 63
11. Differential operators. 67
12. Cone support and wavefront set 83
13. Homogeneous distributions 96
14. Wave equation 97
15. Operators and kernels 98
16. Spectral theorem 99
17. Problems 103
18. Solutions to (some of) the problems 130
References 136
Introduction
These notes are for the course the graduate analysis course (18.155)
at MIT in Fall 2004. They are based on earlier notes for similar courses
in 1997, 2001 and 2002. In giving the lectures I may cut some corners!
I wish to particularly thank Austin Frakt for many comments on,
and corrections to, an earlier version of these notes. Others who made
helpful comments or noted errors include Philip Dorrell, ....
1
2 RICHARD B. MELROSE
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 ([5] 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 con
tinuous 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 ntuples of real
numbers
x = (x1 , . . . , xn ) ∈ Rn , xj ∈ R , j = 1, . . . , n .
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
n
�
2 2 1/2
|x| = (x1 + · · · + xn ) = ( x2j )1/2 ,
j=1
Proof of Lemma 1.7. We have to prove (1.14). Suppose first that the
Ai are open, then so is A = i Ai . If f ∈ C(X) and supp(f ) � A then
supp(f ) is covered by a finite union of the Ai s. Applying Lemma 1.8 we
� fi ’s, all but a finite number identically zero, so supp(fi ) � Ai
can find
and i fi = � 1 in a neighborhood of supp(f ).
Since f = i fi f we conclude that
� �
u(f ) = u(fi f ) =⇒ µ∗ (A) ≤ µ∗ (Ai )
i i
∞
nIf {Ei }i=1 is a sequence of disjoint µ∗ measurable sets, set Fn =
∞
i=1 Ei and F = i=1 Ei . Then for any A,
µ∗ (A ∩ Fn ) = µ∗ (A ∩ Fn ∩ En ) + µ∗ (A ∩ Fn ∩ EnC )
= µ∗ (A ∩ En ) + µ∗ (A ∩ Fn−1 ) .
Iterating this shows that
n
�
µ∗ (A ∩ Fn ) = µ∗ (A ∩ Ej ) .
j=1
µ∗ (A) = µ∗ (A ∩ Fn ) + µ∗ (A ∩ FnC )
�n
≥ 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
�j ,
j=1 j=1
j−1
�j−1 �C
�j = Aj \
A Ai = Aj ∩ Ai
i=1 i=1
∗
is µ measurable since the A
�j are disjoint. �
A measure (sometimes called a positive measure) is an extended func
tion defined on the elements of a σalgebra M:
µ : M → [0, ∞]
such that
(2.3) µ(∅) = 0 and
�∞ � ∞
�
µ Ai = µ(Ai )
(2.4) i=1 i=1
if {Ai }∞
i=1 � j.
⊂ M and Ai ∩ Aj = φ i =
12 RICHARD B. MELROSE
∗ ∗
µ (F ) = µ (Ej ) if F = Ej
j j=1
2Why?
LECTURE NOTES FOR 18.155, FALL 2004 13
3Meaning in a neighborhood of K.
LECTURE NOTES FOR 18.155, FALL 2004 15
3. Measureability of functions
Suppose that M is a σalgebra on a set X 4 and N is a σalgebra on
another set Y. A map f : X → Y is said to be measurable with respect
to these given σalgebras 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 � ; N � ⊃ G, N � a σalgebra}
−1
f Ej = f −1 (Ej )
j=1 j=1
(3.3) �∞ �
� ∞
�
−1
f Ej = f −1 (Ej )
j=1 j=1
−1 −1
f (φ) = φ , f (Y ) = X .
Putting these things together one sees that if M is any σalgebra on
X then
f∗ (M) = E ⊂ Y ; f −1 (E) ∈ M
� �
(3.4)
is always a σalgebra on Y.
In particular if f −1 (A) ∈ M for all A ∈ G ⊂ N then f∗ (M) is a σ
algebra containing G, hence containing N by the generating condition.
Thus f −1 (E) ∈ M for all E ∈ N so f is measurable. �
Proposition 3.2. Any continuous map f : X → Y between metric
spaces is measurable with respect to the Borel σalgebras on X and Y.
4. Integration
The (µ)integral of a nonnegative simple function is by definition
� �
(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.
� 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 se
quence of nonnegative measurable (extended) functions, then f (x) =
limn→∞ fn (x) is measurable and
� �
(4.5) f dµ = lim fn dµ
E n→∞ E
for any measurable set E ⊂ X.
Proof. To see that f is measurable, observe that
Since the sets (a, ∞] generate the Borel σalgebra 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, [5] page 21. Here I
paraphrase it. We can easily see from (4.1) that
� � �
α = sup fn dµ = lim fn dµ ≤ f dµ.
E n→∞ E E
Given a simple measurable function g with 0 ≤ g ≤ f and 0 < c < 1
consider the sets En = {x ∈ E; fn (x)
≥ cg(x)}. These are measurable
and increase with n. Moreover E = n En . It follows that
� � � �
(4.7) fn dµ ≥ fn dµ ≥ c gdµ = ai µ(En ∩ Fi )
E En En i
LECTURE NOTES FOR 18.155, FALL 2004 21
�
in terms of the natural presentation of g = i ai χFi . Now, the fact
that the En are measurable and increase to E shows that
µ(En ∩ Fi ) → µ(E ∩ Fi )
�
as n → ∞. Thus
� the right side of (4.7) tends to c E gdµ as n → ∞.
Hence α ≥ c E gdµ for all 0 < c < 1. Taking the supremum over c and
then over all such g shows that
� � �
α = lim fn dµ ≥ sup gdµ = f dµ.
n→∞ E E E
They must therefore be equal. �
Now for instance the additivity in (4.1) for f ≥ 0 and g ≥ 0 any
measurable functions follows from Proposition 3.3. Thus if f ≥ 0 is
measurable
� and fn is an �approximating sequence as in the Proposition
then E f dµ = limn→∞ E fn dµ. So if f and g are two nonnegative
measurable functions then fn (x) + gn (x) ↑ f + g(x) which shows not
only that f + g is measurable by also that
� � �
(f + g)dµ = f dµ + gdµ.
E E E
As with the definition of u+ long ago, this allows us to extend the
definition of the integral to any integrable function.
Definition 4.4. A measurable extended function f : X −→ [−∞, ∞]
is said to be integrable on E if its positive and negative parts both have
finite integrals over E, and then
� � �
f dµ = f+ dµ − f− dµ.
E E E
�
We further extend the integral to complexvalued functions, just say
ing that
f :X→C
is integrable if its real and imaginary parts are both integrable. Then,
by definition, � � �
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 �� � �
� �
� f dµ� ≤ |f | dµ .
� �
E E
�
This is obvious if E f dµ = 0, and if not then
�
f dµ = Reiθ R > 0 , θ ⊂ [0, 2π) .
E
LECTURE NOTES FOR 18.155, FALL 2004 23
Then
�� � �
� f dµ� = e−iθ
� �
� � f dµ
E E
�
= e−iθ f dµ
�E
= Re(e−iθ f ) dµ
�E
�Re(e−iθ f )� dµ
� �
≤
�E �
� −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
� �
f dµ = lim fn dµ .
n→∞
Proof. First we can make the sequence fn (x) converge by changing all
the fn (x)’s to zero on a set of measure zero outside which they converge.
