Gen Fun D
Gen Fun D
Gen Fun D
Contents
0. Preface 1
1. The space of generalized functions on Rn 1
1.1. Motivation 1
1.2. Basic definitions 2
1.3. Remarks on operations on distributions 5
1.4. Translations of generalized functions 5
1.5. Derivatives of generalized functions 5
1.6. The support of generalized functions 6
1.7. Products and convolutions of generalized functions 7
1.8. Generalized functions on Rn 9
1.9. Generalized functions and differential operators 9
1.10. Regularization of generalized functions 10
2. Topological properties of Cc∞ (Rn ) 12
2.1. Normed spaces 12
2.2. Topological vector spaces 13
2.3. Defining completeness 14
2.4. Fréchet spaces 17
2.5. Sequence spaces 19
2.6. Direct limits of Fréchet spaces 20
2.7. Topologies on the space of distributions 21
3. Geometric properties of C −∞ (Rn ) 22
3.1. Sheaf of distributions 22
3.2. Filtration on spaces of distributions 24
3.3. Functions and distributions on a Cartesian product 26
4. p-adic numbers and `-spaces 27
4.1. Defining p-adic numbers 27
4.2. Misc. -not sure what to do with them (add to an appendix about
p-adic numbers?) 30
4.3. p-adic expansions 32
4.4. Inverse limits 32
4.5. Haar measure and local fields 33
4.6. Some basic properties of `-spaces. 34
0. Preface
??
GENERALIZED FUNCTIONS LECTURES 3
1.1. Motivation. One of the most basic and important examples of a generalized
function is the Dirac delta function. The Dirac delta function on R at point t is
usually denoted
by δt , and while it is not a function, it can be intuitively described
∞ x = t ´∞
by δt (x) := , and by satisfying the equality δt (x)dx = 1. Notice
0 x 6= t −∞
1.2. Basic definitions. In this book we will consider various spaces of functions
and functionals. Unless specified otherwise, all the functions and functionals will be
real-valued. All the statements below are also valid for complex-valued functions.
In order to define what is a generalized function we first need to introduce some
standard notation.
Definition 1.2.2 (Convergence in Cc∞ (R)). Given f ∈ Cc∞ (R) and a sequence
n=1 of smooth functions with compact support we say that {fn }n=1 converges
{fn }∞ ∞
to f in Cc (R) if:
∞
S
(i) There exists a compact set K ⊂ R for which supp(fn ) ⊆ K.
n∈N
(k) ∞
(ii) For every order k ∈ N0 , the derivatives (fn )n=1 converge uniformly to the
derivative f (k) .
We can now define the notion of distributions (cf. [?, ??] and [Kan04, Section 2.3].
have
lim hξ, fm i = hξ, lim fm i.
m→∞ m→∞
We will usually use the notation hξ, f i instead of ξ(f ). We call a continuous linear
functional a distribution or a generalized function. The space of all generalized
functions on R is denoted by C −∞ (R) := (Cc∞ (R))∗ .
Remark 1.2.5. In §?? below we will define a natural topology on the space Cc∞ (R).
The convergence in this topology will be as in Definition 1.2.2, but this does not
GENERALIZED FUNCTIONS LECTURES 5
define the topology uniquely since this topology is not first countable. We will show
that a linear functional on this topological space is continuous if and only if it
satisfies the condition in Definition 1.2.4.
Remark 1.2.6. For now the names generalized functions and distributions are
synonymous as there is no difference for R. We will discuss the difference in a
later part of the manuscript, when it will be relevant.
Recall that a function f is locally-L1 if the restriction to any compact subset in its
domain is an L1 function. We denote the space of such functions L1Loc . Given a
real-valued function f ∈ L1Loc (R) we define ξf : Cc∞ (R) → R to be the generalized
function
ˆ∞
hξf , φi := f (x) ∙ φ(x)dx.
−∞
Note that this integral converges as φ vanishes outside of some compact set K, and
(f φ)|K ∈ L1 (K). These are sometimes called regular generalized functions.
Exercise 1.2.8. For any f ∈ L1Loc (R), show that ξf is a well defined distribution.
where the last embedding is given by f 7→ ξf . This embedding motivates the name
generalized function.
Exercise 1.2.9. Prove that there exists a function f ∈ Cc∞ (R) which is not the
2
zero function. Hint: Use functions such as e−1/(1−x) .
n=1 weakly
Definition 1.2.10. We say that a sequence of generalized functions {ξn }∞
converges to ξ ∈ C −∞
(R) if for every f ∈ Cc (R) we have
∞
Definition 1.2.11. (i) A sequence {fn } in L1Loc (R) is called a weakly Cauchy
sequence if for every g ∈ Cc∞ (R) and > 0 there exists a number N ∈ N such
´∞
that for all m, n > N we have that (fn (x) − fm (x))g(x)dx < .
−∞
GENERALIZED FUNCTIONS LECTURES 6
(ii) Two weakly Cauchy sequences are called equivalent if their difference weakly
converges to zero.
However, weakly Cauchy sequences in Cc∞ (R) do not necessarily converge in Cc∞ (R).
Remark 1.2.13. One can define the space of generalized functions C −∞ (R) as the
space of equivalence classes of weakly Cauchy sequences in Cc∞ (R). As we will show
in ??, this definition is equivalent to Definition 1.2.4. It is important that we take
weakly Cauchy sequences rather than weakly Cauchy nets, since otherwise we would
get the full completion of Cc∞ (R), which is larger than C −∞ (R), as we will see in
??.
One can find a weakly Cauchy sequence that converges to the Dirac’s delta.
The reason for the name “approximation of identity” is that δ0 is the identity for
the convolution operation that we will define later.
Such sequences can be generated, for example, by starting with a non-negative,
continuous, compactly supported function φ1 of total integral 1, and by then setting
φn (x) = nφ1 (nx).
Note that given η ∈ C −∞ (R) of the form η = ξf , we can recover the value of f at
´∞
t via lim hξf , φn (x − t)i = lim f (x)φn (x − t)dx = f (t).
n→∞ n→∞ −∞
ˆ∞
hξ , φi = f (x) ∙
f0 φ(x)|∞
−∞ − f (x) ∙ φ0 (x)dx.
−∞
−∞ = 0.
However, since φ and f have compact support, we know that f (x) ∙ φ(x)|∞
Thus, hξf 0 , φi = −hf, φ i. This motivates the following definition.
0
hξ 0 , φi := −hξ, φ0 i
1.6. The support of generalized functions. Let U ⊂ R be an open set and let
Cc∞ (U ) be the space of smooth functions f : U → R supported in some compact
subset of U . Given a compact subset K ⊂ R, we denote by CK ∞
(R) the space of
smooth functions f : R → R with supp(f ) ⊆ K. In particular CK ∞
(X) ⊆ Cc∞ (R)
for every compact K ⊆ R.
We cannot evaluate a generalized function at a point. Therefore, we cannot just
define its support as we did before for a function by supp(f ) := {x ∈ R | f (x) 6= 0}.
However, if for some neighborhood U ⊂ R we have for every f ∈ Cc∞ (U ) that
hξ, f i = 0, then it is natural to say that supp(ξ) ⊆ R r U . This leads us to the
following definition:
(iii) Denote by Cc−∞ (R) the space of distributions with compact support.
Remark 1.6.5. While the support of δ00 is also {0}, given some f ∈ C ∞ (R) for
which f (0) = 0 but f 0 (0) 6= 0, we get that hδ00 , f i = −hδ0 , f 0 i = −f 0 (0) 6= 0. In
other words, having f (0) = 0 is not enough to get hξ, f i = 0, for a distribution ξ
supported at {0}. However, if f vanishes at 0 with all its derivatives, it will imply
hξ, f i = 0 for any ξ supported at {0}, as follows from Exercise 1.6.7 below.
Exercise 1.6.7. Show that all the generalized functions ξ ∈ C −∞ (R) which are
P
n
supported on {0} are of the form ci δ (i) for some n ∈ N and ci ∈ R.
i=0
Hint: prove this in three steps.
(i) Show that there exists n such that ξ is bounded on the set {f | f (i) (x) < 1∀x ∈
R, ∀i < n}.
(ii) Show that there exists k ∈ N such that ξxk = 0, that is hξxk , f i = hξ, xk f i = 0
for every f ∈ Cc∞ (R).
P
k−1
(i)
(iii) From ξxk = 0 deduce that ξ = ci δ0 for some ci ∈ R.
i=0
While we can multiply every smooth function f by any generalized function ξ, the
product of two generalized functions is not always defined. Notice that indeed the
product of two weakly Cauchy sequences is not always a weakly Cauchy sequence,
so we might not be able to approximate the product of two generalized functions
by the product of their approximations.
Recall that given two functions f, g, their convolution is defined by
ˆ∞
(f ∗ g)(x) := f (t) ∙ g(x − t)dt.
−∞
Exercise 1.7.2.
(1) Show that supp(f ∗ g) ⊆ supp(f ) + supp(g), where supp(f ) + supp(g) is
the Minkowski sum of supp(f ) and supp(g). Thus f, g ∈ Cc∞ (R) implies
f ∗ g ∈ Cc∞ (R).
(2) Find an example in which the left hand side is strictly contained in the right
hand side.
GENERALIZED FUNCTIONS LECTURES 10
Given f, g ∈ Cc∞ (R) we can write (f ∗ g)(x) = hξf , g̃x i, where g̃x (t) := g(x − t).
This motivates us to define the convolution ξ ∗ g as the function (ξ ∗ g)(x) = hξ, g˜x i.
Note that the convolution of a smooth function and a generalized function is always
a smooth function:
Exercise 1.7.3. Show that for φ ∈ Cc∞ (R) and ξ ∈ C −∞ (R) we get that ξ ∗ φ is a
smooth function.
Let us now define the convolution of two generalized functions. This will not be
defined for every pair of generalized functions, but for pairs such that at least one
of the generalized functions have compact support. Firstly, for f, g ∈ Cc∞ (R) we
would like to have ξf ∗ ξg = ξf ∗g . This means
ˆ∞ ˆ∞ ˆ∞
hξf ∗ ξg , φi = (f ∗ g)(x) ∙ φ(x)dx = f (t) ∙ g(x − t) ∙ φ(x)dtdx.
−∞ −∞ −∞
We would like like to express the right-hand side in terms of convolutions of dis-
tributions with functions. For this purpose, for a function h ∈ C ∞ (R) denote
h := h(−x). In these terms we have
ˆ∞ ˆ∞ ˆ∞ ˆ∞ ˆ∞
f (t)∙g(x−t)∙φ(x)dtdx = f (t) ˉ
g(x−t)∙φ(−x)dxdt = ˉ
f (t)∙(ξg ∗φ)(−t)dt ˉ
= ξf (ξg ∗ φ).
−∞ −∞ −∞ −∞ −∞
ˉ
Definition 1.7.4. We define hξf ∗ ξg , φi := hξf , ξg ∗ φi.
Exercise 1.7.5. Let K and V be as above. Show that there exists a continuous
cutoff function. Hint: use Urysohn’s Lemma.
Thus, given some ξ ∈ Cc−∞ (R) with supp(ξ) ⊂ K we have that ξ(φ) = ξ(ρK ∙ φ).
This enables us to define ξ as a functional over all C ∞ (R) and not only on Cc∞ (R).
For every φ ∈ C ∞ (R) we define ξ(φ) = ξ(ρK ∙ φ) where K := supp(ξ) ⊂ R.
Exercise 1.7.7. Show the following identities for any compactly supported distri-
butions ξ1 , ξ2 and ξ3 in C −∞ (R).
GENERALIZED FUNCTIONS LECTURES 11
(1) δ0 ∗ ξ1 = ξ1 .
(2) δ00 ∗ ξ1 = ξ10 .
(3) ξ1 ∗ ξ2 = ξ 2 ∗ ξ 1 .
(4) ξ1 ∗ (ξ2 ∗ ξ3 ) = (ξ1 ∗ ξ2 ) ∗ ξ3 .
(5) (ξ1 ∗ ξ2 )0 = ξ1 ∗ ξ20 = ξ10 ∗ ξ2 .
1.8. Generalized functions on Rn . All the notions above make sense for func-
tions and generalized functions in several variables. The definitions and the state-
ments literally generalize to this case. For example, let us restate the definition of
convergence in Cc∞ (Rn ).
Definition 1.8.1 (Convergence in Cc∞ (Rn )). Given f ∈ Cc∞ (R) and a sequence
n=1 of smooth functions with compact support, we say that {fn }n=1 converges
{fn }∞ ∞
and define the family by ξλ := xλ+ for Re(λ) > −1. The behavior of the function
changes as λ changes: When Re(λ) > 0 we have a continuous function; if Re(λ) = 0
we get a step function and for Re(λ) ∈ (−1, 0), the function xλ+ will not be locally
bounded. However, for any λ with Re(λ) > −1, ξλ ∈ L1loc (R).
We would like to extend the definition analytically to the entire complex plane. To
do that we will have to leave the world of locally L1 functions.
Deriving xλ+ (both as a complex function or as we defined for generalized function)
we obtain ξλ0 = λ ∙ ξλ−1 . This is a functional equation which enables us to define
ξ0
ξλ := λ+1
λ+1
, and in this way extend ξλ to every λ ∈ C with Re(λ) > −2, except
λ = −1. We can use this procedure interatively to extend ξλ to the entire complex
plane. This extension is not analytic, but it is meromorphic: it has a pole in
λ = −1, and by the extension formula, in λ = −2, −3, . . ..
Exercise 1.10.4. Show that any meromorphic family of generalized functions can
be expanded into Laurent series in a punctured neighborhood of every pole, with
coefficients being generalized functions.
Exercise 1.10.5. Find the order and the leading coefficient of every pole of ξλ :=
xλ+ .
Exercise 1.10.7. Solve the problem of finding an analytic continuation for p+ (x1 , . . . xn )λ
in the following cases:
(1) p(x, y, z) := x2 + y 2 + z 2 − a and a ∈ R.
(2) p(x, y, z) := x2 + y 2 − z 2 .
1There was a slightly earlier proof in [?, ?], using resolution of singularities.
GENERALIZED FUNCTIONS LECTURES 14
We want to analyze the space of distributions C −∞ (Rn ). For this aim, we want to
introduce a topology on this space.
