Manifolds. Lecture Notes - I. Marcut.
Manifolds. Lecture Notes - I. Marcut.
Manifolds. Lecture Notes - I. Marcut.
1
Contents
Manifolds. Lecture Notes - Fall 2017 1
Lecture 1. 9
1.1. Brief review of notions from Topology and Analysis 9
1.2. Manifolds 12
1.3. Exercises 17
Lecture 2. 21
2.1. Smooth maps 21
2.2. How many smooth structures? 22
2.3. C k -manifolds 23
2.4. Embedded manifolds in Rn 24
2.5. Exercises 25
Lecture 3. 27
3.1. The topology is determined by the atlas 27
3.2. The real projective space 28
3.3. Smooth manifolds as quotients of group actions 30
3.4. Manifolds of dimensions 1 and 2 32
3.5. Exercises 33
Lecture 4. 37
4.1. Bump functions 37
4.2. Partitions of unity 37
4.3. Proof of Theorem 4.2.4 39
4.4. Corollaries of Theorem 4.2.4 41
4.5. Exercises 42
Lecture 5. 43
5.1. The tangent space 43
5.2. Notations for tangent vectors 45
5.3. The commutative algebra C ∞ (M ) 46
5.4. Tangent vectors as derivations 46
5.5. Exercises 49
Lecture 6. 53
6.1. The inverse function theorem 53
6.2. Immersions and submersions 53
3
4 IOAN MĂRCUT
, , MANIFOLDS
Lecture 7. 61
7.1. Immersed submanifolds 61
7.2. A weak version of Whitney’s Embedding Theorem 65
7.3. Exercises 66
Lecture 8. 69
8.1. The tangent bundle 69
8.2. Vector bundles 70
8.3. Sections of vector bundles 71
8.4. Frames 72
8.5. Vector fields as sections of the tangent bundle 73
8.6. Exercises 75
Lecture 9. 77
9.1. Vector fields as derivations 77
9.2. The Lie bracket 78
9.3. Related vector fields 79
9.4. Change of coordinates rule for vector fields 81
9.5. Integral curves of a vector field 81
9.6. Exercises 83
Lecture 10. 85
10.1. The flow of a vector field 85
10.2. Complete vector fields 87
10.3. The flow as the exponential of the Lie derivative 88
10.4. The Lie bracket as infinitesimal commutator of flows 89
10.5. Exercises 90
Lecture 11. 93
11.1. Lie groups 93
11.2. Exercises 97
Bibliography 147
Exercises and assignments 149
Problems for exercise classes 149
Homework assignments 149
6 IOAN MĂRCUT
, , MANIFOLDS
• The symbol ♣ next to the statement of a theorem indicates that the proof
will not be given in this course, mostly because it is too involved or too
long.
LECTURE 1
We start by recalling some notions from Analysis and Topology, which will be
used throughout the course.
Some Topology
Definition 1.1.1. A topology on a set X is a set T of subsets of X, whose
elements are called open subsets, satisfying the axioms:
[
∅, X ∈ T , U, V ∈ T ⇒ U ∩ V ∈ T , (Ui ∈ T , ∀i ∈ I) ⇒ Ui ∈ T .
i∈I
Definition 1.1.6. (1) The open ball in Rn of radius ε > 0 and center x0 ∈ Rn
is the set
Bε (x0 ) = {x ∈ Rn : |x − x0 | < ε},
where | · | is the usual Euclidean length:
p
|x − y| := (x1 − y 1 )2 + . . . + (xn − y n )2 .
(2) A subset U ⊂ Rn is called open if for every x0 ∈ U there exists a positive
number ε > 0 such that Bε (x0 ) ⊂ U . The collection of all such sets forms the
so-called Euclidean topology on Rn .
Smooth functions
Let U ⊂ Rn be an open set. A function f : U → Rm is called differentiable at a
point x ∈ U , if there exists a linear map
dx f : Rn −→ Rm
such that:
|f (x + h) − f (x) − dx f (h)|
lim = 0.
|h|→0 |h|
The map dx f is called the differential of f at x. If dx f exists, then it is also
uniquely determined, since it can be described as the directional derivative of f :
f (x + εv) − f (x)
dx f (v) = lim , v ∈ Rn .