This does not change the conclusions. Moreover, it suffices to suppose
that the fn are realvalued. Then consider
hk = g − fk ≥ 0 .
Now, lim inf k→∞ hk = g − f by the convergence of fn ; in particular f
is integrable. By monotone convergence and Fatou’s lemma
� � �
(g − f )dµ = lim inf hk dµ ≤ lim inf (g − fk ) dµ
k→∞ k→∞
� �
= g dµ − lim sup fk dµ .
k→∞
Similarly, if Hk = g + fk then
� � � �
(g + f )dµ = lim inf Hk dµ ≤ g dµ + lim inf fk dµ.
k→∞ k→∞
It follows that
� � �
lim sup fk dµ ≤ f dµ ≤ lim inf fk dµ.
k→∞ k→∞
Thus in fact � �
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.,
�
|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) �= 0} = 0 .
p
We denote by L (X, µ) the space of such equivalence classes. It is a
linear space and the function
�� �1/p
p
(4.10) �f �p = |f | dµ
X
1
Since q(p − 1) = p and 1 − q
= 1/p this is just (4.14). �
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.
26 RICHARD B. MELROSE
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
RadonNikodym 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) �µ�1 = 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):
� �
(4.18) · dµ : C0 (X) � f �−→ f · dµ .
X
Proof. We have done half of this already. Let me remind you of the
steps.
We started with u ∈ (C0 (X))� and showed that u = u+ − u− where
u± are positive continuous linear functionals; this is Lemma 1.5. Then
we showed that u ≥ 0 defines a finite positive Radon measure µ. Here
µ is defined by (1.11) on open sets and µ(E) = µ∗ (E) is given by (1.12)
28 RICHARD B. MELROSE
5. 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 special 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
(5.1) (a1 u1 + a2 u2 , v) = a1 (u1 , v) + a2 (u2 , v),
antilinearly in the second
(5.2) (u, a1 v1 + a2 v2 ) = a1 (u, v1 ) + a2 (u, v2 )
and the conjugacy condition
(5.3) (u, v) = (v, u) .
Notice that (5.2) follows from (5.1) and (5.3). If we assume in addition
the positivity condition8
(5.4) (u, u) ≥ 0 , (u, u) = 0 ⇒ u = 0 ,
then
(5.5) �u� = (u, u)1/2
is a norm on V , as we shall see.
Suppose that u, v ∈ V have �u� = �v� = 1. Then (u, v) = eiθ |(u, v)|
for some θ ∈ R. By choice of θ, e−iθ (u, v) = |(u, v)| is real, so expanding
out using linearity for s ∈ R,
0 ≤ (e−iθ u − sv , e−iθ u − sv)
= �u�2 − 2s Re e−iθ (u, v) + s2 �v�2 = 1 − 2s|(u, v)| + s2 .
The minimum of this occurs when s = |(u, v)| and this is negative
unless |(u, v)| ≤ 1. Using linearity, and checking the trivial cases u =
or v = 0 shows that
(5.6) |(u, v)| ≤ �u� �v�, ∀ u, v ∈ V .
This is called Schwarz’9 inequality.
8Notice that (u, u) is real by (5.3).
9No ‘t’ in this Schwarz.
LECTURE NOTES FOR 18.155, FALL 2004 31
�
The fundamental fact about a Hilbert space is that each element
v ∈ H defines a continuous linear functional by
H � u �−→ (u, v) ∈ C
and conversely every continuous linear functional arises this way. This
is also called the Riesz representation theorem.
Proposition 5.3. If L : H → C is a continuous linear functional on
a Hilbert space then this is a unique element v ∈ H such that
(5.9) Lu = (u, v) ∀ u ∈ H ,
Proof. Consider the linear space
M = {u ∈ H ; Lu = 0}
the null space of L, a continuous linear functional on H. By the as
sumed continuity, M is closed. We can suppose that L is not identically
zero (since then v = 0 in (5.9)). Thus there exists w ∈
/ M . Consider
w + M = {v ∈ H ; v = w + u , u ∈ M } .
This is a closed convex subset of H. Applying Lemma 5.2 it has a
unique smallest element, v ∈ w + M . Since v minimizes the norm on
w + M,
�v + su�2 = �v�2 + 2 Re(su, v) + �s�2 �u�2
is stationary at s = 0. Thus Re(u, v) = 0 ∀ u ∈ M , and the same
argument with s replaced by is shows that (v, u) = 0 ∀ u ∈ M .
Now v ∈ w + M , so Lv = Lw = � 0. Consider the element w� =
�
w/Lw ∈ H. Since Lw = 1, for any u ∈ H
L(u − (Lu)w� ) = Lu − Lu = 0 .
It follows that u − (Lu)w� ∈ M so if w�� = w� /�w� �2
(w� , w� )
(u, w�� ) = ((Lu)w� , w�� ) = Lu = Lu .
�w� �2
The uniqueness of v follows from the positivity of the norm. �
Corollary 5.4. For any positive measure µ, any continuous linear
functional
L : L2 (X, µ) → C
is of the form �
Lf = f g dµ , g ∈ L2 (X, µ) .
X
LECTURE NOTES FOR 18.155, FALL 2004 33
6. 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
� � �
2 2
L (Bn ) = u : Bn → C; Borel measurable with |u| dx < ∞ .
Definition 6.4. The space C0k (Rn ) ⊂ C01 (Rn ) k ≥ 1 is defined induc
tively by requiring that
∂u
∈ C0k−1 (Rn ) , j = 1, . . . , n .
∂xj
The norm on C0k (Rn ) is taken to be
n
� ∂u
(6.8) �u�C k = �u�C k−1 + � � k−1 .
j=1
∂xj C
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 6.3.
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
(6.9) �x� = (1 + |x|2 )1/2
a function which is the size of |x| for |x| large, but has the virtue of
being smooth10
Definition 6.5. For any k, l ∈ N = {1, 2, · · · } set
�x�−l C0k (Rn ) = u ∈ C0k (Rn ) ; u = �x�−l v , v ∈ C0k (Rn ) ,
� �
is a metric on any normed space, since then we may sum over k. Thus
we consider
�u − v� �v − w�
+
1 + �u − v� 1 + �v − w�
�u − v�(1 + �v − w�) + �v − w�(1 + �u − v�)
= .
(1 + �u − v�)(1 + �v − w�)
(1 + �u − v�)(1 + �v − w�)�u − w�
≤ (�u − v�(1 + �v − w�) + �v − w�(1 + �u − v�))(1 + �u − w�).
Thus, d is a metric.
Suppose un is a Cauchy sequence. Thus, d(un , um ) → 0 as n, m →
∞. In particular, given
�un − um �(k)
< � ∀ n, m > N.
1 + �un − um �(k)
so the sequence is Cauchy in �x�−k C0k (Rn ) for each k. From the com
pleteness of these spaces it follows that un → u in �x�−k C0k (Rn )j for
each k. Given � > 0 choose k so large that 2−k < �/2. Then ∃ N s.t.
n>N
⇒ �u − un �(j) < �/2 n > N, j ≤ k.