The norm defines a Hausdorff topology on V be letting the basis for open sets be
the open balls B(x, ε) := {v ∈ V | ||v − x|| < ε} for all x ∈ V , ε ∈ R>0 .
P
Example 2.1.2. (i) lp := {sequences xn in R | |xn |p < ∞}
(ii) Lp (R) := {measurable f : R → R | |f |p is integrable on R}.
(iii) C k (R) := functions with k continuous bounded derivatives,
k
X
||f || := sup |f (i) (x)|.
i=1 x∈R
There are several facts that we are used to in the Euclidean space Rn . For example,
the closed bounded sets are compact, linear subspaces are always closed, and linear
operators are always continuous. These properties still hold for finite-dimensional
normed spaces, but fail for infinite-dimensional ones.
Let us now discuss this in more detail. Fix a normed space V .
Exercise 2.1.3. If dim V is finite then the closed unit ball B := {v ∈ V | ||v|| ≤ 1}
is compact.
Exercise 2.1.5. Show that C 1 (R) ⊂ C 0 (R) is dense (and thus not closed).
Exercise 2.1.7. Show that for any two normed spaces of the same finite dimension
there exists a linear homeomorphism between them.
GENERALIZED FUNCTIONS LECTURES 15
Proposition 2.1.8. If the closed unit ball B ⊂ V is compact then dim V is finite.
S
Proof. Note that B can be covered by open balls of radius 1/2: B ⊂ x∈B B(x, 1/2).
If B is compact then this cover has a finite subcover. Denote the centers of the
subcover by {xi }ni=1 , and let W := Span({xi }ni=1 ). Then
n
[
B⊂ B(xi , 1/2) ⊂ W + 1/2B ⊂ W + 1/4B ⊂ . . .
i=1
Definition 2.2.1. A topological vector space (or linear topological space) is a linear
space with a topology such that multiplication by scalar and vectors addition are
continuous. More precisely, the addition and the multiplication define continuous
maps:
(i) + : V × V → V ,
(ii) ∙ : R × V → V .
Remark 2.2.2. In this definition V is a vector space over R, but in the same way
one defines topological vector spaces over any topological field, e.g. over C or over
the field of p-adic numbers that we will define later.
Remark 2.2.5. Note that the previous exercise shows that a linear topological space
satisfies the T1 separation axiom ⇐⇒ it satisfies T2 .
Exercise 2.2.6. Let C be an balanced open convex set in a topological vector space
V . Show that
(i) NC (x) < ∞ for all x ∈ V .
(ii) NC (x) is a semi-norm.
In a locally convex space we have a basis for the topology consisting of convex sets.
By Exercise 2.2.4(i) we can assume that all the sets in this basis are balanced.
Thus we obtain the next corollary.
Exercise 2.2.8. Find a locally convex (Hausdorff ) topological vector space V such
that V has no continuous norm on it. That is, every convex open set C contains a
line span{v}, so NC (v) = 0.
Note however that by Hausdorffness and Corollary 2.2.7, for every vector v ∈ V
there exists a semi-norm n on V such that n(v) > 0.
Generalizing the proof of Proposition 2.1.8, one can prove the following theorem.
Theorem 2.2.9 ([Rud06, Theorem 1.22]). Every locally compact topological vector
space has finite dimension.
2.3. Defining completeness. Recall that one can define the notion of Cauchy
sequence for a metric space, but not for a general topological space. However, one
can define it for any topological vector space V , in the following way.
GENERALIZED FUNCTIONS LECTURES 17
Remark 2.3.2. More generally, if X has a uniform topology, then we can define a
notion of a Cauchy sequence. We will not give the definition of a uniform topology,
but we note that any topological group possesses a uniform topology, and indeed one
can define a notion of a left (resp. right) Cauchy sequence as follows: (xn )∞n=1 is
a Cauchy sequence if for every neighborhood U of e ∈ G there is an index n0 ∈ N
such that m, n > n0 implies x−1m xn ∈ U (resp. xn xm ∈ U ).
−1
The notion of sequentially complete is well suited for topological vector spaces in
which the topology can be defined by sequences. Such are all the first-countable
(Hausdorff) topological spaces, but not all Hausdorff topological spaces, as we will
see from the following discussion.
Exercise 2.3.7 (*). (i) Show that a complete topological vector space is always
sequentially complete.
(ii) Show that a first countable topological vector space V is complete if and only
if it is sequentially complete.
sec sec
??Solution: (i) Let V be the sequential completion of V . The map V → V is
a strict embedding ?? check how to define the topology on it.
(ii): Suppose that V is sequentially complete. Let κ be an ordered filtered set and
ϕ : κ → V s.t. for every neighborhood of zero U ⊂ V there exists α ∈ κ s.t. the
GENERALIZED FUNCTIONS LECTURES 18
Minkowsky difference ϕ(κ≥α ) − ϕ(κ≥α ) lies in U . We want to prove that any such
?? converges. This means that there exists y ∈ V for any open neighborhood U 3 y
there exists β ∈ κ with ϕ(κ≥β ) ⊂ U . Indeed, choose a descending basis Ui of the
topology around zero. Let αi ∈ κ be such that ϕ(κ≥αi ) − ϕ(κ≥αi ) ⊂ h(Ui ), where
h(Ui ) is an open neighborhood of zero such that h(Ui ) + h(Ui ) ⊂ Ui . On the other
hand, ϕ(αi ) is a Cauchy sequence. Let y be its limit. Let us show that y satisfies
the condition. For that purpose, let U be a neighborhood of y. Let i be such that
ϕ(αi ) ∈ h(U ) and Ui ⊂ h(U ). We set β = αi . It is easy to see that it satisfies the
conditions above.
Now let ι : V → W be a strict embedding. We have to show that the image is
closed. Let y ∈ Imι. Let {Uα |α ∈ I} be a basis for topology at y. Let F (I) be
the set of all finite subsets of I, and for any f ∈ F (I) denote Vf := ∩α∈f Uα . For
any f ∈ F (I) choose xf ∈ f ∩ Imι. By the discussion above there exists z ∈ Imι
such that xf converges to z in the sense we discussed above. This implies (??) that
z = y.
Remark 2.3.8.
(i) Equivalently, we can define that a space V is complete if every Cauchy net is
convergent. From this definition it can be easily seen that any complete space
X is also sequentially complete.
(ii) In the category of first countable topological vector spaces, completeness is
equivalent to sequentially completeness, and indeed, there the notion of a
Cauchy net is equivalent to the notion of a Cauchy sequence, and a set Y ⊆ X
is closed ⇐⇒ it is sequentially complete.
Exercise 2.3.9. Find a sequentially complete space which is not complete. Hint:
see the above example.
Remark 2.3.11. We can also use a universal property in order to define the
completion of V . A strict (?) embedding i : V → Vˉ is a completion of V if:
(1) Vˉ is complete.
(2) For every map ψ : V → W where W is complete, there is a unique map
φW : Vˉ → W , such that ψ ≡ φW ◦ i.
Exercise 2.3.12 (*). Show that these two definitions of completeness are equiva-
lent.
It is often easier to show that a space is complete using the universal property. In
this way we avoid dealing with Cauchy nets or filters. However, in order to show
such completion exists one has to use these notions.
GENERALIZED FUNCTIONS LECTURES 19
Exercise 2.3.13.
(1) (∗) Show that every Hausdorff topological vector space has a completion.
(2) Show that in the category of first countable topological vector spaces both
definitions of completion are equivalent to being sequentially complete.
Exercise 2.4.2. Let V be a locally convex topological vector space (i.e, not neces-
sarily normed), and let f : W → R be a continuous linear functional, where W ⊆ V
is a closed linear subspace of V . Show that f can be extended to V .
Remark 2.4.6. Every normed space is Hausdorff and locally convex, since there is
a basis of its topology consisting of open balls, which are convex. We also know that
every normed space is metric. However, metrizability does not force local convexity
and vice versa.
Exercise 2.4.7. Show that for a locally convex topological vector space V the fol-
lowing three conditions are equivalent.
(1) V is metrizable.
(2) V is first countable (that is it has a countable basis of its topology at every
point).
(3) There is a countable collection of semi-norms {ni }i∈N that defines the basis
of the topology of V , i.e, Ui, = {x ∈ V |ni (x) < } is a basis of the topology
at 0.
GENERALIZED FUNCTIONS LECTURES 20
Exercise 2.4.8. Let V be a locally convex metrizable space. Prove that V is com-
plete (and consequentially is a Fréchet space) ⇐⇒ it is sequentially complete.
Recall that the completion of a normed space V with respect to its norm is the
quotient space Vˉ of all Cauchy sequences in X under the equivalence relation
(xn )∞
n=1 ∼ (yn )n=1
∞
⇐⇒ lim kxn − yn k = 0. In particular, Vˉ is a Banach
n→∞
space. Completing V with respect to a semi-norm N results in the elimination of
all elements {x ∈ V | n(x) = 0}. The quotient space equipped with the induced
norm on the quotient then yields a Banach space.
Example 2.4.9. Let V be the space of step functions on R, and consider the norm
kf k1 := R |f (x)|dx. The completion of V with respect to k ∙ k1 is isomorphic to the
´
Example 2.5.1. The space `p is the space of all sequences (xn )∞n=1 with values in
P
∞
R, such that |xn |p < ∞. It is a Banach space, and for p = 2 it is a Hilbert
n=1
space.
Let SW(N) be the space of all the sequences which decay to zero faster than any
polynomial, i.e. ∀n ∈ N, lim xi ∙ in = 0. A family of norms one can consider when
i→∞
analyzing these spaces is ||(xi )∞
n=1 ||n = sup{|xi ∙ i |}. It is not hard to see that with
n
i∈N
respect to these norms every Cauchy sequence converges. Define the topology on
SW(N) using by the family of norms || ∙ ||n , then SW(N) is a Fréchet space. This
is an example of a Fréchet space which is not a Banach space.
Remark 2.5.2. How can we see every Cauchy sequence converges? Why is not it
a Banach space?
The dual space SW(N)∗ is {(xi )∞i=1 | ∃n, c : |xi | < c ∙ i }. This is a union of Banach
n
Exercise 2.5.4.
(1) Show that the Fourier series map F : C ∞ (S 1 ) → SW(Z) via f 7→ an
is an isomorphism of Fréchet spaces. In other words, show that it is a
bijection and that for any semi-norm Pi of SW(Z) there exists a semi-
norm Sj of C ∞ (S 1 ) and C ∈ R such that for any f ∈ C ∞ (S 1 ) we have
that kF (f )kPi < C ∙ kf kSj (and vice-versa, or use Banach’s open mapping
theorem for Fréchet spaces).
GENERALIZED FUNCTIONS LECTURES 22
Definition 2.6.1. The direct limit of an ascending sequence of vector spaces is the
S
space V∞ := Vn . This is not a Fréchet space, but a locally convex topological
n∈N T
vector space. A convex subset U ⊆ V∞ is open ⇐⇒ U Vn is open in Vn for all
n.
Every space of the form C ∞ (K) can be given the induced topology from C ∞ (R).
Taking the union of the ascending chain
gives all smooth functions with compact support Cc∞ (R) = lim C ∞ ([−n, n]) as a
n→∞
direct limit. However, this is not a Fréchet space - it is a direct limit and not an
inverse limit. A basis of the topology of Cc∞ (R) at 0 is given by the sets:
X
U(n ,kn ) := {f ∈ C ∞ (R) | supp(f ) ⊆ [−n, n] and |f (kn ) | < n },
n∈N
Exercise 2.6.2. Show that a sequence (fn )∞ n=1 in Cc (R) converges to f ∈ Cc (R)
∞ ∞
with respect to the topology defined above if and only if it converges as was defined
in the first lecture (Definition 1.2.2), i.e.,
S
∞
(1) There exists a compact set K ⊆ R s.t. supp(fn ) ⊆ K.
n=1
(k)
(2) For every k ∈ N the derivatives fn (x) converge uniformly to f (k) (x).
Remark 2.6.3. Notice that the topology on Cc∞ (R) is complicated- it is a direct
limit of an inverse limit of Banach spaces!
Exercise 2.6.4. Show that taking a convex hull instead of a Minkowski sum (i.e.,
defining U(n ,kn ) := convn∈N {f ∈ C ∞ (R) | supp(f ) ⊆ [−n, n], f (kn ) < n }) will
result in the same topology. This shows that Cc∞ (R) is a locally convex topological
vector space (Note that this follows since this is a direct limit of Fréchet spaces).
Definition 2.7.2.
(1) Let V be a topological vector space. A subset B ⊆ V is called bounded if for
every open U ⊆ V there exists λ ∈ R such that B ⊆ λ ∙ U .
(2) Denote V ∗ = {f : V → R : f is linear and continuous}. There are many
topologies one can define on V ∗ , we mention here two of those. Let > 0
and S ⊆ V , and set U,S = {f ∈ V ∗ : ∀x ∈ S, |f (x)| < }.
(a) The weak topology on V ∗ , denoted Vw∗ . The basis for the topology on
Vw∗ at 0 is given by:
(b) The strong topology on V ∗ , denoted VS∗ . The basis for the topology on
VS∗ at 0 is given by:
U,S ∈ B there exists N ∈ N s.t. (fn −f ) ∈ U,S for n > N . That is, ∀x ∈ S we have
that |fn (x) − f (x)| < . Therefore (fn )∞
n=1 converges to f w.r.t the weak topology
⇐⇒ it converges point-wise, and it converges to f w.r.t the strong topology ⇐⇒
it converges uniformly on every bounded set.
Example 2.7.5 (Fleeing bump function). Let V = R and let ψ be a bump function.
Notice that gn (x) = ψ(x + n) converges pointwise to 0 (and hence also in the weak
GENERALIZED FUNCTIONS LECTURES 24
topology). Note that gn does not converge uniformly to 0, but it does converge
uniformly on bounded sets to 0, so it strongly converges to 0.
Exercise 2.7.6. Let S ⊆ Cc∞ (R) be a bounded set, then there exists a compact
K ⊂ R such that S ⊆ CK
∞
(R).
Remark 3.1.2.
(1) For an open U ⊂ Rn , the topology we defined on Cc∞ (U ) is generally not the
same as the induced topology when considering it as a subspace of Cc∞ (Rn ).