ε→0 ε
Denote the components of f by f = (f 1 , . . . , f m ). The matrix of dx f in the standard
bases of Rn and Rm is given by the Jacobian matrix:
∂f 1 ∂f 1 ∂f 1
∂x12 ∂x22 . . . ∂x n
∂f ∂f ∂f 2
dx f = ∂x1
∂x2 . . . ∂x n
.
... ... ... ...
∂f m ∂f m m
∂x1 ∂x2 . . . ∂f
∂xn
Recall also the chain rule for the differential:
Proposition 1.1.7 (The chain rule). Consider two maps f : U → Rm , g : V → U ,
where U ⊂ Rl and V ⊂ Rn are open sets. If g is differentiable at x ∈ V and f
is differentiable at g(x) ∈ U then f ◦ g is differentiable at x, and its differential is
given by
dx (f ◦ g) = (dg(x) f ) ◦ dx g : Rn −→ Rm .
In terms of partial derivatives the chain rule takes the form:
l
∂(f ◦ g) X ∂f ∂g j
i
(x) = j
(g(x)) i (x), for all 1 ≤ i ≤ n.
∂x j=1
∂y ∂x
The first, and most general class of maps are continuous maps; these are called
also functions of class zero, or just C 0 -functions.
Denote the space of all linear maps from Rn to Rm by Lin(Rn , Rm ). Writing
the linear maps in the standard bases, this space can be identified with the space of
n×m-matrices, hence with Rn×m . Using this identification, we endow Lin(Rn , Rm )
with the Euclidean topology on Rn×m .
LECTURE 1. 11
df : U → Lin(Rn , Rm ), x 7→ dx f ∈ Lin(Rn , Rm ).
function.
f = (f 1 , . . . , f m ) : U −→ Rm
be a function.
(1) If f is a C k -function then it is also a C k−1 -function.
(2) The function f is of class C k if and only if its partial derivatives of order ≤ k:
∂lf i ∂f i
∂ ∂
= . . . . . . : U → R,
∂xj1 ∂xj2 . . . ∂xjl ∂xj1 ∂xj2 ∂xjl
for 1 ≤ i ≤ m, 0 ≤ l ≤ k, 1 ≤ j1 , . . . , jl ≤ n
exist and are continuous on U .
(3) If f is of class C k then its partial derivatives “commute”, meaning that:
∂lf i ∂lf i
= ,
∂xj1 ∂xj2 . . . ∂xjl ∂xjσ(1) ∂xjσ(2) . . . ∂xjσ(l)
for 1 ≤ i ≤ m, 1 ≤ l ≤ k, 1 ≤ j1 , . . . , jl ≤ n, and every permutation σ of the
set {1, . . . , l}.
The following lemma, which will be repeatedly used throughout this course,
illustrates the richness of the class of smooth functions.
Lemma 1.1.11. Fix a point a ∈ Rm , and positive numbers 0 < < δ. There exists
a smooth map χ : Rm → R such that
• 0 ≤ χ(x) ≤ 1, for all x ∈ Rm ,
• χ(x) = 1, for all x ∈ B (a),
• χ(x) = 0, for all x ∈
/ Bδ (a).
This expression belongs to the vector space of formal power series in the Ti ’s with
coefficients in Rm , denoted:
X
Rm [[T1 , . . . , Tn ]] := ci1 ,...,in T1i1 . . . Tnin : ci1 ,...,in ∈ Rm .
i1 ,...,in ≥0
Definition 1.1.12. The function f is called analytic around a if there exists > 0
such that for all x ∈ B (a), the Taylor series evaluated on T = x − a converges
absolutely to f (x):
X 1 ∂ i1 +...+in f
f (x) = (a)(x1 − a1 )i1 . . . (xn − an )in .
i1 ! . . . in ! (∂x )i1 . . . (∂xn )in
1
i1 ,...,in ≥0
1.2. Manifolds
Topological manifolds
Definition 1.2.1. An m-dimensional topological manifold is a Hausdorff, sec-
ond countable topological space, for which every point has an open neighborhood
homeomorphic to an open set in Rm .
The last condition means that a topological manifold looks locally like Rm .
The following terminology is used:
LECTURE 1. 13
Smooth manifolds
Let us recall:
Definition 1.2.4. A map f : U → V between open sets U, V ⊂ Rm is called a
diffeomorphism if f is bijective and both f and f −1 are smooth.