40 RICHARD B. MELROSE
Hence
� �u − un �(j)
d(un , u) = 2−j
j≤k
1 + �u − un �(j)
� �u − un �(j)
+ 2−j
j>k
1 + �u − un �(j)
Cc∞ (Rn ) = C˙c∞ (B(n)), C˙c∞ (B(n)) = {u ∈ S(Rn ); u = 0 in |x| > n}.
n∈N
Consider
(6.16)
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 –
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))
LECTURE NOTES FOR 18.155, FALL 2004 41
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 7.3 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.
42 RICHARD B. MELROSE
7. Tempered distributions
A good first reference for distributions is [2], [4] gives a more exhaus
tive 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 � (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
(7.1) 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:
(7.2) 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 ‘multiindex’ and if ϕ is
at least k times continuously differentiable, we set12
1 ∂ α1 ∂ αn
(7.3) Dα ϕ = · · · ϕ whenever |α| = α1 +α2 +· · · +αn ≤ k.
i|α| ∂x1 ∂xn
Now we have defined the spaces.
C0k (Rn ) = ϕ : Rn → C ; Dα ϕ ∈ C00 (Rn ) ∀ |α| ≤ k .
� �
(7.4)
Notice the convention is that Dα ϕ is asserted to exist if it is required
to be continuous! Using �x� = (1 + |x|2 ) we defined
�x�−k C0k (Rn ) = ϕ : Rn → C ; �x�k ϕ ∈ C0k (Rn ) ,
� �
(7.5)
and then our space of test functions is
�
S(Rn ) = �x�−k C0k (Rn ) .
k
12Periodically
there is the possibility of confusion between the two meanings of
|α| but it seldom arises.
LECTURE NOTES FOR 18.155, FALL 2004 43
Thus,
(7.6) ϕ ∈ S(Rn ) ⇔ Dα (�x�k ϕ) ∈ C00 (Rn ) ∀ |α| ≤ k and all k .
Lemma 7.1. The condition ϕ ∈ S(Rn ) can be written
�x�k Dα ϕ ∈ C00 (Rn ) ∀ |α| ≤ k , ∀ k .
Proof. We first check that
ϕ ∈ C00 (Rn ) , Dj (�x�ϕ) ∈ C00 (Rn ) , j = 1, · · · , n
⇔ ϕ ∈ C00 (Rn ) , �x�Dj ϕ ∈ C00 (Rn ) , j = 1, · · · , n .
Since
Dj �x�ϕ = �x�Dj ϕ + (Dj �x�)ϕ
and Dj �x� = 1
x �x�−1 is a bounded continuous function,
i j
this is clear.
Then consider the same thing for a larger k:
(7.7) Dα �x�p ϕ ∈ C00 (Rn ) ∀ |α| = p , 0 ≤ p ≤ k
⇔ �x�p Dα ϕ ∈ C00 (Rn ) ∀ |α| = p , 0 ≤ p ≤ k .
�
I leave you to check this as Problem 7.1.
Corollary 7.2. For any k ∈ N the norms
�
��x�k ϕ�C k and �xα Dxβ ϕ�∞
|α|≤k,
|β|≤k
are equivalent.
Proof. Any reasonable proof of (7.2) shows that the norms
�
��x�k ϕ�C k and ��x�k Dβ ϕ�∞
|β|≤k
Proof. This is just the equivalence of the norms, since we showed that
u ∈ S � (Rn ) if and only if
|u(ϕ)| ≤ C��x�k ϕ�C k
for some k.
�
Lemma 7.4. A linear map
T : S(Rn ) → S(Rn )
is continuous if and only if for each k there exist C and j such that if
|α| ≤ k and |β| ≤ k
� �
� � � � � � �
(7.8) sup �xα Dβ T ϕ� ≤ C sup �xα Dβ ϕ� ∀ ϕ ∈ S(Rn ).
Rn
|α� |≤j, |β � |≤j
�
≤ C�
� �
sup �xα Dβ ψ � .
Rn
|α|≤k+1, |β|≤k
LECTURE NOTES FOR 18.155, FALL 2004 45
To see that this is small for x� small, we split the integral into two
pieces. Since ψ is very small near infinity, given � > 0 we can choose
R so large that
�
(8.5) �v�∞ · |ψ(y)| dy ≤ �/4 .
|y]|≥R
This reverses the role of v and ψ and shows that if both v and ψ are in
S(Rn ) then v ∗ ψ = ψ ∗ v.
Using this formula on (8.4) we find
(8.11) �
�
v ∗ ψ(x + x ) − v ∗ ψ(x) = v(y)(ψ(x + x� − y) − ψ(x − y)) dy
n
� � n
�
�
= xj v(y)ψ̃j (x − y, x ) dy = xj (v ∗ ψj (·; x� )(x) .
j=1 Rn j=1
LECTURE NOTES FOR 18.155, FALL 2004 49
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 8.3. C0k (Rn ) is dense in C0p (Rn ) for any k ≥ p.
Proposition 8.4. S(Rn ) is dense in C0k (Rn ) for any k ≥ 0.
50 RICHARD B. MELROSE
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 8.5. The map from finite Radon measures
(8.19) Mfin (Rn ) � µ �−→ uµ ∈ S � (Rn )
is injective.
Now, we want the same result for L2 (Rn ) (and maybe for Lp (Rn ),
1 ≤ p < ∞). I leave the measuretheoretic part of the argument to
you.
Proposition 8.6. Elements of L2 (Rn ) are “continuous in the mean”
i.e.,
�
(8.20) 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 � ) · |u(x − y � ) − u(x)| ϕt (y)ϕt (y � ) .
Thus we now know that
L2 (Rn ) �→ S � (Rn ) is injective.
This means that all our usual spaces of functions ‘sit inside’ S � (Rn ).
Finally we can use convolution with ϕt to show the existence of
smooth partitions of unity. If K � U ⊂ Rn is a compact set in an
open set then we have shown the existence of ξ ∈ Cc0 (Rn ), with ξ = 1
in some neighborhood of K and ξ = 1 in some neighborhood of K and
supp(ξ) � U .
Then consider ξ ∗ ϕt for t small. In fact
supp(ξ ∗ ϕt ) ⊂ {p ∈ Rn ; dist(p, supp ξ) ≤ 2t}
and similarly, 0 ≤ ξ ∗ ϕt ≤ 1 and
ξ ∗ ϕt = 1 at p if ξ = 1 on B(p, 2t) .
Using this we get:
52 RICHARD B. MELROSE
Clearly the right side of (8.25) is contained in the left. To see the
converse, suppose first that
(8.26) ϕ ∈ S(Rn ) , ϕ = 0 in |x| < 1 .
Then define
�
0 |x| < 1
ψ= 2
ϕ/ |x| |x| ≥ 1 .
54 RICHARD B. MELROSE
this shows that ϕ of the form (8.26) is in the right side of (8.25). In
general suppose ϕ ∈ S(Rn ). Then
� t
d
ϕ(x) − ϕ(0) = ϕ(tx) dt
0 dt
(8.27) n � t
� ∂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
(8.27) becomes, if ϕ(0) = 0,
ϕ = µϕ + (1 − µ)ϕ
n � t
� ∂ϕ
= xj ψj + (1 − µ)ϕ , ψj = µ (tx) dt ∈ S(Rn ) .
j=1 0 ∂xj
→ 0 as ξ � → ξ .