(2) For every compact K ⊂ U , we have CK ∞
(U ) ⊂ Cc∞ (U ). Here the topology
on CK∞
(U ) is indeed the induced topology from Cc∞ (U ).
Next we prove that with respect to the restriction operation for distributions defined
above, the space of distributions is equipped with a natural structure of a sheaf.
Lemma 3.1.3 (Locally finite partition of unity). Let I be an indexing set and
S
U = Ui be a union of open sets. Then there exist functions λi ∈ Cc∞ (U ) such
i∈I
that:
(i) supp(λi ) ⊂ Ui
(ii) For every x ∈ U , there exists an open neighborhood Ux of x in U and a finite
set Sx of indices such that λi |Ux ≡ 0 for all i ∈
/ S.
GENERALIZED FUNCTIONS LECTURES 25
P
(iii) For every x ∈ U , i∈I λi (x) = 1.
Note that the sum in the denominator is finite. Now, for every i define
X
λi (x) = fj (x).
j∈α−1 (i)
Theorem 3.1.4. With respect to the restriction map defined above, distributions
S
form a sheaf, that is given an open U ⊆ Rn , and open cover U = Ui , we have:
i∈I
then ξ|U ≡ 0.
(2) (Glueability axiom) Given a collection of distributions {ξi }i∈I , where ξi ∈
C −∞ (Ui ), that agree on intersections, i.e. ∀i, j ∈ I we have that ξi |Ui ∩Uj ≡
ξj |Ui ∩Uj , there exists ξ ∈ C −∞ (U ) satisfying ξ|Ui ≡ ξi for any i.
P
Proof. Choose a locally finite partition of unity 1 = i∈I λi corresponding to the
cover Ui by Lemma 3.1.3.
(1) Given f ∈ Cc∞ (U ) we need to show hξ, f i = 0. Let fi := λi f . Then
Pn
f= fi , and
i=1
n
X n
X
hξ, f i = hξ, fi i = hξ, fi i = 0.
i=1 i=1
(2) Note that for any compact K ⊆ U we then have that λi |K ≡ 0 for all but
finitely many i. Now suppose we are given ξi ∈ C −∞ (Ui ) which agree on
pairwise intersections. For any f ∈ Cc∞ (U ) define
X
hξ, f i := hξi , λi f i.
i∈I
where the second equality follows from the fact that λi f ∈ Cc∞ (Uj ∩ Ui )
and ξi |Ui ∩Uj ≡ ξj |Ui ∩Uj .
A second way to prove continuity of ξ is working with the open sets
in the topology of Cc∞ (U ). As ξi are continuous, they are bounded in
some convex open set 0 ∈ Bi , so |hξi , f i| < for every f ∈ Bi . Notice
S L
that conv( Bi ) is open in i∈I Cc∞ (Ui ), where each Bi is an open set in
i∈I L S
Cc∞ (Ui ) and hence a set in i∈I Cc∞ (Ui )) as conv( Bi ) ∩ Cc∞ (Ui ) = Bi .
L i∈I
Consider the map ϕ : i∈I Cc∞ (Ui ) → Cc∞ (U ) given by extension by zero
S
and summation. Note that B := ϕ(conv( Bi )) is open. Now let f ∈ B.
i∈I
P
n P
We can write f = ai fi where fi ∈ Bji and ai = 1. Therefore
P jP
i =1
ξ(f ) := ξi (ai fi ) < ai ∙ = and ξ is bounded on B.
CR−∞
k (R ) = {ξ ∈ C
n −∞
(Rn )|∀f ∈ Cc∞ (Rn \Rk ) we have hξ, f i = 0}
{ξ ∈ C −∞ (Rn )|∀f ∈ Cc∞ (Rn \Rk ) we have hξ, f i = 0} = {ξ ∈ C −∞ (Rn )| ξ|V = 0},
GENERALIZED FUNCTIONS LECTURES 27
Fm (CR−∞ n ⊥ −∞ n −∞
k (R )) = Vm = {ξ ∈ CRk (R ) : ξ|Vm = 0} ⊆ CRk (R ).
n
locally.
(4) Show that Fn is stable under coordinate changes. More generally, let ϕ :
Rn → Rn be a smooth proper map that fixes Rk . Show that for every ξ ∈ Fi ,
ϕ∗ (ξ) ∈ Fi , where hϕ∗ (ξ), f i := hξ, f ◦ ϕi.
L ∂ i (C −∞ (Rk ))
Theorem 3.2.3. As vector spaces we have Fm ' ∂xi where i ∈ Nn−k
0
|i|≤m
∂ i (C −∞ (Rk )) ∂i
and ∂xi is the image of C −∞
(R ) under the differential operator
k
∂xi (note
that we only differentiate with respect to coordinates not lying in Rk ).
Proof. We prove here the statement for m = 0, and return to the case where m > 0
in section 5. Define a map res∗ : C −∞ (Rk ) → F0 by hres∗ ξ, f i = hξ, f |Rk i for every
ξ ∈ C −∞ (Rk ). Notice that res∗ ξ(f ) = 0 for any f ∈ F0 by definition so it is well
defined.
Furthermore, res∗ is injective since if hres∗ ξ, f i = hξ, f |Rk i = 0 for all f ∈ Cc∞ (Rn )
then ξ = 0 since the restriction res : Cc∞ (Rn ) → Cc∞ (Rk ) is surjective.
It is left to prove surjectivity. Define an extension map ext∗ : F0 → C −∞ (Rk )
by hext∗ η, f i = hη, ext(f )i where ext(f )|Rk = f for every f ∈ Cc∞ (Rk ). Note
that this is well defined since if we choose a different extension ext0 (f ) we get
that hη, ext0 (f ) − ext(f )i = 0 since (ext0 (f ) − ext(f ))|Rk ≡ 0 and thus ext∗ (η) =
ext0∗ (η). Also, we have that ext∗ η is a continuous functional, since we can choose
the extension ext(f ) in such a way that if (fn )∞ n=1 converges to f then (ext(fn ))n=1
∞
converges to ext(f ). Finally, note that since res ext (η) = η, we have that ext∗ is
∗ ∗
i −∞ k
Exercise 3.2.5. Show that Gm and G(i) = ∂ (C ∂xi(R )) , where i is some multi-
index, are not invariant under changes of coordinates, that is we might have that
ϕ(Gm ) 6= Gm and ϕ(G(i) ) 6= G(i) for a diffeomorphism ϕ : Rn → Rn .
ϕ : Cc∞ (Rn ) ⊗ Cc∞ (Rk ) → Cc∞ (Rn × Rk ), given by ϕ(f ⊗ g)(x, y) 7→ f (x)g(y).
Exercise 3.3.1. Show that this map is continuous and has a dense image.
Exercise 3.3.2. Show that this map is continuous and has a dense image.
Let us now denote by L(C −∞ (Rn ), C −∞ (Rk )) the space of all continuous linear
operators, and define a natural map
Remark 3.3.4. (i) Note that the map S is similar to the matrix multiplication.
(ii) The map S is in fact an isomorphism. This statement is the Schwartz kernel
theorem, see [Tre67, Theorem 51.7]
(iii) There are two natural topologies one can define on a tensor product: the injec-
tive one and the projective one. If the spaces are nuclear these two topologies
coincide. We will not define these notions in the present course, but all the
topological vector spaces we consider are nuclear, and thus our tensor products
possess natural topology. If we complete C −∞ (Rn ) ⊗ C −∞ (Rk ) with respect
to this topology, the map Φ will extend, and will become an isomorphism.
The analogous statement for the map ϕ does not hold, but it will hold if we
omit the compact support assumption. In other words, the extension of ϕ to
the completed tensor product C ∞ (Rn )⊗C ˆ ∞ (Rk ) by the same formula is an
isomorphism with C (R × R ), see [Tre67, Theorem 51.6].
∞ n k
GENERALIZED FUNCTIONS LECTURES 29
One motivation to define the p-adic numbers comes from number theory. Assume
we are given a polynomial equation p(x) = 0 where p ∈ Z[x]. If it has an integral
solution x0 ∈ Z, then surely it satisfies the equation p(x) = 0 mod n for every
n ∈ N. Now, consider the converse question - if we know that it has a solution
modulo n for every n ∈ N, does it have an integral solution in characteristic zero?
In some cases, such as for quadratic forms, the answer, together with demanding
that there also exists a real solution, is yes (see the Hasse principle for more on
this). To know whether there exists a real solution, we can use simple methods
from analysis. The question of whether an equation has a solution mod n for every
n ∈ N can be simplified in two steps. Firstly, by the Chinese remainder theorem
it is enough to check whether the equation has a solution mod pn for every n ∈ N.
The second step is then by defining a ring Zp , called the ring of p-adic integers,
such that if there exists a solution x ∈ Zp it implies that there is a solution mod
pn for every n ∈ N. The field Q of p-adic numbers is then defined to be the field of
fractions of Zp .
A different motivation for introducing the p-adic numbers comes from a more an-
alytic point of view. One construction of the real numbers is via completing Q
with respect to its absolute value. An interesting question is whether this can be
generalized, that is what are the possible absolute value-like functions on Q and
their completions. It turns out that besides the standard and the trivial absolute
values, every absolute value (up to equivalence) is a p-adic absolute value (this is
essentially Theorem 4.1.8 below). The p-adic numbers are then obtained as the
completion of Q with respect to such an absolute value.
In this manuscript we take the second approach, starting with defining what prop-
erties we demand from an absolute value function.
For topological fields we demand the absolute value to be a continuous map. Notice
that every absolute value satisfies |1| = 1 (as |1| = |1| ∙ |1|, and |1| 6= 0).
Definition 4.1.4. Let p be a prime number. We define the p-adic absolute value
of x ∈ Q by
p−n , for x 6= 0,
|x|p =
0, for x = 0,
where x = pn ab and a, b ∈ Z, are coprime to p.
Definition 4.1.6. Two absolute values | ∙ | and | ∙ |0 on F are called equivalent, and
denoted | ∙ | ∼ | ∙ |0 if they induce the same topology on F .
Exercise 4.1.7. Let | ∙ | and | ∙ |0 be two absolute values on a field F . Show that
the following are equivalent:
(1) | ∙ | and | ∙ |0 are equivalent.
(2) There exists α ∈ R>0 such that | ∙ | = (| ∙ |0 )α .
(3) Every sequence which is Cauchy with respect to | ∙ | is Cauchy with respect
to | ∙ |0 .
(1) |x + . . . + x| ≤ |x|,
|p|
pZ ⊆ a ( Z and consequentially pZ = a. Now, if we put s = − log log p , we see that
|p| = p−s . Taking x = pn ab where a, b, n ∈ Z and a and b are coprime to p we get:
a a
|x| = |pn | = |pn | | | = |p|n = p−ns = |x|sp ,
b b
|{z}
=1
showing | ∙ | is equivalent to | ∙ |p .
Now, assume | ∙ | is an Archimedean absolute value. We must have that |n| ≥ 1 for
all non-zero integers n ∈ Z. Otherwise, let n be the smallest positive number such
that |n| < 1, and for every n < x ∈ N write it in base n:
(2) x = a0 + a1 n + a2 n2 + . . . + ar nr , for 0 ≤ ai ≤ n − 1, nr ≤ x.
Xr
log x log x
|x| ≤ |ai ||n| ≤ 1 +
i
n|n| log n .
i=0
log n
|x| = xs = |x|s∞ ,
Exercise 4.1.9. Show that given a field F and an absolute value | ∙ | on it the
topology it defines makes F a topological field, i.e. that addition, multiplication,
and the inverse operations are continuous.
Definition 4.1.11. Let p be a prime number. We define the field of p-adic numbers
Qp to be the completion of Q with respect to the absolute value | ∙ |p .
Remark 4.1.12.
(1) The completion is defined just as we did in the case of the Archimedean
norm on Q; by equivalence classes of Cauchy sequences. Therefore, any
element a ∈ Qp is represented by a Cauchy sequence {an }∞ n=1 ⊂ Q with
respect to | ∙ |p .
(2) We get a space which is an uncountable field of characteristic 0, not alge-
braically closed, locally compact (every point has a compact neighborhood)
and totally disconnected, i.e. every connected component is a point.
Exercise 4.1.14. Show that the p-adic absolute value extends from Q to Qp , that is
n=1 of elements in Q the limit lim |an |p
show that for every Cauchy sequence {an }∞
n→∞
exists.
Remark 4.1.15. Notice that just like R, this completion is not algebraically closed.
Try to find an equation in Qp for some p which does not have a solution Qp .
4.2. Misc. -not sure what to do with them (add to an appendix about
p-adic numbers?)
Theorem 4.2.1 (Taken from Koeblitz Theorem 2 page 11). Every equivalence
class a ∈ Qp for which |a|p ≤ 1 has exactly one representative Cauchy sequence of
the form {ai }∞
i=1 for which:
1) 0 ≤ ai < p for i = 1, 2, . . .
i
Lemma 4.2.2 (Taken from Koeblitz page 12). If x ∈ Q and kxkp ≤ 1, then for
any i there exists an integer α ∈ Z such that kα − xkp ≤ p−i . The integer α can be
chosen in the set {0, 1, 2, . . . pi − 1}.
Proof. Let x = a/b written in the form where (gcd(a, b) = 1). Since kxkp ≤ 1 it
follows that p does not divide b and therefore b and pi are relatively prime. Then
we can find m, n ∈ Z such that bm + npi = 1 . The intuition is that bm is close to
1 up to a small p-adic length so it is a good approximation to 1 so am is a good
approximation to a/b. So we pick α = am and get:
So {ai } v {bi }.
Now, if we have some {a} ∈ Qp with kak ≥ 1 then there exists some m such that
ka ∙ pm k ≤ 1 and we have numbers with negative powers. Therefore we can present
GENERALIZED FUNCTIONS LECTURES 34
4.5. Haar measure and local fields. Let X be a topological space and let Cc (X)
be the space of continuous functions of compact support on X. Recall that the space
of continuous linear functionals Cc (X)∗ can be identified with the space of regular
Borel measures on X.
Exercise 4.5.2.
(1) Prove Haar’s theorem for (Qp , +).
(2) Given a Haar measure μ, we can define another invariant measure μa (B) =
μ(aB) for any a ∈ Qp . Show that μa = |a| ∙ μ.