Note that (as in the case of homeomorphisms) the condition that f −1 be smooth
is not automatically satisfied. The standard example of a smooth bijection which
is not a diffeomorphism is
f : R −→ R, f (t) = t3 .
In a local chart a topological manifold is described as an open piece of Rm . To
develop analysis on manifolds, one needs to introduce derivatives and integration of
functions, notions which, locally in charts, should coincide with those from multi-
variable calculus. However, a function that is differentiable in one chart might fail
to be differentiable in different chart! To circumvent this, on a smooth manifold
one only works with mutually compatible charts.
Definition 1.2.5. Let M be a topological space. Let (U, ϕ) and (V, ψ) be two m-
dimensional charts on M . The map
ϕ ◦ ψ −1 : ψ(U ∩ V ) −→ ϕ(U ∩ V )
14 IOAN MĂRCUT
, , MANIFOLDS
is a called the change of coordinates map or the transition map between the
two charts. The two charts are said to be compatible if the transition map is a
diffeomorphism.
Definition 1.2.6. An m-dimensional topological atlas A on an m-dimensional
topological manifold M is said to be an m-dimensional smooth atlas (or C ∞ -
atlas) if every two charts in A are compatible.
Here are some simple examples of C ∞ -atlases:
Example 1.2.7. (1) On Rm there is a smooth atlas with only one chart
A = {(Rm , idRm )}.
Another smooth atlas is the collection of all diffeomorphisms between open
subsets of Rm :
B = { (U, ϕ) : U, V ⊂ Rm are open and ϕ : U → V is a diffeomorphism}.
In fact, B consists of all charts compatible with (Rm , idRm ).
(2) Consider the m-dimensional sphere with the induced topology from Rm+1 :
S m = (x0 , x1 , . . . , xm ) : (x0 )2 + (x1 )2 + . . . + (xm )2 = 1 ⊂ Rm+1 .
We construct an atlas on S m with only two charts. The “south pole” and the
“north pole” of S m are the points
s := (0, 0, . . . , 0, −1) ∈ S m , resp. n := (0, 0, . . . , 0, 1) ∈ S m .
The stereographic projection through the north pole is the map
πn : S m \{n} −→ Rm
which sends a point p ∈ S m , p 6= n, to the point q = πn (p) which is the
intersection of the plane xm = 0 and the line through n and p.
n
Sm
q
xm = 0
p
and so,
1
πn (x0 , . . . , xm ) = (x0 , . . . , xm−1 ).
1 − xm
Similarly, we have the stereographic projection through the south pole:
1
πs : S m \{s} → Rm , πs (x0 , . . . , xm ) = (x0 , . . . , xm−1 ).
1 + xm
We claim that the following is a smooth atlas on S m :
A = {(S m \{n}, πn ), (S m \{s}, πs )}.
LECTURE 1. 15
of πn is given by:
πn−1 : Rm −→ S m \{n},
2y 1 2y 2 2y m |y|2 − 1
−1 1 2 m
πn (y , y , . . . , y ) = , ,..., 2 , .
|y|2 + 1 |y|2 + 1 |y| + 1 |y|2 + 1
All functions appearing in this expression are continuous, and this proves that
πn is a homeomorphism. A similar calculation shows that also πs is a homeo-
morphism.
Finally, we need to check that the charts are C ∞ -compatible. Since πn (s) =
πs (n) = 0, the transition map is defined as follows:
πs ◦ πn−1 : Rm \{0} −→ Rm \{0},
and using the formulas above, we obtain that this map is given by:
|y|2 − 1
2y 1 2y y
πs ◦ πn−1 (y) = πs 2
, 2
= |y|2 −1 2
= 2.
|y| + 1 |y| + 1 1 + 2 |y| + 1 |y|
|y| +1
This map is clearly smooth. Note that the map satisfies (πs ◦ πn−1 )2 = idRm \{0} .
This shows that it equals its own inverse, hence it is a diffeomorphism.
We introduce a relation on atlases:
Definition 1.2.8. Let M be an m-dimensional topological manifold. Consider the
following relation on m-dimensional C ∞ -atlases on M :
def.
A1 ∼ A2 ⇐⇒ A1 ∪ A2 is a C ∞ − atlas.
As expected:
Proposition 1.2.9. The relation ∼ is an equivalence relation on the set of m-
dimensional C ∞ -atlases of M .