In fact
Proposition 8.11. Fourier transformation, (8.28), defines a continu
ous linear map
(8.29) F : S(Rn ) → S(Rn ) , Fϕ = ϕˆ .
Proof. Differentiating under the integral15 sign shows that
�
∂ξj ϕ(ξ)
ˆ = −i e−ix·ξ xj ϕ(x) dx .
Since the integral on the right is absolutely convergent that shows that
(remember the i’s)
(8.30) xj ϕ , ∀ ϕ ∈ S(Rn ) .
Dξj ϕˆ = −�
Similarly, if we multiply by ξj and observe that ξj e−ix·ξ = i ∂x∂ j e−ix·ξ
then integration by parts shows
�
∂ −ix·ξ
(8.31) ξj ϕ̂ = i ( e )ϕ(x) dx
∂xj
�
∂ϕ
= −i e−ix·ξ dx
∂xj
n
D j ϕ = ξj ϕ̂ , ∀ ϕ ∈ S(R ) .
�
15See [5]
56 RICHARD B. MELROSE
n
ˆ j ) , ψ(x) = e−x2 /2 ,
�
ϕ(ξ)
ˆ = ψ(ξ
j=1
16Really by Fubini’s theorem, but here one can use Riemann integrals.
LECTURE NOTES FOR 18.155, FALL 2004 57
Since this is true for ϕ = exp(− |x|2 /2). The identity allows us to
invert the Fourier transform.
58 RICHARD B. MELROSE
9. Fourier inversion
It is shown above that the Fourier transform satisfies the identity
�
−n
(9.1) ϕ(0) = (2π) ϕ̂(ξ) dξ ∀ ϕ ∈ S(Rn ) .
Rn
= eiy·ξ ϕ(ξ)
ˆ .
�
60 RICHARD B. MELROSE
Thus �ξ�m û = α m α m
�
|α|≤−m ξ �ξ� v̂α . Since all the factors ξ �ξ� are
2 n m 2 n
bounded, each term here is in L (R ), so �ξ� û ∈ L (R ) which is the
definition, u ∈ �ξ�−m L2 (Rn ).
Conversely, suppose u ∈ H m (Rn ), i.e., �ξ�m uˆ ∈ L2 (Rn ). The func
tion ⎛ ⎞
�
⎝ |ξ α |⎠ · �ξ�m ∈ L2 (Rn ) (m < 0)
|α|≤−m
The Schwartz space S(Rn ) �→ H m (Rn ) is dense for each m and the
pairing
(9.18) H m (Rn ) × H −m (Rn ) � (u, u� ) �−→
�
�
((u, u )) = uˆ� (ξ)û� (·ξ) dξ ∈ C
Rn
Proof. The Hilbert space property follows essentially directly from the
definition (9.14) since �ξ�−m L2 (Rn ) is a Hilbert space with the norm
(9.17). Similarly the density of S in H m (Rn ) follows, since S(Rn ) dense
in L2 (Rn ) (Problem L11.P3) implies �ξ�−m S(Rn ) = S(Rn ) is dense in
�ξ�−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 (9.18) makes sense, since �ξ�−m û(ξ),
�ξ�m uˆ� (ξ) ∈ L2 (Rn ) implies
û(ξ))û� (−ξ) ∈ L1 (Rn ) .
Furthermore, by the selfduality of L2 (Rn ) each continuous linear func
tional
U : H m (Rn ) → C , U (u) ≤ C�u�H m
can be written uniquely in the form
U (u) = ((u, u� )) for some u� ∈ H −m (Rn ) .
�
Notice that if u, u� ∈ S(Rn ) then
�
�
((u, u )) = u(x)u� (x) dx .
Rn
This is always how we “pair” functions — it is the natural pairing on
L2 (Rn ). Thus in (9.18) what we have shown is that this pairing on test
function �
S(Rn ) × S(Rn ) � (u, u� ) �−→ ((u, u� )) = u(x)u� (x) dx
Rn
extends by continuity to H m (Rn ) × H −m (Rn ) (for each fixed m) when
it identifies H −m (Rn ) as the dual of H m (Rn ). This was our ‘picture’
at the beginning.
For m > 0 the spaces H m (Rn ) represents elements of L2 (Rn ) that
have “m” derivatives in L2 (Rn ). For m < 0 the elements are ?? of “up
to −m” derivatives of L2 functions. For integers this is precisely ??.
LECTURE NOTES FOR 18.155, FALL 2004 63
�� �1/2 �� �1/2
≤ �ξ�2m |uˆ(ξ)|2 dξ · �ξ�−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
(10.2) sup |u(x)| = �u�L∞ ≤ (2π)−n �u�H m , m > n/2 .
Rn
Proof. This follows directly from (10.5) since the left side is contained
in �
�x�−k C0k−n (Rn ) ⊂ S(Rn ).
k
�
Theorem 10.5 (Schwartz representation). Any tempered distribution
can be written in the form of a finite sum
�
(10.8) u= xα Dxβ uαβ , uαβ ∈ C00 (Rn ).
|α|≤m
|β|≤m
or in the form
�
(10.9) u= Dxβ (xα vαβ ), vαβ ∈ C00 (Rn ).
|α|≤m
|β|≤m
To get (10.9) we ‘commute’ the factor �x�k to the inside; since I have
not done such an argument carefully so far, let me do it as a lemma.
17This is probably the most useful form of the representation theorem!
66 RICHARD B. MELROSE
Lemma 10.6. For any γ ∈ Nn0 there are polynomials pα,γ (x) of degrees
at most |γ − α| such that
�
�x�k Dγ v = Dγ−α pα,γ �x�k−2|γ−α| v .
� �
α≤γ
Any polynomial on Rn
�
p(ξ) = pα ξ α , pα ∈ C
|α|≤m
Proof. Since (x+iy)−1 is smooth and bounded away from the origin the
local integrability follows from the estimate, using polar coordinates,
� � 2π � 1
dx dy r dr dθ
(11.11) = = 2π .
|(x,y)|≤1 |x + iy| 0 0 r
Differentiating directly in the region where it is smooth,
∂x (x + iy)−1 = −(x + iy)−2 , ∂y (x + iy)−1 = −i(x ∈ iy)−2
� 0.20
so indeed, ∂E = 0 in (x, y) =
The derivative is really defined by
(11.12) (∂E)(ϕ) = E(−∂ϕ)
�
1
= lim − (x + iy)−1 ∂ϕ dx dy .
�↓0 2π |x|≥�
|y|≥�
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
� �
−1
− (x + iy) ∂x ϕ dx dy = (∂x (x + iy)−1 )ϕ dx dy
|x|≥� |x|≥�
|y|≥� |y|≥�
� �
−1
+ (x + iy) ϕ(x, y) dy − (x + iy)−1 ϕ(x, y) dy .
|y|≤� |y|≤�
x=� x=−�
(11.13) 2π∂E(ϕ) =
�
lim [(� + iy)−1 ϕ(�, y) − (−� + iy)−1 ϕ(−�, y)] dy
�↓0 |y|≤�
�
+ i lim [(x + i�)−1 ϕ(x, �) − (x − i�)−1 ϕ(x, �)] dx ,
�↓0 |x|≤�
Thus we get the same result in (11.13) by replacing ϕ(x, y) by ϕ(0, 0).