Theorem 4.5.4. Any local field F is isomorphic as a topological field to one of the
following :
(1) R or C (if F is Archimedean).
(2) A finite extension of Qp for some prime p (if F is non-Archimedean of
characteristic 0).
P∞
(3) The field of formal Laurent series Fps ((t)) = { i=−k ai ti | ai ∈ Fps } for
some prime p and natural number s (if F is non-Archimedean of charac-
teristic p).
GENERALIZED FUNCTIONS LECTURES 36
Remark 4.6.2. We usually add the demand X is σ-compact, that is it is the union
of countably many compact spaces. Such a space is also sometimes called countable
at ∞.
Exercise 4.6.3. Find a compact `-space X and U ⊆ X such that U is not countable
at ∞.
Now, we claim that for every closed subset F of K such that F ∩ P = ∅ there
exists some W ∈ Px such that W ∩ F = ∅. Indeed, set η = {U ∩ F : U ∈ Px }. By
assumption, it is a family of non-empty closed subsets of F , and since F is compact if
T Tn T
n
V = ∅, then there is a finite collection of Vi such that Vi = Ui ∩ F = ∅
V ∈η i=0 i=0
(note that this is an equivalent characterization of compactness via closed sets).
Tn
Now set W := Ui ∈ Px . Since Px is closed under finite intersections, W ∈ Px .
i=0
Now we wish to show that P = {x}. Assume the contrary, i.e. P 6= {x}. P
is disconnected since X is totally disconnected, so there exists non-empty closed
x ∈ A and B such that A ∪ B = P and A ∩ B = ∅ which are open in K. Since K
is regular, (Hausdorff + locally compact implies regular), there exist open disjoint
sets U 3 U and V ⊇ B in K, where we have F = K\(U ∪ V ) closed in K and
P ∩ F = ∅. We showed that for such F we can find W ∈ Px such that F ∩ W = ∅.
Now, observe that the open set G = U ∩ W is also closed in K as,
Exercise 4.6.6. Show that every σ-compact, first countable `-space X is homeo-
morphic to one of the following:
(1) Countable (or finite) discrete space.
(2) Cantor set.
(3) Cantor set minus a point.
(4) Disjoint union of (2) or (3) with (1).
S
Definition 4.6.7. A refinement of an open cover Ui = X is an open cover
i∈I
{Vj }j∈J such that for any j, we have that Vj ⊆ Ui for some i.
Exercise 4.6.8.
(1) Let C ⊆ X be a compact subset of an `-space. Then any open cover has an
open compact disjoint refinement.
(2) Let X be a σ-compact `-space, then any open cover has an open compact
disjoint refinement.
GENERALIZED FUNCTIONS LECTURES 38
Proposition 4.7.2. Let X be an `-space. Show that smooth functions separate the
points in X. Assuming this, the Stone-Weierstrass theorem implies that C ∞ (X) is
dense in the space of all continuous functions C(X).
Definition 4.7.3. The space of smooth functions with compact support, Cc∞ (X) ⊂
C ∞ (X), are called Schwartz functions. We denote them by S(X). We also denote
Dist(X) = Cc∞ (X)∗ = S ∗ (X). We consider S(X) as a vector space, without any
topology.
Remark 4.7.5. In Rn , the Schwartz functions are the functions whose derivatives
decrease faster than any polynomial, and there is a strict containment Cc∞ (Rn ) ⊂
S(X) ⊂ C ∞ (Rn ). We will define them in the next lectures.
Proof. It is clear that extension by zero S(U )→S(X) is injective, we show that
S(X)→S(Z) is onto. Let f ∈ S(Z). As f is locally constant and compactly
supported, we may assume that Z is compact and has a covering by a finite number
of open sets Ui (open in Z) with f |Ui = ci . Notice that each Ui is of the form
S
n
Ui = Wi ∩ Z, where Wi is open in X. Therefore, Z ⊆ Wi , and as Z is compact,
S i=1
m
we may refine {Wi }ni=1 and get that Z ⊆ j=1 Vj where Vj are open, compact
GENERALIZED FUNCTIONS LECTURES 39
the fact that suppf is compact). Since each term cUi ×Vi ∈ Imφ we are done.
To show φ is injective, assume that
k
X k
X
φ( fi ⊗ gi ))(x, y) := fi (x) ∙ gi (y) = 0.
i=1 i=1
We can assume that {fi } are linearly independent and that {gi } are non zero and
that k is minimal with respect to these demands. If we take some y such that
Pk
g1 (y) 6= 0 we get that for any x ∈ X, we have i=1 fi (x) ∙ gi (y) = 0. This implies
that fi are linearly dependent. Contradiction. Hence gi ≡ 0 for all i, implying
fi ⊗ gi ≡ 0, contradicting the assumption that k is minimal.
S : S ∗ (X × Y ) → HomF (S(X), S ∗ (Y )) by
GENERALIZED FUNCTIONS LECTURES 40
Exercise 4.8.7.
(i) Show that the map S is a linear isomorphism.
(ii) Assume X = F n , Y = F m and let μ and ν be Haar measures on X and
Y . Embed C ∞ (X × Y ) ,→ S ∗ (X × Y ) and C ∞ (Y ) ,→ S ∗ (Y ) by multipli-
cation by the corresponding Haar measures. Then S maps C ∞ (X × Y ) to
HomF (S(X), C ∞ (Y )) by the formula
ˆ
(S(f )(g))(y) = f (x, y)g(x)μ.
X
Φ : S ∗ (X) ⊗ S ∗ (Y ) → S ∗ (X × Y ) given by
Exercise 5.0.2. Let V be a topological vector space over F . Prove that Cc∞ (X, V ) ∼
=
Cc (X) ⊗F V as topological vector spaces, where the topology on Cc (X) ⊗F V is
∞ ∞
given by choosing a basis to identify V with F n and by then taking the product topol-
ogy on Cc∞ (X) ⊗F F n ∼ = (Cc∞ (X))n . In particular, this topology is independent of
a choice of a basis.
5.1. Smooth measures. Recall that a measure is a σ additive map from the σ-
algebra of Borel subsets of X into R. For us, the following characterization is
better:
Definition 5.1.1. Let X be a locally compact topological space. The space of signed
measures on X is Cc (X)∗ , i.e. all continuous functionals on Cc (X) (and all linear
functionals if X is an `-space). A signed measure is a measure if it is non-negative
on non-negative functions.
As the space Cc (X) is larger than Cc∞ (X), its dual is smaller. Explicitly, Cc (X)∗ ⊆
Cc∞ (X)∗ where the inclusion is the dual of the dense embedding Cc∞ (X) ,→ Cc (X).
If X is a group then in Cc (X)∗ there is a one-dimensional space of Haar measures,
which for X = Rn is just the space of multiples of the Lebesgue measure.
GENERALIZED FUNCTIONS LECTURES 41
Definition 5.1.3. Let V be a locally compact vector space (note that it must be
finite dimensional as otherwise it is not locally compact). The space of Haar mea-
sures on V , denoted Haar(V ) ⊆ Cc (V )∗ , is the space of translation invariant mea-
sures (which exists by Haar’s theorem).
The fact that this space is one dimensional is non-trivial, but the intuition is as
follows: A Borel measure on V is determined by its value on cubes with rational
coordinates, as they form basis of the topology. It is not hard to see that if the
measure is translation invariant, the measures of these cubes are determined by the
measure of the unit cube.
Exercise 5.1.5 (*). Let V be a vector space over a local field and let ξ ∈ Cc∞ (V )∗
be translation invariant. Prove that ξ is a Haar measure, i.e. show it is a mea-
sure (note that Cc∞ (V )∗ ! Cc (V )∗ so a-priori there might be translation invariant
distributions which are not measures).
is not canonical.
where the diagonals are dual to each other. Both inclusions i and j are obtained
via the pairing h∙, ∙i.
Exercise 5.2.1. Show that Haar(V ) ' Dist(V )V or equivalently that Dist(V )V is
one dimensional, for any finite dimensional vector space V over a local field F .
Definition 5.2.2. We can also define generalized functions with values in a vector
space by either:
1) C −∞ (V, E) := C −∞ (V ) ⊗ E
2) C −∞ (V, E) := Cc∞ (V, Haar(V ) ⊗ E ∗ )∗
and then C −∞ (V, Haar(V )) := C −∞ (V ) ⊗ Haar(V ) = Cc∞ (V )∗ = Dist(V ).
Exercise 5.2.3.
(1) Show that the two definitions of C −∞ (V, E) are equivalent.
(2) Describe an embedding Cc∞ (V, E) ,→ C −∞ (V, E).
Definition 5.3.1. Let V be a finite dimensional vector space over a local field F .
(1) We define the exterior algebra as
∞
M
Λ(V ) = Λk (V ),
k=0
N
k
where Λ0 (V ) = F , and for k > 0 we have Λk (V ) = ( V )/Jk where Jk is
j=1
N
k
the vector space generated in V by the set
j=1
N
k
where Sym0 (V ) = F , and for k > 0 we have Symk (V ) = ( V )/Ik where
j=1
N
k
Ik is the vector space generated in V by the set
j=1
Note that this implies that the elements of the exterior algebra are anti-symmetric
(i.e. v ⊗ u = −u ⊗ v), and that Λk (V ) = 0 if k > dim V , since after choosing a
basis for V and decomposing an element in Λk (V ) to basic tensors, there must be
a basis element which appears at least twice.
Definition 5.3.2. Let V be an n-dimensional vector space over a local field F with
absolute value | ∙ |.
(1) We define the space of k-forms Ωk (V ) = Λk (V ∗ ).
(2) For a 1-dimensional space V we define a real vector space
Definition 5.3.4. For a finite dimensional real vector space V define the orienta-
tion line
Exercise 5.3.5. Using the tensor product of the natural maps Ωtop (V ) → Dens(V )
and Ωn (V ) → Ori(V ) show that Ωtop (V ) = Dens(V ) ⊗ Ori(V ).
Note that the orientation line is a linear space and not just two points as one is
used to think about orientations. However, we have two distinguished points in
Ori(V ), the two functions with absolute value 1. These are the usual orientations
we are used to thinking about.
Proposition 5.3.6. Show that there is a canonical isomorphism Dens(V ) ' Haar(V ).
ϕ ∈ |Ωn (V )| such that ϕ(e1 , . . . , en ) = 1 with the Haar measure normalized such
that it has the value 1 on the parallelogram spanned by the vectors {ei }ni=1 . This is
independent of choice of basis since given a different basis both elements would be
GENERALIZED FUNCTIONS LECTURES 44
multiplied by the same factor of | det(M )|, where M is the change of basis matrix
with respect to these two bases.
Fm (W )/Fm−1 (W ) ∼
=can Cc∞ (W, Symm (W ⊥ ))∗ ' Dist(W ) ⊗ Symm (V /W ).
Vm−1 (W )/Vm (W ) ∼
= Cc∞ (W, Symi (W ⊥ )).
GENERALIZED FUNCTIONS LECTURES 45
For this we do the natural thing- we attach to f ∈ Vm−1 (W )/Vm (W ) its i-th
derivatives. Explicitly, we define:
Exercise 5.4.3.
(1) Finish the proof of the lemma - show that Φ is onto, hence an isomorphism.
(2) Show that the isomorphism Fm (W )/Fm−1 (W ) ∼ =can Cc∞ (W, Symm (W ⊥ ))∗
is invariant with respect to diffeomorphism of (V, W ).
(3) Find ξ ∈ Dist(V \W ) such that there is no η ∈ Dist(V ) such that η|V \W =
ξ. That is, show that the natural map Dist(V ) → Dist(V \W ) is not onto.
To get a similar result for generalized functions, we twist by the one dimensional
space of Haar measures:
Fm (W )/Fm−1 (W ) ∼
= Cc∞ (W, Symm (W ⊥ ))∗ = C −∞ (W, Symm (W ⊥ ) ⊗ Haar(W )).
Take Gm (W ) = Fm (W ) ⊗ Haar(W )∗ ⊆ C −∞ (V ). We get by the compatibility of
tensor product and quotient the following:
Gm (W )/Gm−1 (W ) ∼
= C −∞ (W, Symm (W ⊥ ) ⊗ Haar(W )) ⊗ Haar(V )∗
∼
= C −∞ (W, Symm (W ⊥ ) ⊗ Haar(W ) ⊗ Haar(V )∗ ).
The next exercise shows that Haar(W ) ⊗ Haar(V )∗ can be presented in a simpler
manner:
We arrive at the following corollary, yielding the desired description for generalized
functions.
Gm (W )/Gm−1 (W ) ∼
= C −∞ (W, Symm (W ⊥ ) ⊗ Haar(W ⊥ )).
6. Manifolds
Exercise 6.0.2.
(1) Find a space X which is locally homeomorphic to Rn at every point and is
paracompact but is not Hausdorff.
(2) Find a space which is Hausdorff, locally isomorphic to Rn but is not para-
compact.
We now give a definition of a smooth manifold which is different than the usual
definition in differential topology and uses sheaves of functions.
Definition 6.0.6. A smooth manifold is a space with a sheaf of functions (X, C ∞ (X))
such that X is a topological manifold and for every point x ∈ X there is an open
neighborhood U such that (U, C ∞ (X)|U ) ' (Rn , C ∞ (Rn )) as sheaves of functions.
Remark 6.0.7. The usual definition of manifolds adds an atlas to the structure
S
of X, that is an open cover X = Ui with diffeomorphisms φi : Ui → Rn . We
i∈I
also demand that φi ◦ φ−1
j is differentiable, so it seems like an additional demand
GENERALIZED FUNCTIONS LECTURES 47
with respect to the definition above. Alas, further rumination shows that a pair
of isomorphisms ϕi : (Ui , C ∞ (Ui )) → (Rn , C ∞ (Rn )) and ϕj : (Uj , C ∞ (Uj )) →
(Rn , C ∞ (Rn )) implies that the following composition is an isomorphism:
#
ϕi ◦ ϕ−1
j |Ui ∩Uj : (Rn , C ∞ (Rn ))|ϕj (Ui ∩Uj ) → ϕi ◦ ϕ−1
j ∗
(Rn , C ∞ (Rn ))|ϕi (Ui ∩Uj ) .
Exercise 6.0.8.