Proof. Reflexivity and symmetry of the relation ∼ are obvious. We will check that
transitivity holds. Consider three m-dimensional C ∞ -atlases on M : A1 , A2 and
A3 such that A1 ∼ A2 and A2 ∼ A3 . To check that A1 ∼ A3 we need to show
that each pair of charts (U1 , ϕ1 ) ∈ A1 and (U3 , ϕ3 ) ∈ A3 are compatible, i.e. that
the map
(1) ϕ1 ◦ ϕ−1
3 : ϕ3 (U1 ∩ U3 ) −→ ϕ1 (U1 ∩ U3 )
So, it suffices to prove that the maps (1) and (2) are smooth. Let p ∈ U1 ∩ U3 .
Since A2 is an atlas, there exists a chart (U2 , ϕ2 ) ∈ A2 such that p ∈ U2 . Since
A1 ∼ A2 and A2 ∼ A3 is follows that the following maps are diffeomorphisms:
ϕ1 ◦ ϕ−1
2 : ϕ2 (U1 ∩ U2 ) −→ ϕ1 (U1 ∩ U2 ),
ϕ2 ◦ ϕ−1
3 : ϕ3 (U2 ∩ U3 ) −→ ϕ2 (U2 ∩ U3 ).
In particular, the restriction of their composition is a diffeomorphism:
ϕ1 ◦ ϕ−1
3 : ϕ3 (U1 ∩ U2 ∩ U3 ) −→ ϕ1 (U1 ∩ U2 ∩ U3 ).
Hence, the map (1) and its inverse (2) are smooth when restricted to the open
neighborhoods ϕ3 (U1 ∩ U2 ∩ U3 ) and ϕ1 (U1 ∩ U2 ∩ U3 ), respectively, of ϕ3 (p) and
ϕ1 (p), respectively. Since p was chosen arbitrary in U1 ∩ U3 , it follows that (1) and
(2) are smooth everywhere. This finishes the proof.
Next, we define:
Definition 1.2.10. An m-dimensional C ∞ -atlas A on the topological manifold M
is said to be maximal if
A ∼ B =⇒ B ⊂ A
∞
for any m-dimensional C -atlas B on M .
We have that:
Proposition 1.2.11. Any equivalence class of smooth atlases has a unique maximal
representative. The maximal C ∞ -atlas equivalent to the C ∞ -atlas A is given by:
Amax = {(U, ϕ) : A ∪ {(U, ϕ)} is a C ∞ -atlas} .
Proof. Let A be an m-dimensional C ∞ -atlas on M . It is clear that there exists
at most one maximal smooth atlas which is equivalent to A, because if there were
two, A1max and A2max , then by transitivity of the relation ∼ we would have that
A1max ∼ A2max , and since both atlases are maximal, this implies that
A1max ⊂ A2max ⊂ A1max ;
thus A1max = A2max .
To prove existence of a maximal C ∞ -atlas equivalent to A, let Amax be the set
constructed in the statement. We check that Amax is indeed a C ∞ -atlas. First,
since A ⊂ Amax , it follows that the open sets in Amax cover M . Second, we check
that each pair of charts (U, ϕ) and (V, ψ) in Amax are compatible. Clearly A ∼
A ∪ {(U, ϕ)} and A ∼ A ∪ {(V, ψ)}, therefore, by transitivity of ∼, A ∪ {(U, ϕ)} ∼
A ∪ {(V, ψ)}, hence A ∪ {(U, ϕ)} ∪ {(V, ψ)} is a C ∞ -atlas. This implies that the
charts (U, ϕ) and (V, ψ) are compatible, and proves that Amax is indeed a C ∞ -atlas.
Finally, we prove that Amax is maximal. Note first that A ∼ Amax . Let B be
a C ∞ -atlas such that B ∼ Amax . Then, by transitivity of ∼, B ∼ A, which implies
that every chart in B is compatible with every chart in A; hence A ∪ {(U, ϕ)} is an
atlas for every (U, ϕ) ∈ B. The definition of Amax implies now that B ⊂ Amax .
Definition 1.2.12. An m-dimensional differentiable structure (or smooth
structure) on a topological manifold M is an equivalence class of m-dimensional
C ∞ -atlases on M . Equivalently, by Proposition 1.2.11, a differentiable structure is
the same as a maximal atlas on M .
We are ready to give the general definition of a smooth manifold:
LECTURE 1. 17
1.3. Exercises
Exercise 1.1. (a) Show that Rn is Hausdorff and second countable.