Then 2π∂E(ϕ) = cϕ(0),
� �
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
�
u ∗ v(x) = u(x − y)v(y) dy = v ∗ u(x) .
so it follows that
(11.15) |u ∗ ϕ(x)| = |�u, ϕ(x − ·)�| ≤ C(1 + |x|2 )k/2 .
LECTURE NOTES FOR 18.155, FALL 2004 71
where ϕ(z)
ˇ = ϕ(−z). In fact using Problem 35,
(11.19) u ∗ v(ϕ) = ((u ∗ v) ∗ ϕ)(0)
ˇ = (u ∗ (v ∗ ϕ))(0)
ˇ .
Here, v, ϕ are both smooth, but notice
Lemma 11.7. If v ∈ S � (Rn ) has compact support and ϕ ∈ S(Rn ) then
v ∗ ϕ ∈ S(Rn ).
Proof. Since v ∈ S � (Rn ) has compact support there exists χ ∈ Cc∞ (Rn )
such that χv = v. Then
v ∗ ϕ(x) = (χv) ∗ ϕ(x) = �χv(y), ϕ(x − y)�y
= �u(y), χ(y)ϕ(x − y)�y .
Thus, for some k,
|v ∗ ϕ(x)| ≤ C�χ(y)ϕ(x − y)�(k)
where � �(k) is one of our norms on S(Rn ). Since χ is supported in
some large ball,
�χ(y)ϕ(x − y)�(k)
� �
≤ sup ��y�k Dα y (χ(y)ϕ(x − y))�
|α|≤k
≤ CN sup (1 + |x − y |2 )−N/2
|y|≤R
≤ 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 11.6, so
v ∗ ϕ ∈ S(Rn ) .
�
In fact we get a little more, since we see that for each k there exists
�
k and C (depending on k and v) such that
�v ∗ ϕ�(k) ≤ C�ϕ�(k� ) .
LECTURE NOTES FOR 18.155, FALL 2004 73
�
Proof. We must show that x ∈ / sing supp(u) ⇒ x ∈/ sing supp(P (D)u).
/ sing supp(u) we can find ϕ ∈ Cc∞ (Rn ), ϕ ≡ 1 near x, such
Now, if x ∈
that ϕu ∈ Cc∞ (Rn ). Then
P (D)u = P (D)(ϕu + (1 − ϕ)u)
= P (D)(ϕu) + P (D)((1 − ϕ)u) .
The first term is C ∞ and x ∈
/ supp(P (D)((1−ϕ)u)), so x ∈
/ sing supp(P (D)u).
�
It remains to show the converse of (11.26) where P (D) is assumed to
be hypoelliptic. Take F , a parametrix for P (D) with sing supp u ⊂ {0}
and assume, or rather arrange, that F have compact support. In fact
if x ∈
/ sing supp(P (D)u) we can arrange that
(supp(F ) + x) ∩ sing supp(P (D)u) = φ .
Now P (D)F = δψ with ψ ∈ Cc∞ (Rn ) so
u = δ ∗ u = (P (D)F ) ∗ u − ψ ∗ u.
Since ψ ∗ u ∈ C ∞ it suffices to show that x¯ ∈
/ sing supp ((P (D)u) ∗ f ) .
Take ϕ ∈ Cc∞ (Rn ) with ϕf ∈ C ∞ , f = P (D)u but
(supp F + x) ∩ supp(ϕ) = 0 .
Then f = f1 + f2 , f1 = ϕf ∈ Cc∞ (Rn ) so
f ∗ F = f1 ∗ F + f2 ∗ F
where f1 ∗ F ∈ C ∞ (Rn ) and x ∈
/ supp(f2 ∗ F ). It follows that x ∈
/
sing supp(u).
Example 11.1. If u is holomorphic on Rn , ∂u = 0, then u ∈ C ∞ (Rn ).
Recall from last time that a differential operator P (D) is said to be
hypoelliptic if there exists F ∈ S � (Rn ) with
(11.27) P (D)F − δ ∈ C ∞ (Rn ) and sing supp(F ) ⊂ {0} .
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 � (Rn ) then
sing supp(u) = sing supp(P (D)u) .
LECTURE NOTES FOR 18.155, FALL 2004 75
However even the Laplacian, Δ = nj=1 Dj2 , does not satisfy this rather
�
stringent condition.
It is reasonable to expect the top order derivatives to be the most
important. We therefore consider
�
Pm (ξ) = Cα ξ α
|α|=m
Now x� α F = (−1)|α| D α F
ξ , so what we need to cinsider is the behaviour
�
of the derivatives of F�, which is just Q(ξ) in (11.32).
Lemma 11.14. Let P (ξ) be a polynomial of degree m satisfying
(11.35) |P (ξ)| ≥ C |ξ|m in |ξ | > 1/C for some C > 0 ,
then for some constants Cα
� �
� α 1 � −m−|α|
(11.36) � P (ξ) � ≤ Cα |ξ | in |ξ | > 1/C .
�D �
for 0 �= x̄ ∈ Rn . Since two points give the same halfline if and only if
they are positive multiples of each other, this means we think of the
sphere as the quotient
(12.1) 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 (12.1)
does not require a choice of metric.
Now, suppose we consider functions on Rn \{0} which are (positively)
homogeneous of degree 0. That is f (a¯ x) = f (¯
x), for all a > 0, and
they are just functions on Sn−1 . Smooth functions on Sn−1 correspond
(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
(12.2) 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,
(12.3) 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.
84 RICHARD B. MELROSE
for some seminorm on S(Rn ). Thus the map (12.3) is actually continu
ous. This continuity means that ψR is a multiplier on S � (Rn ), defined
as usual by duality:
(12.7) ψR u(f ) = u(ψR f ) ∀ f ∈ S(Rn ).
Definition 12.2. The conesupport and conesingularsupport of a tem
pered distribution are the subsets Csp(u) ⊂ Rn ∪ Sn−1 and Css(u) ⊂
Rn ∪ Sn−1 defined by the conditions
(12.8)
Csp(u) ∩ Rn = supp(u)
(Csp(u))� ∩ Sn−1 ={ω ∈ Sn−1 ;
∃ R > 0, ψ ∈ C ∞ (Sn−1 ), ψ(ω) =
� 0, ψR u = 0},
Css(u) ∩ Rn = singsupp(u)
(Css(u))� ∩ Sn−1 ={ω ∈ Sn−1 ;
∃ R > 0, ψ ∈ C ∞ (Sn−1 ), ψ(ω) =
� 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
(12.9) Rn � x �−→ ∈ {y ∈ Rn ; |y| ≤ 1} = Bn .
�x�
LECTURE NOTES FOR 18.155, FALL 2004 85
Clearly |x| < �x� and for 0 ≤ a < 1, |x| = a�x� has only the solution
1
|x| = a/(1 − a2 ) 2 . Thus if we combine (12.9) with the identification of
Sn with the unit sphere we get an identification
(12.10) Rn ∪ Sn−1 � Bn .
Using this identification we can, and will, regard Csp(u) and Css(u) as
subsets of Bn .21
Lemma 12.3. For any u ∈ S � (Rn ), Csp(u) and Css(u) are closed
subsets of Bn and if ψ˜ ∈ C ∞ (Sn ) has supp(ψ̃) ∩ Css(u) = ∅ then for R
sufficiently large ψ˜R u ∈ S(Rn ).