(1) Show that C ∞ (Rn , Rk ) = {f : Rn → Rk : f ∗ (μ) ∈ C ∞ (Rn ) ∀μ ∈ C ∞ (Rk )}.
(2) Let M and N be smooth manifolds. Show that 6.0.6(1) is equivalent to the
usual definition of a morphism of smooth manifolds. That is, that a map
f : M → N is a smooth map of manifolds ⇐⇒ it is a morphism of ringed
spaces (where the sheaf is a sheaf of smooth functions).
6.1. Tangent space of a manifold. There are several equivalent ways to define
the tangent space to a smooth manifold M at a point x ∈ M . We first give a
categorical definition and then construct several objects which satisfy this definition.
Definition 6.1.2. A tangent space is a functor Tan : ptMan → Vect from pointed
smooth manifolds to vector spaces which satisfies the following conditions:
(1) The restriction of T an to the subcategory V ect ⊂ ptM an is the identity
functor.
(2) If f, g : (R, 0) → (C, 0) satisfy f 0 (0) = g 0 (0) then Tan(f ) = Tan(g).
(3) If ϕ : U ,→ M is an open embedding, then Tan(ϕ) is an isomorphism.
Exercise 6.1.3. Show the constructions of the tangent space given above are equiv-
alent.
Exercise 6.1.5.
(1) Show the differential is well defined, i.e. it does not depend on the repre-
sentative γ ∈ [γ].
(2) Show that given manifolds M, N, and K and maps φ : M → N and ψ :
N → K, the differentials satisfy dx (ψ ◦ φ) = dφ(x) (ψ) ◦ dx (φ).
Example 6.2.2.
(1) Let φ : [−1, 1] → R2 be a smooth path that slows to a stop at φ(0) = (0, 0),
but spends no time at (0, 0). That is, φ0 (0) = (0, 0), but φ(x) 6= 0 for all
x 6= 0 in some neighborhood [−ε, ε] of 0. Such a φ is locally injective at 0,
but since d0 φ = 0 it is not an immersion at 0.
(2) An immersion is not necessarily one-to-one. As an example, consider a
self-intersecting path φ : R → R2 with constant speed.
(3) Let L and D be finite dimensional linear spaces. The differential of a linear
map φ : L → V is φ itself. Thus, a one-to-one φ will be an immersion, an
onto φ will be a submersion, and an isomorphism of linear space will be an
étale map.
(2) Show that every proper map which is an injective immersion is a closed
embedding.
(3) Show that a proper map which is étale is a covering map, and that a covering
map with finite fibers is proper and étale.
6.3. Analytic manifolds and vector bundles. We would like to be able to talk
about manifolds for a general local field. In order to do so, for a non-archimedean
local field F we introduce the notion of an analytic F -manifold.
Remark 6.3.2. By Exercise 4.6.8 there is no need to use partition of unity for
F -analytic manifolds.
Example 6.3.3. There exist singular analytic manifolds, and any singular affine
algebraic variety is an example for such a manifold.
Exercise 6.3.5. It is known that the Mobius strip M is not homeomorphic to the
(finite) cylinder I × S 1 . By extending each segment I of M to R, we can define a
GENERALIZED FUNCTIONS LECTURES 50
vector bundle E over the manifold S 1 . This way, points of E are such that the fiber
over θ ∈ S 1 is a line in R3 which intersects the z-axis with angle 0.5 ∙ θ. Define the
vector bundle above rigorously and show it is not diffeomorphic to the trivial bundle
S 1 × R.
Definition 6.3.7. Let (M, E) be a k-dimensional real vector bundle and π be its
projection. Given trivializing neighborhoods U and V , and trivializations ϕU : U ×
∼ ∼
Rk − → π −1 (U ) and ϕV : V × Rk − → π −1 (V ), one can consider ϕ−1 V ◦ ϕU : (U ∩
V ) × R → (U ∩ V ) × R . We can then write ϕ−1
k k
V ◦ ϕ U (x, v) = (x, gU,V (v)) where
gU,V ∈ GL(R ). The maps gU,V are called transition functions.
k
Notice that the set of transition functions gU,V , satisfy the cocycle conditions
Proof. First, take a cover {Uα }α∈I of M which is a local trivialization of E (that
is, p−1 (Uα ) ' V × Uα ). Define the total space F (E) over each Uα by F (V ) × Uα ,
where the surjection q will be projecting to M , and glue every two pieces q −1 (Uα )
and q −1 (Uβ ) by setting (v, x) ∼ (gα,β (v), x) for every x ∈ Uα ∩ Uβ and v ∈ V ,
where gα,β = F (ϕ−1 Uβ ϕUα ). Finally, note that for any two elements of the cover
−1
gα,β = gβ,α , and that in order for our construction to be well defined we need to
show the cocycle condition, namely that gβ,γ gα,β = gα,γ when restricted to triple
intersections. This holds since
Note that if we want F (E) to have a smooth structure we need to demand that F
preserves smooth maps.
Example 6.3.9. Let E1 and E2 be two vector bundles over M . The direct sum
E1 ⊕ E2 is defined by applying our construction above to the direct sum functor
L
: Vect2 → Vect.
GENERALIZED FUNCTIONS LECTURES 51
Exercise 6.4.2.
GENERALIZED FUNCTIONS LECTURES 52
(1) Show the every manifold has a Riemannian metric, i.e , an inner product
on tangent spaces
h∙, ∙ip : Tp M × Tp M → R
Remark 6.4.3. We do not always have non-zero top differential forms on a man-
ifold M , and the Mobius strip is an example of a manifold with no non-zero top
differential form. However, we can always find a non-zero density on M . Since
with a density we can define a measure on the manifold, we can define integration
over manifolds.
Exercise 6.5.5. Complete the proof by showing that this is indeed an equivalence
of categories.
Exercise 6.5.6.
(1) Show that covering spaces correspond to locally constant sheaves, and that
a covering space is trivial when it corresponds to a constant sheaf.
(2) Give an example of a locally constant sheaf arising from a covering space
which is not constant.
Definition 7.0.3. We can now define the density bundle over an F -analytic man-
ifold X in two ways:
Def 1 (Leray): Dens(X) := |Ωtop (X)|.
Def 2 (Grothendieck):
Dens(X)(U ) := {μ ∈ Measures(U )|∀ϕ ∈ OFn → U, there exists f ∈ C ∞ (OFn ) such that μ = ϕ∗ (f ∙Haar)}.
Cc∞ (M, E) := {f : M → E : π◦f = IdM and ∃K compact such that f |K C (m) = (m, 0)}.
Recall that Cc∞ (Rn , Rk ) = lim
−→CKm (R , R ) where Km is an increasing sequence of
∞ n k
S
∞
compact sets such that Km = Rn . We define a topology on Cc∞ (M, E) using
n=1
the topology on Cc∞ (Rn , Rk ):
Case 1- The trivial case: M ' Rn and E ' Rn × Rk with the projection to
the first component. Note that continuous sections from Rn to Rn × Rk are just
functions in Cc∞ (Rn , Rk ). Hence we give Cc∞ (M, E) the topology of Cc∞ (Rn , Rk ).
Exercise 7.1.2. Show that the above definition is well defined, i.e. it does not
depend on the isomorphism M ' Rn and E ' Rn × Rk → Rn . In other words,
show that:
(1) Given a diffeomorphism ϕ : Rn → Rn it induces a homeomorphism ϕ∗ :
Cc∞ (Rn , Rn × Rk ) → Cc∞ (Rn , Rn × Rk ) via precomposition.
(2) Given a smooth map ψ ∈ C ∞ (Rn , GLk (R)) we have that ψ∗ : Cc∞ (Rn , Rn ×
Rk ) → Cc∞ (Rn , Rn × Rk ) by ψ∗ (f ) = ψ ◦ f is a homeomorphism.
S
Case 2- General case: We can choose trivializing {Ui }i∈I such that M = Ui
i∈I
' '
where ϕi : Ui → Rn and ψi : E|Ui → Rn × Rk . We have a surjective map
M
ϕ: Cc∞ (Ui , E|Ui ) Cc∞ (M, E)
i∈I
by summation where surjectivity follows from partition of unity. We define the quo-
tient topology on Cc∞ (M, E) according to the map ϕ, that is, a set U ⊆ Cc∞ (M, E)
L ∞
is open if ϕ−1 (U ) is open in Cc (Ui , E|Ui ), where the latter is endowed with the
i∈I
direct sum topology.
Proposition 7.1.3. The topology on Cc∞ (M, E) is well defined. That is, the defi-
nition does not depend on the choice of the cover {Ui }i∈I of M .
Proof. We need to show that given a different cover {Vβ }β∈J of M which locally
trivializes M and E, we get the same topology.
Consider the cover {Wα,β } for Wα,β = Uα ∩ Vβ which refines both covers. We need
to show that for the addition map,
MM +
M
Cc∞ (Wα,β , E|Wα,β ) −
→ Cc∞ (Uα , E|Uα )
α∈I β∈J α∈I
a set in the range is open if and only if its preimage is open, where Wα,β ⊆ Uα ' Rn
and E|Wα,β ' E|Uα ' Rk . In order to show the above, it is enough to handle each
GENERALIZED FUNCTIONS LECTURES 55
L +
case Cc∞ (Wα,β , Rk ) −
→ Cc∞ (Uα , Rk ) ' Cc∞ (Rn , Rk ) separately, since in the
β∈J L
direct sum topology a set is open if all the injections Di ,→ Di are continuous
(we furthermore assume our open sets are convex).
Given a basic open set U(Lm ,m ,Bm ) ⊆ Cc∞ (Uα , Rk ) where Lm are mixed differen-
S
∞
tiations, m ∈ R>0 and Bm are compact sets such that Bm = Rn , it is of the
P m=1
form U(Lm ,m ,Bm ) = VLm ,m ,Bm , where
m∈N
VLm ,m ,Bm = f ∈ C ∞ (Rn , Rk ) : supp(f ) ⊆ Bm , sup ||Lm (f )|| < m .
x∈Rn
P P P
l
Now, take a finite sum fβ ∈ +−1 (U{Lm ,m ,Bm } ) for fβ = f = fmi and
i=0
fmi ∈ VLmi ,mi ,Bmi . Let fβ = prβ (f ) be the projection of f into Cc∞ (Wα,β , Rk ),
mi −sup ||Lmi (fmi )||
and define N = #{β : prβ (f ) 6= 0} and 0mi = N and set 0m = Nm
if m 6= mi for all 0 ≤ i ≤ l. For Bm,β 0
⊆ Wα,β , compact sets which exhaust
Wα,β and such that Bm,β
0
⊆ Bm , the sets U(Lm ,0m ,Bm,β
0 ) are basic open sets in each
Cc (Wα,β , R ), and their direct sum is open in the direct sum. Now, we claim that,
∞ k
M
f∈ fβ + U(Lm ,0m ,Bm,β
0 ) ⊆+
−1
(U(Lm ,m ,Bm ) ).
β:fβ 6=0
P lβ
P
Given g = gβ where gβ ∈ U(Lm ,0m ,Bm,β
0 ) , then gβ = gβ,iβ where gβ,iβ ∈
β:fβ 6=0 iβ =1
V Ln i ,0n 0
,Bn .
β iβ iβ ,β
mi −sup ||Lmi (f )||
Thus if niβ = mi for some i, we have sup ||Lmi (gβ,mi )|| < 0mi = N
x∈Bmi
implying that,
X X X
sup Lmi (fmi ,β + gβ,mi ) ≤ sup Lmi (fmi ,β ) + sup ||Lmi (gβ,mi )||
x∈Bmi x∈Bmi x∈Bmi
β:fβ 6=0 β:fβ 6=0 β:fβ 6=0
X m − sup ||Lm (fm )||
< sup Lmi (fmi ) + i i i
x∈Bmi N
β:fβ 6=0
= mi .
Otherwise, if niβ 6= mi for all i, set n0 = niβ , and using the requirement sup ||Ln0 (gβ,n0 )|| <
x∈Bn0
n 0
N we note that:
X X n0
sup Ln0 (gβ,n0 ) ≤ sup ||Ln0 (gβ,n0 )|| < N = n0 .
x∈Bn0 x∈Bn0 N
β:fβ 6=0 β:fβ 6=0
GENERALIZED FUNCTIONS LECTURES 56
P P l P P lβ
This allows us to conclude that f + g = fmi ,β + gβ,iβ lie in
β:fβ 6=0 i=1 β:fβ 6=0 iβ =1
P
U(Lm ,m ,Bm ) = VLm ,m ,Bm for all such functions g, implying that the ad-
m∈N
dition is continuous. For a less cumbersome approach, note that the embed-
L ∞
dings Cc (Wα,β , Rk ) →
− Cc∞ (Rn , Rk ) are continuous (a cookie for the person
β∈J
who finds a quick proof for this), so it is enough to show that the addition map
L ∞ n k + Lk
Cc (R , R ) − → Cc∞ (Rn , Rk ) ' Cc∞ (Rn ) is continuous. Since the domain
β∈J i=k
has the direct sum topology, it is enough to check this for a finite direct sum, which
follows by the continuity of addition in a topological vector space.
L ∞ +
To show the map is open, it is enough to consider CK,c (Wα,β , Rk ) − → CK ∞
(Rn , Rk ) '
β∈J
L
k
∞
CK (Rn ), for every compact K, and since the domain has the direct sum topol-
j=1
ogy and the basic open sets are finite sums of open sets in each coordinate, it is
L
m
+ L
k
enough to show it for a finite direct sum ∞
CK,c (Wi , Rk ) −
→ CK∞
(Rn ) where
i=1 j=1
S
m
K ⊂ Wi . Now, use partition of unity fi , with Ci = supp(fi ) ⊂ Wi where
i=1
P
m
fi |K ≡ 1 to get an onto map via the composition,
i=1
m
M m
M +
∞
CK∩C i
(Wi , Rk ) ,→ ∞
CK,c (Wi , Rk ) − ∞
→ CK (Rn ).
i=1 i=1
Since this is a continuous surjective map of Fréchet spaces, it must be open, implying
that the addition is open since the embedding is continuous.
We now give a different description of the topology of Cc∞ (Rn ). First observe that
f ∈ C(Rn ) is compactly supported if and only if f g is bounded for any g ∈ C(Rn ).