(b) Show that if X is a Hausdorff topological space and Y ⊂ X is a subset, then
the induced topology on Y is Hausdorff.
(c) Show that if X is a second countable topological space and Y ⊂ X is a subset,
then the induced topology on Y is second countable.
Exercise 1.2. (a) Prove that the following function is smooth but is not analytic
0, x ≤ 0;
f : R −→ R, f (x) =
e−1/x , x > 0.
(b) Show that g : R → R, g(x) := f (x)f (1 − x) is a smooth function which is
positive on (0, 1) and zero elsewhere. Show also that the function h : R → R,
Rx
g(y)dy
h(x) := R01
0
g(y)dy
is a smooth and it satisfies: h(x) = 0 for x < 0, 0 < h(x) < 1 for 0 < x < 1,
and h(x) = 1 for x > 1.
(c) Use (a) and (b) to prove Lemma 1.1.11.
Exercise 1.3. Prove that there are no analytic functions χ : Rm → R with the
properties from Lemma 1.1.11.
The following exercise is used to prove Borel’s Lemma:
Exercise 1.4. Let Cbk (Rn ) denote the set of C k -function f : Rn → R which have
all partial derivatives up to order k bounded. On this space we define the so-called
C k -norm:
∂ i1 +...+in f
: x ∈ Rn , i1 + . . . + in ≤ k .
kf kk := sup (x)
(∂x1 )i1 . . . (∂xn )in
Let Cb∞ (Rn ) denote the set of smooth maps with all partial derivatives bounded.
18 IOAN MĂRCUT
, , MANIFOLDS
(a) Prove that (Cbk (Rn ), k · kk ) is a Banach space, i.e. prove that: every sequence
of functions in Cbk (Rn ) which is Cauchy with respect to the norm k · kk is in
fact convergent to an element of Cbk (Rn ).
(b) Prove that Cb∞ (Rn ) endowed with the family of norms {k · kk }k≥0 is a Fréchet
space, i.e. prove that: every sequence of functions in Cb∞ (Rn ) which is Cauchy
with respect to all the norms {k · kk }k≥0 is in fact convergent with respect to
all the norms {k · kk }k≥0 to an element in Cb∞ (Rn ).
In the next exercise you are asked to prove Borel’s Lemma; note that it uses
the previous exercise.
Exercise 1.5. Consider a formal power series
X
Q(T1 , . . . , Tn ) := ci1 ,...,in T1i1 . . . Tnin ∈ R[[T1 , . . . , Tn ]].
i1 ,...,in ≥0
n
Let χ : R → R be a smooth function such that χ(x) = 1, for x ∈ B1 (0), and
χ(x) = 0 for x ∈
/ B2 (0) (by Lemma 1.1.11, such functions exist).
For k ≥ 0, define the following function depending on a positive number k > 0:
X x
fk (x) := χ · ci1 ,...,in (x1 )i1 . . . (xn )in .
k
i1 +...+in =k
(b) Prove that the atlas A from Example 1.2.7 (2) and the atlas B define the same
differentiable structure on S m .
LECTURE 2
Let us be a bit more precise about the second part of the questions. Assume that
M is a smooth manifold in two different ways, corresponding to the differentiable
structures D1 and D2 . We regard these differentiable structures are equivalent, if
there exists a diffeomorphism relating the two structure:
ϕ : (M, D1 ) −→ (M, D2 ).
Thus, the second part of the questions asks how many non-equivalent such differ-
entiable structures exist.
For small dimension, we have existence and uniqueness:
Theorem ♣ 2.2.1. For m ≤ 3, every m-dimensional topological manifold has a
smooth structure, and any two such smooth structures are equivalent (for m = 1, 2
Radó, 1920s, for m = 3 Moise, 1953, [9]).
However, in larger dimension:
Theorem ♣ 2.2.2. There exist compact topological manifolds of dimension ≥ 4
which do not admit a differentiable structure (Kervaire 1960, [6] first example in
dimension 10; nowadays many examples are known even of dimension 4).
Finally, also uniqueness fails in dimension m ≥ 4; there are topological man-
ifolds which admit non-equivalent smooth structures. The non-standard ones are
usually called exotic. The first examples of non-equivalent smooth structures on a
topological manifold were Milnor’s exotic 7-spheres:
Theorem ♣ 2.2.3. The topological manifold S 7 admits exactly 28 non-equivalent
smooth structures (Milnor 1956, [7]).