Proof. Directly from the definition we know that Csp(u) ∩ Rn is closed,
as is Css(u) ∩ Rn . Thus, in each case, we need to show that if ω ∈ Sn−1
and ω ∈/ Csp(u) then Csp(u) is disjoint from some neighbourhood of ω
n
in B . However, by definition,
� 0} ∪ {ω � ∈ Sn−1 ; ψ(ω � ) =
U = {x ∈ Rn ; ψR (x) = � 0}
is such a neighbourhood. Thus the fact that Csp(u) is closed follows
directly from the definition. The argument for Css(u) is essentially the
same.
The second result follows by the use of a partition of unity on Sn−1 .
Thus, for each point in supp(ψ) ⊂ Sn−1 there exists a conic localizer for
which ψR u ∈ S(Rn ). By compactness we may choose a finite number of
these functions ψj such that the open sets {ψj (ω) > 0} cover supp(ψ̃).
By assumption (ψj )Rj u ∈ S(Rn ) for some Rj > 0. However this will
remain true if Rj is increased, so we may suppose that Rj = R is
independent of j. Then for function
�
µ= |ψj |2 ∈ C ∞ (Sn−1 )
j
(12.13)
{u ∈ S � (Rn ); Css(u) ⊂ K1 } × {u ∈ S � (Rn ); Css(u) ⊂ K2 } � (u1 , u2 )
−→ u1 (u2 ) ∈ C.
One can also give a similar discussion of the convolution of two tem
pered distributions. Once again we do not have a definition of u∗v as a
tempered distribution for all u, v ∈ S � (Rn ). We do know how to define
the convolution if either u or v is compactly supported, or if either is
in S(Rn ). This leads directly to
Lemma 12.6. If Css(u) ∩ Sn−1 = ∅ then u ∗ v is defined unambiguously
by
x
(12.21) u ∗ v = u1 ∗ v + u2 ∗ v , u1 = (1 − χ( ))u, u2 = u − u1
r
∞ n
where χ ∈ Cc (R ) has χ(x) = 1 in |x| ≤ 1 and R is sufficiently large;
there is a similar definition if Css(v) ∩ Sn−1 = ∅.
Proof. Since Css(u) ∩ Sn−1 = ∅, we know that Css(u1 ) = ∅ if R is large
enough, so then both terms on the right in (12.21) are welldefined. To
see that the result is independent of R just observe that the difference
of the righthand side for two values of R is of the form w ∗ v − w ∗ v
with w compactly supported. �
Now, we can go even further using a slightly more sophisticated
decomposition based on
Lemma 12.7. If u ∈ S � (Rn ) and Css(u) ∩ Γ = ∅ where Γ ⊂ Sn−1 is a
closed set, then u = u1 + u2 where Csp(u1 ) ∩ Γ = ∅ and u2 ∈ S(Rn ); in
fact
(12.22) u = u�1 + u��1 + u2 where u�1 ∈ Cc−∞ (Rn ) and
/ supp(u��1 ), x ∈ Rn \ {0}, x/|x| ∈ Γ =⇒ x ∈
0∈ / supp(u��1 ).
Proof. A covering argument which you should provide. �
Let Γi ⊂ Rn , i = 1, 2, be closed cones. That is they are closed sets
such that if x ∈ Γi and a > 0 then ax ∈ Γi . Suppose in addition that
(12.23) Γ1 ∩ (−Γ2 ) = {0}.
That is, if x ∈ Γ1 and −x ∈ Γ2 then x = 0. Then it follows that for
some c > 0,
(12.24) x ∈ Γ1 , y ∈ Γ2 =⇒ |x + y| ≥ c(|x| + |y |).
To see this consider x + y where x ∈ Γ1 , y ∈ Γ2 and |y | ≤ |x|. We
can assume that x �= 0, otherwise the estimate is trivially true with
c = 1, and then Y = y/|x| ∈ Γ1 and X = x/|x| ∈ Γ2 have |Y | ≤ 1 and
|X| = 1. However X + Y �= 0, since |X | = 1, so by the continuity of the
sum, |X + Y | ≥ 2c > 0 for some c > 0. Thus |X + Y | ≥ c(|X| + |Y |)
and the result follows by scaling back. The other case, of |x| ≤ |y|
LECTURE NOTES FOR 18.155, FALL 2004 89
Cc−∞ (Rn ). In that case we know that the singular support of the con
volution is contained in the first term in (12.29), so it is enough to
consider the conic singular support in the sphere at infinity. Thus, if
ω ∈/ Css(v) we need to show that ω ∈ / Css(u ∗ v). Using Lemma 12.7
we can decompose v = v1 + v2 + v3 as a sum of a Schwartz term, a
compact supported term and a term which does not have ω in its conic
support. Then u ∗ v1 is Schwartz, u ∗ v2 has compact support and sat
isfies (12.29) and ω is not in the cone support of u ∗ v3 . Thus (12.29)
holds in general. �
Lemma 12.10. If u, v ∈ S � (Rn ) and ω ∈ Css(u) ∩ Sn−1 =⇒ −ω ∈ /
Css(v) then their convolution is defined unambiguously, using the pair
ing in Lemma 12.5, by
(12.30) v ∗ φ) ∀ φ ∈ S(Rn ).
u ∗ v(φ) = u(ˇ
Proof. Since v(x)
ˇ v) = − Css(v) so applying Lemma 12.8
= v(−x), Css(ˇ
we know that
(12.31) Css(v̌ ∗ φ) ⊂ − Css(v) ∩ Sn−1 .
Thus, Css(v) ∩ Css(v̌ ∗ φ) = ∅ and the pairing on the right in (12.30)
is welldefined by Lemma 12.5. Continuity follows from your work in
Problem 78. �
In Problem 79 I ask you to get a bound on Css(u ∗ v) ∩ Sn−1 under
the conditions in Lemma 12.10.
Let me do what is actually a fundamental computation.
Lemma 12.11. For a conic cutoff, ψR , where ψ ∈ C ∞ (Sn−1 ),
(12.32) �R ) ⊂ {0}.
Css(ψ
Proof. This is actually much easier than it seems. Namely we already
know that Dα (ψR ) is smooth and homogeneous of degree −|α| near
infinity. From the same argument it follows that
(12.33) Dα (xβ ψR ) ∈ L2 (Rn ) if |α| > |β | + n/2
since this is a smooth function homogeneous of degree less than −n/2
near infinity, hence squareintegrable. Now, taking the Fourier trans
form gives
(12.34) ξ α Dβ (ψ
�R ) ∈ L2 (Rn ) ∀ |α| > |β | + n/2.
If we localize in a cone near infinity, using a (completely unrelated)
cutoff ψR� � (ξ) then we must get a Schwartz function since
(12.35)
�R ) ∈ L2 (Rn ) ∀ |α| > |β | + n/2 =⇒ ψ � � (ξ)ψ
|ξ||α| ψR� � (ξ)Dβ (ψ R
�R ∈ S(Rn ).
LECTURE NOTES FOR 18.155, FALL 2004 91
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) =
� 0
∞ n−1
or a ψR where ψ ∈ C (S ) has ψ(p) = � 0 – such that q ∈/ Css(Φu).