Now let D ∈ Diff(Rn ) be a differential operator on Cc∞ (Rn ). Define a seminorm
kf kD by sup |D(f )(x)|.
x∈Rn
Exercise 7.1.4. The topology on Cc∞ (Rn ) can be defined by the seminorms k kD .
0
Case 1- the trivial case: Assume that E ' M × Rk and E 0 ' M × Rk . Then
0
Diff(C ∞ (M, E), C ∞ (M, E 0 )) ' Diff(C ∞ (M )k , C ∞ (M )k ) and the latter space is
isomorphic as a vector space to the space of k×k 0 matrices with values in Diff(C ∞ (M )).
Exercise 7.1.6. Show that the definition of the space of differential operators
Diff(C ∞ (M, E), C ∞ (M, E 0 )) does not depend on the isomorphisms E ' M × Rk
0
and E 0 ' M × Rk .
Case 2- the general case: Let A ∈ Hom(C ∞ (M, E), C ∞ (M, E 0 )). Then we say
that A ∈ Diff(C ∞ (M, E), C ∞ (M, E 0 )) if:
(1) For any f1 , f2 ∈ C ∞ (M, E) such that f1 |U = f2 |U , we have Af1 |U = Af2 |U .
(2) If E 0 |U is a trivialization then A|U ∈ Diff(U, E|U , E 0 |U ).
Definition 7.1.7 (Second definition to the topology on Cc∞ (M, E)). For D ∈
Diff(C ∞ (M, E), C ∞ (M, E)) define kf kD = sup |D(f )(x)|. Define the topology on
x∈M
Cc∞ (M, E) via
Cc∞ (M, E) = lim (Cc∞ (M, E)k∙kD ).
←D
Exercise 7.1.8. Given a manifold M and a vector bundle E over it show that the
two definitions of the topology on Cc∞ (M, E) are equivalent (one defined via taking
a cover of M and trivialization of E and the other through differential operators).
Although we do not have a natural injection from Cc∞ (M, E) to Cc∞ (M, E)∗ , we
have a natural injection
as follows: let μ ∈ Cc∞ (M, E ∗ ⊗ Dens(M )) and f ∈ Cc∞ (M, E). Note that
that is, f ⊗ μ(m) = f (m) ⊗ μ(m). Note that we have a natural map
by integrating over M according to the measure defined by the section of the density
bundle. We define ˆ
hi(f ), μi := q(f ⊗ μ).
M
Therefore, the definition of generalized sections indeed generalizes smooth sections.
Proof. Recall that C −∞ (X) = μ∞ c (X) . Given a topological vector space V , for
∗
W ⊆ V the space W is dense with respect to the weak topology if and only if
∗
then η = 0.
Assume X is a smooth manifold. Given a non-zero measure η, there exists some
Rn ' U ⊂ X such that η|U 6= 0, to see this either use the fact that distributions
form a sheaf, or view it as a positive function on Borel sets. Now, since U ' Rn
we must have that η|U = g ∙ μHaar where g ∈ C ∞ (Rn ). Taking some cutoff function
ψ ∈ Cc∞ (Rn ) such that ψ|B1 (0) ≡ 1 and ψ ≥ 0 implies the desired result as hgψ, ηi =
hgψ, g ∙ μHaar i = hg 2 ψ, μHaar i > 0 as this is an integral of a positive function.
For an F -analytic manifold we do the same procedure only this time ψ is the
indicator function of the open unit ball in F n .
Exercise 7.2.3. Let M, N be either smooth or F -analytic manifolds and let E and
I be complex vector bundles over M and N respectively.
(i) Show that the natural map Cc∞ (M, E) ⊗ Cc∞ (N, I) → Cc∞ (M × N, E I) is an
embedding with dense image, and is an isomorphism if M, N are F -analytic
manifolds.
(ii) Show that the natural map
given by
In order to prove the theorem, we would like to define the notion of derivatives of
smooth sections f ∈ Cc∞ (M, E). Alas, the value of the derivative depends on the
chart defined on M , so it is not well defined. Fortunately, the notion of vanishing
of derivatives of certain order is well defined as the following exercise shows:
Exercise 7.2.7. Let f ∈ C ∞ (Rn ) such that f (α) (0) = 0 for every multi-index α
with |α| < k, and ϕ : Rn → Rn a diffeomorphism such that ϕ(0) = 0. Furthermore
let g ∈ C ∞ (Rn ) be a nowhere vanishing function, and set fe(x) = g(x)(f ◦ ϕ−1 (x)).
(1) Show that fe(α) (0) = 0 for every multi-index α with |α| < k.
(2) Show that:
∂k ∂k
fe (0) = f (0)g(0).
∂v1 . . . ∂vk ∂ ((dϕ)v1 ) . . . ∂ ((dϕ)vk )
(3) Find a counter example for part (1) if f (i) (0) 6= 0 for some |i| < k.
Remark 7.2.8. As a consequence of this exercise, given any f ∈ Cc∞ (M, E) whose
first k − 1 derivatives vanish we can define the k-th differential symbol of f denoted
GENERALIZED FUNCTIONS LECTURES 60
dkx f : Tx M × . . . × Tx M → Ex by
∂k
dkx f (ξ1,i , . . . , ξk,i ) = (f ◦ ϕ−1
i ) (0),
∂ξ1,i . . . ∂ξk,i
0
where ϕi is a local chart and ξ1,i = (ϕi ◦ γ1 ) (0) are tangent vectors. If we choose
a different chart ϕj we get that
∂k ∂k
dkx f (ξ1,j , . . . , ξk,j ) = (f ◦ ϕ−1
j ) (0) = (f ◦ ϕ −1
i ◦ ϕ) (0)
∂ξ1,j . . . ∂ξk,j ∂ξ1,j . . . ∂ξk,j
where ϕ := ϕi ◦ ϕ−1 j . By the discussion above, we get that
∂k ∂k
(f ◦ ϕ−1
j ) (0) = (f ◦ ϕ −1
i ) (0),
∂ξ1,j . . . ∂ξk,j ∂(dϕ)ξ1,j . . . ∂(dϕ)ξk,j
but as
0
dx ϕ(ξ1,j ) = dx ϕ ∙ (ϕi ◦ γ1 ) (0) = (ϕ ◦ ϕi ◦ γ1 )0 (0) = ξ1,i
we have that dkx f (ξ1,j , . . . , ξk,j ) = dkx f (ξ1,i , . . . , ξk,i ) so this is well defined.
Proof of Theorem 7.2.6. Note that we can identify dkx f ∈ Symk (Tx∗ M ) ⊗ Ex . Let
N ⊆ M be a submanifold. Define:
FNi (Cc∞ (M, E)) = {f ∈ Cc∞ (M, E) : ∀x ∈ N, the first k − 1 derivatives of f vanish}.
i−1
Choose trivializations M |U ' Rn and N |U ∩N ' Rk . We showed that FW (V )/FW
i
(V ) ∼
=
Cc (W, Sym (W ) ⊗ Ex ) using the map f 7→ dx f . Hence we get that:
∞ i ⊥ k
Proof. We have
In this section we assume X and Y are either `-spaces, analytic F -manifolds (with
or without complex bundles over them), or smooth manifolds.
GENERALIZED FUNCTIONS LECTURES 61
Note that if ϕ is not proper then we can define ϕ∗ : Dist(X)prop → Dist(Y ) where
Dist(X)prop := {ξ ∈ Dist(X)| ϕ|supp(ξ) is proper}. We would like to set hϕ∗ ξ, f i =
hξ, f ◦ϕi, but f ◦ϕ might not be compactly supported. Therefore we choose a cutoff
function ρ such that ρ|supp(ξ) = 1 and ρ|U C = 0 where U is a small neighborhood
of supp(ξ) and ϕ|U is proper (it is a hard task to find such a function). Hence we
can define
hϕ∗ ξ, f i := hξ, ρ ∙ (f ◦ ϕ)i.
Note that
Since ϕ|supp(ρ) is proper, and f is compactly supported, this is well defined. The
definition clearly does not depend on the choice of ρ.
Recall that for vector spaces we had that Dens(V ) ' Haar(V ) canonically. Hence
we can identify the space of smooth measures μ∞ c (X) with the space of smooth
sections of the density bundle Cc (X, Dens(X)). Note that we can define ϕ∗ :
∞
(2) ϕ∗ (f ∙ |ωX |) = g ∙ |ωY |, where |ωX | and |ωY | are non-vanishing densities on
X and Y respectively and
|ωX |
ˆ
g(y) = f ∗
ϕ−1 (y) |ϕ ωY |
|ωX |
where |ϕ∗ ωY | ⊗ |ωY | is the image of |ωX | under the natural isomorphism
Proof.
(1) We prove the first statement in two steps. Case 1: X = F n , Y = F m
where n ≥ m and ϕ : F n → F m is the natural projection ϕ(x1 , . . . , xn ) =
x1 , . . . , xm . Recall that Haar(X) ' Haar(Y ) ⊗ Haar(X/Y ) or equivalently
GENERALIZED FUNCTIONS LECTURES 62
that
Dens(X) ' Dens(Y ) ⊗ Dens(X/Y ).
Let φ ∈ Cc∞ (X, Dens(X))and note that φ = f ∙ dμX where f ∈ Cc∞ (X)
and μX is a normalized Haar measure (taking the value 1 on the unit ball
of X = F n ), so we can write μX = μY ⊗ μX/Y . By definition, for any
g ∈ Cc∞ (Y ) we have:
ˆ ˆ ˆ
hϕ∗ (φ), gi = hφ, g ◦ ϕi = f ∙ (g ◦ ϕ)dμX = f ∙ (g ◦ ϕ)dμY ⊗ μX/Y .
X Y X/Y
Since this is an exact sequence of vector spaces it splits so Tx (X) = Tx ϕ−1 (y)⊕
Tϕ(x) (Y ) and by dualizing we get that
In (1) we showed the case where ϕ is a projection. Using the fact that
for diffeomorphisms pushforward and pullback are inverse to one another
and using part (1) we get that:
Definition 8.0.4. By the proposition, the map ϕ∗ : Cc∞ (X, Dens(X)) → Cc∞ (Y, Dens(Y ))
gives rise to a pullback ϕ∗ : C −∞ (Y ) → C −∞ (X).
Proof. As in the proof of the last proposition, we may reduce to the case where
ϕ : X → Y is the natural projection, and X = F n , Y = F m , and E ' F m × F k
is trivial (resp OFn ,OFn and OFm × OFk for F -analytic manifolds). As a consequence,
ϕ∗ (E) = F n × F k (resp. OFn × OFk ). Note reducing is possible since the notion of
smoothness of a distribution (that is, it is a smooth measure) is local.
Let φ = f μ ∈ Cc∞ (X, ϕ∗ (E) ⊗ Dens(X)). Then we have for any g ∈ C ∞ (Y, E),
ˆ ˆ ˆ !
hϕ∗ (φ), gi = hφ, g ◦ ϕi = f ∙ (g ◦ ϕ)μX = f ∙ μX/Y ∙ gμY
X Y X/Y
ˆ
= fe ∙ gμY = hf˜μY , gi,
Y
9. Fourier transform
Definition 9.0.1. Let G be a locally compact Hausdorff abelian group. Define its
Pontryagin dual by,
The topology on G∨ is the compact open topology, i.e. a sub-basis of the topology is
comprised of sets M (K, V ) = {χ ∈ G∨ : χ(K) ⊆ V } where K ⊆ G is compact and
V ⊆ S 1 is open.
χ(x) ∈ [−n , n ] and x2 ∈ Un we get χ(x2 ) = χ(x)2 ∈ [−2n , 2n ] ⊆ [− 2n , 2n ],
implying that V (Un , n ) ⊆ V (Un+1 , n+1 ).
Now, take χ ∈ V (U, ) and a basic open set (−δ, δ) ⊂ S 1 for δ > 0. We have
that [−n , n ] ⊆ (−δ, δ) for n big enough, implying that e ∈ Un ⊆ χ−1 ((−δ, δ))
which means that χ is continuous at e. Since χ is a homomorphism, we can show
it is continuous everywhere; if χ(g) ∈ W ⊂ S 1 and W is open, we have that
(−δ, δ) ⊆ χ(g −1 )W for some δ > 0 and that,
Now, for some m ∈ N0 big enough, the following implies that g ∈ gUm ⊆ χ−1 (W ):
We know that V (U, ) is compact in the product topology, and want to show it is
compact with respect to the compact open topology. For this, it is enough to show
that any net in V (U, ) has a converging subnet in the compact open topology.
Assume we are given some net (xα ) ∈ V (U, ), then it has a subnet (fβ ) → f
converging in the product topology with fβ , f ∈ V (U, ). Now, note that V (U, )
is uniformly equicontinuous, that is if g1 , g2 ∈ G and g1 g2−1 ∈ Un then for any
GENERALIZED FUNCTIONS LECTURES 66
χ ∈ V (U, ),
Given a basic open neighborhood of the identity character 1G ∈ M (K, B0 (0)),
where K is compact, for every g ∈ K we have that g ∈ Un g (for n big enough).
Now, taking any g 0 ∈ Un g, we get that g 0 g −1 ∈ Un implying that for some big
enough β we have that |f (g) − fβ (g)| < n and that,
Exercise 9.0.3. Let G be a locally compact, Hausdorff abelian group. Show that if
G is compact then G∨ is discrete, and that if G is discrete then G∨ is compact.
Theorem 9.0.4. For a locally compact abelian group G, we have that the natural
map ϕ : G → G∨∨ defined by g 7→ ϕg , where ϕg (χ) = χ(g), is an isomorphism
G∨∨ ' G.
Example 9.0.6.
(1) For any finite abelian group G we have that G ' G∨ .
(2) The dual of U1 (C) = S 1 is Z.
(3) We have that R∨ ' R.
Exercise 9.0.7. Let V be a topological vector space over a local field F . Then
V ∗ ⊗F F ∨ ' V ∨ .
Definition 9.0.8. Let G be a locally compact Hausdorff abelian group. The map
F : μc (G) → C(G∨ ) defined by F (μ)(χ) = χdμ is called the Fourier transform.
´
(1) F is continuous.
(2) Let G be a locally compact abelian group. For a character τ : G → S 1 define
shh (τ )(x) = τ (x + h). Show that for η ∈ μ∞c (G) and g ∈ G:
(a) F (shg (η))(χ) = χ(g)F (η)(χ) for all χ ∈ G∨ .