LECTURE 2. 23
dimension 1 2 3 4 5 6 7 8 9 10
C ∞ − str.’s 1 1 1 ? 1 1 28 2 8 6
dimension 11 12 13 14 15 16 17 18 19 20
C ∞ − str.’s 992 1 3 2 16256 2 16 16 523264 24
Even more surprisingly are the exotic R4 ’s:
Theorem ♣ 2.2.4. For m 6= 4, the topological manifold Rm admits a unique
smooth structure (Stallings, 1962 [12]). The topological manifold R4 admits un-
countably many non-equivalent smooth structures (Taubes 1987, [13]).
2.3. C k -manifolds
Throughout this section we fix k which is either an integer k ≥ 1, or k ∈ {∞, ω}.
We briefly introduce the notion of C k -manifolds.
All notions discussed so far can be extended directly to the C k -setting, by
simply replacing the requirement that a map be smooth with that of being C k .
For example, a C k -diffeomorphism is a map f : U → V between open sets
U, V ⊂ Rm which is bijective and both f and f −1 are C k -maps. Note the following:
Lemma 2.3.1. If f : U → V is a C 1 -diffeomorphism and a C k -map, then f is a
C k -diffeomorphism.
Proof. We need to check that the inverse of f , denoted by g : V → U , is also of
class C k . For k = ω, we omit the proof (one should write the power series expansion
of the equation f ◦ g = id, and determine a term-by-term formula for g, and then
show that this is convergent). Let k 6= ω. Differentiating the relation f ◦ g = id, by
the chain rule, we obtain that
dg(x) f ◦ dx g = id.
This gives that the differential of g is the inverse of the differential of f :
dx g = (dg(x) f )−1 .
This relation shows that if g is C l and f is C l+1 then g is C l+1 (the right hand side
is a composition of C l -maps, and the smooth (!) operation of taking the inverse of
a matrix). An argument by induction, for 1 ≤ l ≤ k, proves the statement.
One introduces C k -compatible charts, C k -atlases, C k -differentiable structures,
and finally, C k -manifolds, all in perfect analogy with the smooth setting.
Proof. According to Exercise 1.1, the induced topology is Hausdorff and second
countable. We need to check that A is a smooth atlas on M . By definition, M is
covered by adapted charts. Let (U, f ) be an adapted chart, and denote V := f (U )
and ϕ := prnm ◦ f |M ∩U . First, note that M ∩ U is open in the induced topology
on M , and the map ϕ is continuous for the induced topology, because it is the
restriction of a continuous map on U . We have that
ϕ(M ∩ U ) = prnm ◦ f (M ∩ U ) = prnm ((Rm × {0}) ∩ V ) ⊂ Rm .
Since prnm restricts to a homeomorphism between Rm × {0} (with the induced
topology) and Rm , we have that ϕ(M ∩ U ) is open in Rm . Finally, the map ϕ is a
homeomorphism onto its image, because its inverse is given by as a composition of
continuous maps:
ϕ−1 : ϕ(M ∩ U ) −→ M ∩ U, ϕ−1 = f −1 ◦ im
n |ϕ(M ∩U ) .
This proves that (M ∩ U, ϕ) is an m-dimensional chart on M . Thus, M is a
topological m-dimensional manifold.
To show that A is a smooth atlas, we need to check that any two charts are
C ∞ -compatible. Let f : U → V and f 0 : U 0 → V 0 be two adapted charts, inducing
the following charts in A:
(M ∩ U, ϕ := prnm ◦ f |M ∩U ), (M ∩ U 0 , ϕ0 := prnm ◦ f 0 |M ∩U 0 ).
The corresponding transition map is smooth since it is given by:
ϕ0 ◦ ϕ−1 : ϕ(M ∩ U ∩ U 0 ) −→ ϕ0 (M ∩ U ∩ U 0 ), prnm ◦ f 0 ◦ f −1 ◦ im
n.
By interchanging the roles of f and f 0 we see that its inverse is also smooth. Hence
any two charts in A are C ∞ -compatible.
The converse of this result also holds:
Theorem ♣ 2.4.4 (Whitney’s Embedding Theorem). Any m-dimensional smooth
manifold, for m ≥ 1, is diffeomorphic to an embedded manifold in R2m , which is a
closed subset.