�
One crucial property is
Lemma 12.13. If (p, q) ∈ / WFsc (u) then if p ∈ Rn there exists a
neighbourhood U ⊂ R of p and a neighbourhood U ⊂ Bn of q such
n
Thus
� ��
� u = v ∗ ψR u, v̌ = ω �� .
(12.40) �
ψ � �
R R
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
(16.21) �
Pb (φ, ψ) : H×H � (φ, ψ) �−→ dµφ,ψ ∈ R, |Pb (φ, ψ)| ≤ �A��φ��ψ�.
[m,b]
17. Problems
Problem 1. Prove that u+ , defined by (1.10) is linear.
Problem 2. Prove Lemma 1.8.
Hint(s). All functions here are supposed to be continuous, I just
don’t bother to keep on saying it.
(1) Recall, or check, that the local compactness of a metric space
X means that for each point x ∈ X there is an � > 0 such that
the ball {y ∈ X; d(x, y) ≤ δ} is compact for δ ≤ �.
(2) First do the case n = 1, so K � U is a compact set in an open
subset.
(a) Given δ > 0, use the local compactness of X, to cover K
with a finite number of compact closed balls of radius at
most δ.
(b) Deduce that if � > 0 is small enough then the set {x ∈
X; d(x, K) ≤ �}, where
d(x, K) = inf d(x, y),
y∈K
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 � = K ∩ U1� and show that the
inductive hypothesis applies to K � and the Uj for j > 1; let
fj� , j = 2, . . . , n be the functions supplied by the inductive
assumption and put f = j≥2 fj� .
�
�
with the series on the right always absolutely convergenct (i.e., this is
part of the requirement on µ). Define
∞
�
(17.2) |µ| (E) = sup |µ(Ei )|
i=1
i.e, the inner product is recoverable from the norm, so use the RHS
(right hand side) to define an inner product on the vector space. You
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
Problem 18. Consider the function f (x) = �x�−1 = (1+ |x|2 )−1/2 . Show
that
∂f
= lj (x) · �x�−3
∂xj
with lj (x) a linear function. Conclude by induction that �x�−1 ∈
C0k (Rn ) for all k.
I suggest induction!
Problem 26. Prove the generalization of Proposition 8.10 that u ∈
S � (Rn ), supp(w) ⊂ {0} implies there are constants cα , |α| ≤ m, for
some m, such that
�
u= cα D α δ .
|α|≤m
108 RICHARD B. MELROSE
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, m� > 0 show that u ∈ H m (Rn ) and Dα u ∈
� �
H m (Rn ) for all |α| ≤ m implies u ∈ H m+m (Rn ). Is the converse true?
Problem 28. Show that every element u ∈ L2 (Rn ) can be written as a
sum n
�
u = u0 + Dj uj , uj ∈ H 1 (Rn ) , j = 0, . . . , n .
j=1
This is called the weak topology (because there are very few open
sets). Show that uj → u weakly in S � (Rn ) means that for every open
set U � u ∃N st. uj ∈ U ∀ j ≥ N .
Problem 35. Prove (11.18) where u ∈ S � (Rn ) and ϕ, ψ ∈ S(Rn ).
Problem 36. Show that for fixed v ∈ S � (Rn ) with compact support
S(Rn ) � ϕ �→ v ∗ ϕ ∈ S(Rn )
is a continuous linear map.
Problem 37. Prove the ?? to properties in Theorem 11.6 for u ∗ v where
u ∈ S � (Rn ) and v ∈ S � (Rn ) with at least one of them having compact
support.
Problem 38. Use Theorem 11.9 to show that if P (D) is hypoelliptic
then every parametrix F ∈ S(Rn ) has sing supp(F ) = {0}.
Problem 39. Show that if P (D) is an ellipitic differential operator of
order m, u ∈ L2 (Rn ) and P (D)u ∈ L2 (Rn ) then u ∈ H m (Rn ).
Problem 40 (Taylor’s theorem). . Let u : Rn −→ R be a realvalued
function which is k times continuously differentiable. Prove that there
is a polynomial p and a continuous function v such that
|v(x)|
u(x) = p(x) + v(x) where lim = 0.
|x|↓0 |x|k
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
�
n
(17.3) C(B ) � f �−→ f (x)dx.
Bn
supp(u) ∩ supp(φ) = ∅
then u(φ) = 0.
iii) Consider the space C ∞ (Rn ) of all smooth functions on Rn , with
out restriction on supports. Show that for each N
Show that such a v ‘is’ a distribution and that the map E � (Rn ) −→
C −∞ (Rn ) is injective.
vi) Show that if v ∈ E � (Rn ) satisfies (17.11) and f ∈ C ∞ (Rn ) has
f = 0 in |x| < N + � for some � > 0 then v(f ) = 0.
vii) Conclude that each element of E � (Rn ) has compact support
when considered as an element of C −∞ (Rn ).
viii) Show the converse, that each element of C −∞ (Rn ) with compact
support is an element of E � (Rn ) ⊂ C −∞ (Rn ) and hence conclude
that E � (Rn ) ‘is’ the space of distributions of compact support.
I will denote the space of distributions of compact support by Cc−∞ (R).
� 0 where it
(3) Use an inductive argument to show that, in (τ, ξ) =
makes sense,
|α|
1 � qk,α,j (ξ)
(17.12) Dτk Dξα =
p(τ, ξ) j=1
p(τ, ξ)k+j+1
Problem 61. Wavefront set computations and more – all pretty easy,
especially if you use results from class.
i) Compute WF(δ) where δ ∈ S � (Rn ) is the Dirac delta function
at the origin.
ii) Compute WF(H(x)) where H(x) ∈ S � (R) is the Heaviside func
tion
�
1 x>0
H(x) = .
0 x≤0
Hint: Dx is elliptic in one dimension, hit H with it.
iii) Compute WF(E), E = iH(x1 )δ(x� ) which is the Heaviside in
the first variable on Rn , n > 1, and delta in the others.
iv) Show that Dx1 E = δ, so E is a fundamental solution of Dx1 .
v) If f ∈ Cc−∞ (Rn ) show that u = E � f solves Dx1 u = f.
vi) What does our estimate on WF(E � f ) tell us about WF(u) in
terms of WF(f )?
Problem 62. The wave equation in two variables (or one spatial vari
able).
LECTURE NOTES FOR 18.155, FALL 2004 119
is a nonincreasing function of t.
Hint: Multiply the equation by u and integrate over a slab
[t1 , t2 ] × Rn .
LECTURE NOTES FOR 18.155, FALL 2004 121
where the vj form a basis of Rn and using the dual basis wj (so wj · vi =
δij is 0 or 1 as i �= j or i = j) set
� n
�
�
L◦ = w = 2π kj wj , kj ∈ Z .
j=1
ii) Show that there exists f ∈ Cc∞ (Rn ) such that AL f ≡ 1 is the
costant function on Rn .
iii) Show that the map (17.19) is surjective. Hint: Well obviously
enough use the f in part ii) and show that if u is periodic then
AL (uf ) = u.
iv) Show that the infinite sum
�
(17.20) F = δ(· − v) ∈ S � (Rn )
v∈L
V ⊥ = {f ∈ H; �f, v� = 0 ∀ v ∈ V }.