(b) F (χη) = shχ−1 (F (η)) for all χ ∈ G∨ .
Definition 9.0.10. Let X1 and X2 be locally compact topological vector spaces and
c (X1 ) and μ2 ∈ μc (X2 ). We define the external tensor product of such
let μ1 ∈ μ∞ ∞
Definition 9.0.12. Let V be a finite dimensional vector space over a local field F .
Define the space of Schwartz functions S(V ) on V by:
(1) If F is non-archimedean, then S(V ) = Cc∞ (V ), i.e. locally constant func-
tions on V .
(2) If F is archimedean, then
Proof. Assume V is a real vector space of dimension n, and recall that the topology
β
on S(V ) is determined by the semi-norms kf kα,β = sup |Φα (x) ∂ ∂x
f (x)
β | where α, β ∈
x∈V
Q
n
α
Nn0 and Φα (x) = xj j . It is enough to show that for every f ∈ Cc∞ (V, Haar(V ))
j=1
and semi-norm k ∙ kα,β on S(V ∨ ) there exists a semi-norm k ∙ k0 on S(V, Haar(V ))
and a positive constant C such that kF (f )kα,β ≤ Ckf k0 . Now, recall that,
i∂F (f ) i∂ −iξ∙x
ˆ
= (e f (x))dx = F (xj f ),
∂ξj ∂ξj
Rn
where one can differentiate directly using the definition to verify the above proce-
dure. The other side of the coin is given by integration by parts,
−iξ∙x ∞ ξj −iξ∙x ∂f (x) −i∂(f )
ˆ ˆ
ξj F (f ) = ξj e −iξ∙x
f (x)dx = −e f (x) −∞ − e dx = F ( ).
−iξj ∂xj ∂xj
Rn Rn
|ξ| → ∞ we get that Schwartz measures are mapped into S(V ∨ ). We can now
GENERALIZED FUNCTIONS LECTURES 68
where C = 1
Since the last expression is a linear combination of
´
(1+|x|)n+1 dμ(x).
V
norms of the form kf kα0 ,β 0 for |α0 | ≤ |α| + n + 1 and |β 0 | ≤ |β|, this implies that F
is continuous. Note that we can also use this to show that F (f ) is Schwartz, since
β
if all the norms k ∙ kα,β are bounded then the value of |Φα (x) ∂ ∂x f (x)
β | decays to 0 as
Proof. Let f (x) ∙ dx ∈ S(V, Haar(V )) and g(χ) ∙ dχ ∈ S(V ∨ , Haar(V ∨ )). Then by
definition,
ˆ
hF ∗ (f (x) ∙ dx), g(χ) ∙ dχi := hf (x) ∙ dx, F (g(χ)dχ)i = f (x)F (g(χ) ∙ dχ)(x)dx
V
Remark 9.0.16. We will usually omit the ∗ from the F ∗ notations, this should
cause no confusion.
In the following argument we would like to show the Fourier transform is a unitary
operator. For this we will first need to define a pairing between Haar(V ) and
Haar(V ∨ ). Given α ∈ Haar(V ) and β ∈ Haar(V ∨ ) we can define such a pairing
GENERALIZED FUNCTIONS LECTURES 69
as follows. We choose f ∈ Cc∞ (V ∨ ) such that f (0) = 1 and then define hα, βi :=
hF (α), f ∙ βi.
Exercise 9.0.17.
(1) Show this is well defined. That is, given a different g ∈ Cc∞ (V ∨ ) such that
g(0) = 1, show that hF (α), (f − g) ∙ βi = 0.
(2) Show that Haar(V ∨ ) 'can Haar(V )∗ .
Proposition 9.0.20. We have that F1 ◦F0 = flip where hflip(ξ), f (x)μi = hξ, f (−x)μi.
Proof. Note that span{δx }x∈V is a dense subspace of S ∗ (V ) with respect to the
weak topology. Hence it is enough to show that F1 ◦ F0 (δa ) = δ−a for all a ∈ V .
Note that hF0 (δ0 ), f βi := hδ0 , F0 (f β)i = V ∨ f dβ, this implies F0 (δ0 ) = 1.
´
so F1 ◦ F0 (δ0 ) = δ0 .
Using Exercise 9.0.9, we now see that (here χ(a) is the function which substitutes
the value a in a given character χ):
Exercise 9.0.24. Let V be a real 1-dimensional vector space with a positive struc-
ture.
(1) Show that:
(a) V 'can |V |.
(b) V α+β 'can V α ⊗ V β where α, β ∈ Q× .
(2) Deduce that Haar(V )α ⊗ Haar(V )β ' Haar(V )α+β .
In particular, choosing α = 1
2 we have:
1 1
F 12 : S ∗ (V, Haar(V ) 2 ) → S ∗ (V ∨ , Haar(V ∨ ) 2 ).
i∗ i∗
S(V ) S(W ) S ∗ (V ) S ∗ (W )
F F F F
∗
p∗ p
S(V ∨ , Haar(V ∨ )) S(W ∨ , Haar(W ∨ )) G(V ∨ ) G(W ∨ ).
Note that this is possible since p is a submersion (linear and surjective) so pushing
Schwartz measures along it yields Schwartz measures.
Proof. We start by showing the right hand side diagram commutes. Since i∗ , the
Fourier transform and p∗ are continuous with respect to the weak topology, it is
enough to prove commutativity for a dense set in S ∗ (W ).
First take the delta function δ0 ∈ S ∗ (W ), it is a compactly supported measure, and
it holds that i∗ (δ0 ) = δ0 . Furthermore, since F : S ∗ (V ) → G(V ∨ ) is defined via
GENERALIZED FUNCTIONS LECTURES 71
where the third equality is sensible since F (f μ) ∈ S(V ∨∨ ) and 0V ∨∨ (χ) = 1 for
all χ ∈ V ∨ . We can also show that p∗ (1) = 1. Consider G(W ∨ ) as a subspace of
C −∞ (W ∨ ), there the generalized Schwartz function 1 is a smooth function, and
note that the following diagram, where the horizontal arrows are the inclusions is
commutative:
G(V ∨ ) C −∞ (V ∨ ) C ∞ (V ∨ )
p∗ p∗ p∗
G(W ∨ ) C −∞ (W ∨ ) C ∞ (W ∨ ).
hp∗C −∞ (1), f μi = h1, p∗ (f μ)i = hp∗ (f μ), 1i = hf μ, p∗C ∞ (1)i = hf μ, 1i = h1, f μi.
GENERALIZED FUNCTIONS LECTURES 72
The wave-front set of a generalized function ξ is the collection of all points and
codirections in which ξ is not smooth. It is a very important invariant of the
generalized function. For example, there are some operations on functions, like
product or pull-back, that do not extend to arbitrary generalized functions but
do extend to generalized functions under some conditions on the wave-front set.
The term comes from physics. Every differential equation satisfied by a generalized
function will give a restriction on its wave-front set.
Example 10.0.1. Let ξ be the δ-function of the x-axis on R2 , i.e. the generalized
function given by hξ, f dxdyi := f (x)dxi, let L ⊂ R2 denote the x-axis and L⊥ ⊂
´
Exercise 10.0.4 (*). Check whether the wave-front set is uniquely defined by the
above properties.
Let us now find a necessary condition for (0, v) ∈ WF(ξ). We will use the Radon
transform, which maps f ∈ Cc∞ (V ) to the integrals of f on all affine hyperplanes,
i.e. hyperplanes not necessarily passing through the origin. For a vector space
W , denote W ˉ := P(W ⊕ F ). Note that this is a compact manifold and that the
manifold of all affine hyperplanes in V is V ∗ , and the manifold of all affine lines in
V is Vˉ . Define R := {(l, H) ∈ Vˉ × V ∗ | l ∈ H}, and let p1 : R → Vˉ and p2 : R → V ∗
be the projections. Then the Radon transform is (p2 )∗ ◦ p∗1 . It is known that this
transform is invertible, and thus any distribution is the Radon transform of another
one. Thus, the conditions (9) and (8) give an upper bound on WF. Hopefully, the
two bounds together determine WF.
10.1. Definition of the wave-front set. Let us now give a constructive definition
of WF, following Hörmander, and then prove the above properties. It will take
several steps.
Definition 10.1.1. For any local field F , let V be an F -vector space, v ∈ V and
f ∈ C ∞ (V ). We say that f vanishes asymptotically along v if there exists an
open neighborhood U of v and ρ ∈ Cc∞ (U ) such that p∗ (ρ)m∗ f ∈ S(U × F ), where
m : V × F → V is given by m(v, λ) := λv and p : V × F → F is the projection.
Exercise 10.1.4. For any Lie group G, we have F (Distc (G)) ⊂ C ∞ (Ǧ).
Γ ∩ Sν ⊂ N × {0}.
ν ∗ : CΓ−∞ (N ) → Cν−∞
∗ (Γ) (M ).
Si = CN (V ) = V × W ∗ ⊂ (V ⊕ W ) × (V ∗ ⊕ W ∗ ) = T ∗ (V ⊕ W ).
Also, enough to prove for the case dim W = 1. This we do by twisting pushforward
by F .
More precisely, we use Theorem 9.0.26 on functoriality of Fourier transform. By this
theorem, for f ∈ S(V ⊗W ) we have F (i∗ f ) = p∗ (F (f )), where p : V ∗ ⊕W ∗ → V ∗ is
the projection. This formula extends to ξ ∈ CV−∞ ×W ∗ (V ⊕ W ), since F (ρξ) vanishes
asymptotically along W for any ρ ∈ Cc (V ⊕ W ). More precisely, we can choose a
∗ ∞
Intuitively, this theorem makes sense since the cotangent directions in the wave-
front set form the asymptotic support of the Fourier transform. Let us now give the
proof in the p-adic case, since the proof in the real case is similar, though longer.
Proof. Enough to show that for any ρ ∈ Cc∞ (V ), F (ρξ) is asymptotically supported
in Z. Note that F (ρξ) = F (ρ) ∗ F (ξ) and thus suppF (ρξ) ⊂ suppF (ρ) + Z. Since
suppF (ρ) is a compact set, F (ρξ) is eventually zero in any direction not in Z.
Since Z is closed, this implies that F (ρξ) is asymptotically supported in Z and
thus WF(ξ) ⊂ V × Z.
GENERALIZED FUNCTIONS LECTURES 76
Sketch of proof. Since the question is local, we can assume that M is a vector space
V . Fix x ∈ V and let p be the polynomial on V ∗ given by p(l) := Symb(D)(v, l).
Let l ∈ V ∗ be a cotangent direction such that p(l) 6= 0. We want to show that
p∈/ WFx (ξ). Define
ln |F (ξ)(λw)|
ν(ξ, l) := sup lim
w∈l+B (0) λ→∞ ln |λ|
This is the order of asymptotics as λ → ∞ of F (ξ) near the direction l. It is not
+∞ since F (ξ) is a tempered distribution. We want to show that ν(ξ, l) = −∞ for
small enough. Denote ψ(α) := exp(2πiα), and define fλ ∈ C ∞ (V ) by fλ (v) :=
ψ(λl(v)). We have
0 = hρDξ, fλ i = hξ, D(ρfλ )i.
Using the Leibnitz rule we get
degD
X degD
X
D(ρfλ )(v) = D(ρ(v)ψ(λl(v)) = λm ψ(λl(v))ρm = λm f ρ m
m=0 m=0
Thus
F (ξ) ∗ F (ρdegD )(λl) = −λ−1 (F (ξ) ∗ F (ρdegD−1 ))(λl) + . . .
and thus ν(ξ, l) = ν(ξ, l) − 1. Thus ν(ξ, l) = −∞.
Then
F (ν1∗ ξ)|Bε−1 (0) = F (ν2∗ ξ)|Bε−1 (0) .
Exercise 10.3.6. Let r, ε, α > 0 with ε < 1 and let l ∈ V ∗ s.t. ||l|| = 1. Let
A : V → V be an affine transformation s.t. ||d0 A − Id|| < 1 and A(Br (0)) ⊂ Br (0).
Then
−∞
A∗ (Cl,r,ε,α (V )) ⊂ Cd−∞
∗ A(l),r,ε,α (V ).
0
This exercise follows from the rules for conjugation of linear transformations by
Fourier transform.
The theorem follows now from the following specialization.
Theorem 10.3.7. Let r, ε, α > 0 with ε < 1 and let l ∈ V ∗ s.t. ||l|| = 1. Let
ν : V → V be a diffeomorphism such that
(1) ν(0) = 0
(2) ||dx ν − Id|| < ε
(3) for any x ∈ Br (0) and any r0 ≤ r we have ν(Br0 (x)) = Br0 (ν(x)).
−∞
Let ξ ∈ Cl,r,ε,α (V ). Then
−∞
ν ∗ ξ ∈ Cl,r,ε (V ).
p
Proof. Let C be as in Exercise 10.3.4. Fix λ ∈ F with |λ| > α + C + C(C + 2α).
√
Let δ = 1/( 2λC). Thus α + δ −1 < |λ| < C −1 δ −2 . It is enough to show that for
any l0 ∈ Bε (l) we have F (ν ∗ ξ)(λl0 ) = 0. Present Br (0) as a disjoint union of balls
Ui = Bδ (xi ) of radius δ. We have
X
F (ν ∗ ξ)(λl0 ) = F (ν ∗ (1Ui ξ))(λl0 ).
i
Summarizing, we have
X X
F (ν ∗ ξ)(λl0 ) = F (ν ∗ (1Ui ξ))(λl0 ) = F ((Af fxi ν)∗ (1Ui ξ))(λl0 ) = 0.
i i
Remark 10.3.8. (1) The method in the proof is called the stationary phase
method, which is a central method in microlocal analysis.
P
(2) In order to use the affine approximation we decomposed ξ = i 1Ui ξ. The
multiplication of ξ by 1Ui “damaged" ξ but this “damage" can be controlled
using 10.3.2(i), and is apparently moderate since the affine approximation
is good to the second order.
In this section we show that the Fourier transform is not alone - it is part of an
infinite group of operators. This group is a representation of a double cover of
SL2 (F ). Thus this section requires some knowledge of representation theory. To
motivate the existence of this representation we first describe the Heisenberg group
and its representations.