LECTURE NOTES FOR 18.155, FALL 2004 123
N
� N
�
2
� aj ej � = |aj |2 .
j=1 j=1
124 RICHARD B. MELROSE
Hint: Problem 70 shows that (17.26) holds for each u ∈ L2 (Rn ); check
that (17.27) 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 con
verse is to show that if (17.26) and (17.27) hold then B is bounded
and to use this to extract a weakly convergent sequence from any given
sequence in B. Next show that (17.26) is equivalent to (17.27) for the
set F (B), the image of B under the Fourier transform. Show, possi
bly using Problem 71, that if χR is cutoff to a ball of radius R then
χR G(χR uˆn ) converges strongly if un converges weakly. Deduce from
this that the weakly convergent subsequence in fact converges strongly
so B¯ is sequently compact, and hence is compact.
Each of the space on the right is a Banach space for the supremum
norm.
LECTURE NOTES FOR 18.155, FALL 2004 127
(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 (17.28) 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 (6.16) means
precisely that for each n ∈ N there exists k = k(n) and C = Cn such
that
(17.29) |u(ϕ)| ≤ C�ϕ�C 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 embedding
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 map
1 1
(17.30) H m (Rn ) −→ H m− 2 (Rn ) if m >
2
with the following properties:
(1) On S (Rn ) it is given by u �−→ u(0, x� ), x� ∈ 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
�
� −1
(17.31) �
u(0, ·)(ξ ) = (2π) ˆ 1 , ξ � )dξ1 , ∀ ξ � ∈ Rn−1 .
u(ξ
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 welldefined and unique.
Problem 76. [Restriction by WF] From class we know that the product
of two distributions, one with compact support, is defined provided
128 RICHARD B. MELROSE
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, x� ), (ω1 , 0)) ∈
/ WF(u)
as the product
(17.33) u0 = uδ(x1 ).
supp(ψ � ) ∩ Css(u) = ∅
show that
S(Rn ) � φ �−→ φψ � u ∈ S(Rn )
is continuous and hence (or otherwise) show that the functional u1 u2
defined by (12.20) is an element of S � (Rn ).
Solution 18.3 (To Problem 40). (Matjaž Konvalinka) Let us prove this
� 0)
in the case where n = 1. Define (for b =
(b − x)k−1 (k−1)
U (x) = u(b) − u(x) − (b − x)u� (x) − . . . − u (x);
(k − 1)!
then
(b − x)k−1 (k)
U � (x) = − u (x).
(k − 1)!
For the continuously differentiable function V (x) = U (x)−(1−x/b)k U (0)
we have V (0) = V (b) = 0, so by Rolle’s theorem there exists ζ between
0 and b with
k(b − ζ)k−1
V � (ζ) = U � (ζ) + U (0) = 0
bk
Then
bk
U (0) = − U � (ζ),
k(b − ζ)k−1
u(k−1) (0) k−1 u(k) (ζ) k
u(b) = u(0) + u� (0)b + . . . + b + b .
(k − 1)! k!
The required decomposition is u(x) = p(x) + v(x) for
u�� (0) 2 u(k−1) (0) k−1 u(k) (0) k
p(x) = u(0) + u� (0)x + x + ... + x + x ,
2 (k − 1)! k!
� l! ∂lu
wx(l) (t) = l1 l2 li
(tx)xl11 xl22 · · · xlii
l1 !l2 ! · · · li ! ∂x1 ∂x2 · · · ∂xi
l1 +l2 +...+l =l
i
132 RICHARD B. MELROSE
m� m�
and since �ξ� ξ α u
� = �ξ� D � α u is in L2 (Rn ) (note that u ∈ H m (Rn )
�
follows from Dα u ∈ H m (Rn ), |α| ≤ m). The converse is also true since
Cα in the formula above are strictly positive.
Solution 18.7. Take v ∈ L2 (Rn ), and define subsets of Rn by
E0 = {x : |x| ≤ 1},
Ei = {x : |x| ≥ 1, |xi | = max |xj |}.
j
�n �n
Then obviously we have 1 = i=0 χEj a.e., and v = j=0 vj for vj =
√
χEj v. Then �x� is bounded by 2 on E0 , and �x�v0 ∈ L (Rn ); and on
2
Ej , 1 ≤ j ≤ n, we have
�x� (1 + n|xj |2 )1/2 � �1/2
≤ = n + 1/|xj |2 ≤ (2n)1/2 ,
|xj | |xj |
134 RICHARD B. MELROSE
2 n
�n�x�vj = xj wj for w2j ∈n L (R ). But that means that �x�v = w0 +
so
j=1 xj wj for wj ∈ L (R ).
If u is in L2 (Rn ) then u
� ∈ L2 (Rn ), and so there exist w0 , . . . , wn ∈
L2 (Rn ) so that
� n
�ξ�u
� = w0 + ξj wj ,
j=1
in other words
n
�
u
�=u
�0 + ξj u
�j
j=1
�j ∈ L2 (Rn ). Hence
where �ξ�u
n
�
u = u0 + Dj uj
j=1
where uj ∈ H 1 (Rn ).
Solution 18.8. Since
� ∞ � ∞
�
Dx H(ϕ) = H(−Dx ϕ) = i H(x)ϕ (x) dx = i ϕ� (x) dx = i(0−ϕ(0)) = −iδ(ϕ),
−∞ 0
this is equivalent to finding m such that �ξ�2m has a finite integral over
Rn . One option is to write �ξ� = (1 + r2 )1/2 in spherical coordinates,
and to recall that the Jacobian of spherical coordinates in n dimensions
has the form rn−1 Ψ(ϕ1 , . . . , ϕn−1 ), and so �ξ�2m is integrable if and only
if � ∞
rn−1
dr
0 (1 + r2 )m
converges. It is obvious that this is true if and only if n − 1 − 2m < −1,
ie. if and only if m > n/2.
Solution 18.10 (Solution to Problem31). We know that δ ∈ H m (Rn )
for any m < −n/1. Thus is just because �ξ�p ∈ L2 (Rn ) when p < −n/2.
Now, divide Rn into n + 1 regions, as above, being A0 = {ξ; |ξ | ≤ 1 and
Ai = {ξ; |ξi | = supj |ξj |, |ξ | ≥ 1}. Let v0 have Fourier transform χA0
and for i = 1, . . . , n, vi ∈ S; (Rn ) have Fourier transforms ξi−n−1 χAi .
Since |ξi | > c�ξ� on the support of v�i for each i = 1, . . . , n, each term
LECTURE NOTES FOR 18.155, FALL 2004 135
References
[1] G.B. Folland, Real analysis, Wiley, 1984.
[2] F. G. Friedlander, Introduction to the theory of distributions, second ed., Cam
bridge University Press, Cambridge, 1998, With additional material by M. Joshi.
MR 2000g:46002
[3] J. Hadamard, Le problème de Cauchy et les `equatons aux d´eriv´ees partielles
linéaires hyperboliques, Hermann, Paris, 1932.
[4] L. Hörmander, The analysis of linear partial differential operators, vol. 3,
SpringerVerlag, Berlin, Heidelberg, New York, Tokyo, 1985.
[5] W. Rudin, Real and complex analysis, third edition ed., McGrawHill, 1987.
[6] George F. Simmons, Introduction to topology and modern analysis, Robert E.
Krieger Publishing Co. Inc., Melbourne, Fla., 1983, Reprint of the 1963 original.
MR 84b:54002