Note that the center of Hn is 0 × F , and it acts on σ by the character χ, which can
be trivially extended to a character of V ∗ × F .
It is easy to see that σ is the unitary induction of (the extension of) the character
χ from V ∗ × R to Hn = (V ⊕ V ∗ ) × F .
Lemma 11.0.3. The space of smooth vectors in σ is S(V ), and the Lie algebra of
Hn acts on it by
Proof. Formula (4) is obtained from (3) by derivation. Now, it is known that the
space of smooth vectors in a unitary induction consists of the smooth L2 functions
whose derivatives also lie in L2 .
Idea of the proof. Let me ignore all the analytic difficulties. Consider the normal
commutative subgroup A := V ×F . Conjugation in Hn defines an action of V on the
dual group of A. This action has only two orbits. The closed orbit is the singalton
{1} and the open orbit O is the complement to the closed one. The restriction σ|A
decomposes to a direct integral of characters in O, each “with multiplicity one".
The restriction of any non-zero subrepresentation ρ ⊂ σ to A will also include χ,
and thus the whole orbit O of χ. Thus ρ = σ and σ is irreducible.
Now let τ be any irreducible unitary representation of Hn with central character
χ. Then the restriction of τ to A will again include all the characters in O with
multiplicity one. Thus τ is the induction of an irreducible representation of the
stabilizer of χ in Hn . However, this stabilizer is A and thus τ ' σ.
The uniqueness part of Corollary 11.0.5 follows from Schur’s lemmas. Corollary
11.0.5 defines a projective representation of Sp(V ⊕ V ∗ ) on S(V ), i.e. a map τ :
Sp(V ⊕ V ∗ ) → GL(S(V )) such that τ (gh) = λg,h τ (g)τ (h). It is not possible to
coordinate the scalars in order to obtain an honest representation of Sp(V ⊕ V ∗ ),
f ⊕ V ∗ ), called the
but it is possible to obtain a representation of a double cover Sp(V
metaplectic group. This was shown by A. Weil. We will not give the formulas for
f 2 (F )
the full Weil representation, but rather for its restriction to the subgroup SL
embedded by
where by IdV we mean the operator that is Id on V and zero on V ∗ and by T the
operator that is zero on V ∗ and on V is given by ω.
Suppose that F is non-archimedean, and identify V with V ∗ using a non-degenerate
quadratic form q. Also, identify σ with S(V ). Let n := dim V . Then the Weil
representation is given by
GENERALIZED FUNCTIONS LECTURES 80
!
1 u
(5) π( , ε)f (v) = εn ψ(uq(v))f (v)
0 1
!
t 0
(6) π( , ε)f (v) = εn |t|n/2 γ(q)γ(tq)−1 f (tv)
0 t−1
!
0 1
(7) π( , ε)f (v) = εn γ(q)F (f )(v)
−1 0
Here, γ(q) is a certain eights root of unity that depends on the quadratic form, and
ψ is the non-trivial unitary additive character that we use to identify V̌ with V ∗ .
Note that this representation defines a representation of SL2 (F ) if and only if n is
even. The existence of the Weil representation and the formulas above imply the
following corollary.
In the case F = R, one can prove a slightly weaker lemma. The problem in
generalizing the argument above to the archimedean case is that in this case the
condition suppF (ξ) ⊂ V does not imply ψ(uq(v))ξ = 0, but rather that there exists
k such that ψ(uq(v))k ξ = 0. On the other hand, in this case one can use the Lie
algebra sl2 .
For the Archimedean version of the corollary we will need the following definition.
Theorem 11.0.8. Let L ⊂ SV∗ (Z(q)) be a non-zero subspace such that for all ξ ∈ L
we have Fq (ξ) ∈ L and q ∙ ξ ∈ L (here B is viewed as a quadratic function).
Then there exists a non-zero distribution ξ ∈ L which is adapted to q.
Lemma 12.1.2. The collection of semi-algebraic sets is closed with respect to finite
unions, finite intersections and complements.
Proof. The graph of ν is obtained from the graph of ν −1 by switching the coordi-
nates.
GENERALIZED FUNCTIONS LECTURES 82
One of the main tools in the theory of semi-algebraic spaces is the Tarski-Seidenberg
principle of quantifier elimination. Here we will formulate and use a special case of
it. We start from the geometric formulation.
pr(A) = SΨ where
Since Ψ ∈ L(R), the proposition follows now from the previous corollary.
In fact, it is not difficult to deduce the logical formulation from the geometric one.
Let us now demonstrate how to use the logical formulation of the Seidenberg-Tarski
theorem.
Corollary 12.1.15.
(i) The composition of semi-algebraic mappings is semi-algebraic.
(ii) The R-valued semi-algebraic functions on a semi-algebraic set A form a ring,
and any nowhere vanishing semi-algebraic function is invertible in this ring.
(iii) Images and preimages of semi-algebraic sets under semi-algebraic mappings
are semi-algebraic.
12.2. Nash manifolds. Let us now define the category of Nash manifolds, i.e.
smooth semi-algebraic manifolds. I like this category since the Nash manifolds
behave as tamely as algebraic varieties (e.g. posses some finiteness properties, and
admit an analog of Hironaka’s desingularization theorem), and in addition their
GENERALIZED FUNCTIONS LECTURES 84
Theorem 12.2.2 (Implicit Function Theorem for Nash manifolds, see e.g. [BCR,
Corollary 2.9.8]). Let (x0 , y 0 ) ∈ Rn+p , and let f1 , ..., fp be semi-algebraic smooth
functions on an open neighborhood of (x0 , y 0 ), such that fj (x0 , y 0 ) = 0 for j = 1, .., p
∂f
and the matrix [ ∂yji (x0 , y 0 )] is invertible. Then there exist open semi-algebraic
neighborhoods U (resp. V) of x0 (resp. y 0 ) in Rn (resp. Rp ) and a Nash mapping
φ, such that φ(x0 ) = y 0 and f1 (x, y) = ... = fp (x, y) = 0 ⇔ y = φ(x) for every
(x, y) ∈ U × V.
By the implicit function theorem it is easy to see that this definition is equivalent
to the following one, given in [BCR]:
Example 12.2.10. Any real nonsingular affine algebraic variety has a natural
structure of an affine Nash manifold.
The equivalence of the definitions follows from the following theorem, which imme-
diately follows from [BCR, Theorem 8.4.6] and Proposition 12.1.16.
GENERALIZED FUNCTIONS LECTURES 86
Theorem 12.2.18 ([AG10, Theorem 2.4.3]). Let M and N be Nash manifolds and
ν : M → N be a surjective submersive Nash map. Then locally (in the restricted
S
k
topology) it has a Nash section, i.e. there exists a finite open cover N = Ui such
i=1
that ν has a Nash section on each Ui .
Z (z).
{(z, v) ∈ NZM s.t. ||v|| < ρM
Definition 12.3.1. Let M be a Nash manifold, and let E be a Nash bundle over
S
k
it. Let M = Ui be an affine Nash trivialization of E. Then we have a map
i=1
L
k
φ: S(Ui ) → C ∞ (M, E). We define the space S(M, E) of global Schwartz
n
i=1
sections of E by S(M, E) := Imφ. We define the topology on this space using the
Lk
isomorphism S(M, E) ∼
= S(Ui )n /Kerφ.
i=1
(5) Partition of unity: Let (Ui )ni=1 be a finite cover by open Nash submani-
folds. Then there exist smooth functions α1 , ..., αn such that supp(αi ) ⊂ Ui ,
Pn
αi = 1 and for any g ∈ S(M, E), αi g ∈ S(Ui , E).
i=1
(6) S(M, E) = S(M )S(M, E).
(7) S(M, E) is a nuclear Fréchet space.
ˆ
(8) For any Nash manifold M 0 we have S(M × M 0 ) = S(M )⊗S(M 0
).
0 → SM
∗
\U (M ) → S (M ) → S (U ) → 0.
∗ ∗
Let CNM
N
denote the conormal bundle to N in M . Then
Fi /Fi−1 ∼
= S ∗ (N, Symi+1 (CNM
N
)).
Lemma 12.3.10. Suppose α ∈ C ∞ (R) vanishes at 0 with all its derivatives. Then
for any natural number n, α(t) = (n!)−1 tn α(n) (θ) for some θ ∈ [0, t].
Case 2 M is affine.
Follows from the previous case and property (3) (extension from a closed
Nash submanifold).
Case 3 General case.
Choose an affine cover of M . The theorem now follows from the previous
case and partition of unity.
Recall that this property implies, by the Hahn-Banach theorem, that the restriction
S ∗ (M ) → S ∗ (U ) is onto (Corollary 12.3.3). Let us demonstrate a classical corollary
of this fact.
GENERALIZED FUNCTIONS LECTURES 92
Proof. For the affine case embed M in Rn and extend ξ to Rn . There find f and
D from the previous corollary and restrict them to M . The general case follows by
partition of unity.
Lemma 12.3.13. For any homogeneity degree α, the space of α-homogeneous even
distributions on F is one-dimensional.
The same statement with the same proof holds for odd distributions. Here, even
distribution means invariant under the coordinate change x 7→ −x, and odd means
anti-invariant.
Finally, we would like to remark on a different, extrinsic, approach to Schwartz
functions, applied in [CHM] and [KS]. We can compactify our manifold and de-
fine Schwartz functions on it as smooth functions on the (smooth) compactification
that vanish to infinite order on the complement to M . If both M and the com-
pactification are Nash manifolds then this definition will be equivalent, by Theorem
12.3.2(2,4). This allows to define Schwartz functions on non-Nash (say, subanalytic)
manifolds, but this space will depend on the compactification.
Schwartz functions on non-smooth algebraic varieties and more generally on Nash
varieties (i.e. varieties that can be locally described as zeros of Nash functions) can
be locally defined by restriction from a smooth ambient space, see [ESh, Ela].
Another realm in which one can define Schwartz functions is the category of tem-
pered manifolds. Indeed, our only use of (semi-)algebraicity was to have a scale of
infinitesimals. Such a scale exists in the wider generality of tempered manifolds -
manifolds that can be covered by open subsets that are identified with open subsets
in Rn , and coordinate changes between these subsets are tempered functions. A
tempered manifold is said to be of finite type if this cover is finite. Many properties
of Schwartz functions listed above continue to hold for tempered manifolds of finite
type, see [Sha].
GENERALIZED FUNCTIONS LECTURES 93
Let gln (F ) denote the vector space of all square matrices of order n with coefficients
in F and GLn (F ) ⊂ gln (F ) denote the group of all invertible matrices of!order n.
g 0
Consider also the embedding of GLn (F ) into GLn+1 (F ) by g 7→ .
0 1
For a distribution ξ ∈ S ∗ (GLn+1 (F )) denote by ξ t the distribution given by hξ t , f i :=
hξ, f t i, where f t (X) = f (X t ) and X t denotes the transposed matrix. We let
GLn (F ) act on GLn+1 (F ) by conjugation via the embedding above. This action
defines an action on Schwartz functions and thus also on tempered distributions.
Denote by S ∗ (GLn+1 (F ))GLn (F ) the subspace of distributions invariant under this
action. In this section we sketch the proof of the following theorem
(λ, 1) ∙ (x, y) = (λx, λ−1 y), and (1, −1) ∙ (x, y) = (y, x).
We can thus consider the second copy only, and prove S ∗ (F 2 )G,χ = 0.
Let U := {(x, y) ∈ F 2 | xy 6= 0}. Then U is locally isomorphic as a G-manifold to
F × ×F × , where the action on the first F × is trivial, and the action on the second one
is given by (λ, ε)∙z = (λ2 z)ε . Clearly, with this action we have S ∗ (F × ×F × )G,χ = 0
and thus we obtain
(10) S ∗ (U )G,χ = 0.
Now, note that the Fourier transform preserves the space S ∗ (F 2 )G,χ . By (10), for
any S ∗ (F 2 )G,χ ⊂ SZ∗ (F 2 ), where Z is the zero set of the quadratic form q(x, y) = xy.
Thus, by Corollary 11.0.6, Proposition 11.0.9 and 13.1.2 we have S ∗ (F 2 )G,χ = 0.
References
[AG08] A. Aizenbud, D. Gourevitch: Schwartz functions on Nash Manifolds, International
Mathematics Research Notices IMRN 2008, no. 5, Art. ID rnm 155, 37 pp. See also
arXiv:0704.2891v3 [math.AG].
[AG10] A. Aizenbud, D. Gourevitch:.: De-Rham theorem and Shapiro lemma for Schwartz func-
tions on Nash manifolds, Israel Journal of Mathematics 177 (2010), pp 155-188. See also
arXiv:0802.3305[math.AG].
[AG09] A. Aizenbud, D. Gourevitch:Generalized Harish-Chandra descent, Gelfand pairs and an
Archimedean analog of Jacquet-Rallis’ Theorem, Duke Mathematical Journal, 149, n. 3,
509-567 (2009). See also arXiv: 0812.5063[math.RT].
[AG13] A. Aizenbud, D. Gourevitch: Smooth Transfer of Kloostermann Integrals, American Jour-
nal of Mathematics 135, 143-182 (2013). See also arXiv:1001.2490[math.RT].
[AT08] A. Arhangel’skii and M. Tkachenko, Topological groups and related structures, Atlantis
Studies in Mathematics, Vol 1, Atlantis Press, Paris; World Scientific Publishing Co. Pte.
Ltd., Hackensack, NJ (2008).
[BCR] J. Bochnak; M. Coste, M-F. Roy: Real Algebraic Geometry Berlin: Springer, 1998.
[CHM] W. Casselman, H. Hecht, Henryk; D. Miličić: Bruhat filtrations and Whittaker vectors for
real groups. The mathematical legacy of Harish-Chandra (Baltimore, MD, 1998), 151-190,
Proc. Sympos. Pure Math., 68, Amer. Math. Soc., Providence, RI, (2000).
[dCl] F. du Cloux: Sur les representations differentiables des groupes de Lie algebriques. Annales
scientifiques de l E.N.S. 4e serie, tome 24, no 3, p. 257-318 (1991).
[Ela] B. Elazar: Schwartz functions on quasi-Nash varieties, preprint, arXiv:1711.05804.
GENERALIZED FUNCTIONS LECTURES 95