2.5. Exercises
Exercise 2.1. Prove Proposition 2.1.6.
Exercise 2.2. Let M be an embedded manifold in Rn . Show that the inclusion
map M ,→ Rn , x 7→ x, is smooth.
Exercise 2.3. (a) Prove that S m is an embedded manifold in Rm+1 (as in Defini-
tion 2.4.1).
(b) Prove that the smooth structure on S m resulting from Proposition 2.4.3 co-
incides with the smooth structure on S m coming from the smooth atlas in
Example 1.2.7.
LECTURE 3
then TA is Hausdorff.
• If TA is Hausdorff and second countable, then M has an m-dimensional
differentiable structure for which A is a C ∞ -atlas.
Proof. First, we show that there is at most one topology TA on M for which A is
an atlas: since {Uα }α∈I is an open cover of M w.r.t. TA , we have that
V ∈ TA ⇐⇒ ∀ α ∈ I : V ∩ Uα ∈ TA ;
and since ϕα : Uα → ϕα (Uα ) is a homeomorphism, we have that
V ∈ TA ⇐⇒ ∀ α ∈ I : ϕα (V ∩ Uα ) is open in Rm ).
27
28 IOAN MĂRCUT
, , MANIFOLDS
The above defines TA uniquely. Next we show that TA is indeed a topology. Clearly,
∅ ∈ TA , and by (3), M ∈ TA . Let V1 , V2 ∈ TA . For every α ∈ I, we have that
(2)
ϕα (Uα ∩ V1 ∩ V2 ) = ϕα ((Uα ∩ V1 ) ∩ (Uα ∩ V2 )) = ϕα (Uα ∩ V1 ) ∩ ϕα (Uα ∩ V2 ),
which shows that ϕα (Uα ∩ V1 ∩ V2 ) is open. This implies that V1 ∩ V2 ∈ TA . Finally,
consider an arbitrary family {Vλ }λ∈Λ ⊂ TA . Then, for each α ∈ I,
ϕα (Uα ∩ (∪λ∈Λ Vλ )) = ϕα (∪λ∈Λ Uα ∩ Vλ ) = ∪λ∈Λ ϕα (Uα ∩ Vλ ),
which shows that ϕα (Uα ∩ (∪λ∈Λ Vλ )) is open. Hence, ∪λ∈Λ Vλ ∈ TA . We conclude
that TA is a topology.
Let us see that (Uα , ϕα ) is indeed a chart on M for the topology TA . By (3),
Uα is open. Next, we show that ϕα : Uα → ϕα (Uα ) is a homeomorphism. For any
open set O ⊂ ϕα (Uα ), we have that ϕβ (Uβ ∩ ϕ−1 −1
α (O)) = ϕβ ◦ ϕα (O) which is open
because, by (4), ϕβ ◦ ϕα is a homeomorphism. Since β is arbitrary, ϕ−1
−1
α (O) ∈ TA .
So, ϕα is continuous. By definition of TA , we have that also ϕ−1 α : ϕα (Uα ) → M
is continuous. Thus, the maps ϕα : Uα → ϕα (Uα ) are homeomorphisms, i.e. charts
on M for the topology TA .
Assuming that I is at most countable, we prove that TA is second countable.
Let {Ok }k∈N be a countable basis for the topology of Rm , i.e. every open set U ⊂ Rm
can be represented as U = ∪k∈JU Ok , for a subset JU ⊂ N. We claim that the family
of open sets {ϕ−1
α (Ok )}α∈I,k∈N is a basis for TA . Since I is at most countable, the
index set I × N is countable. Let V ∈ TA be an open set. First, decompose
V = ∪α∈I (V ∩ Uα ). Since ϕα (V ∩ Uα ) is open in Rm , there exists a subset Jα ⊂ N
such that ϕα (V ∩ Uα ) = ∪k∈Jα Ok . Thus we have that
V = ∪α∈I V ∩ Uα = ∪α∈I ∪k∈Jα ϕ−1
α (Ok ).
We claim that the following two sets are open and separate p and q:
U (p) := {[x0 : . . . : xm ] : |xi | > |xj |}, U (q) := {[x0 : . . . : xm ] : |xi | < |xj |}.
Note first that these sets are well-defined. Clearly, p ∈ U (p), q ∈ U (q) and U (p) ∩
U (q) = ∅. Next, note that U (p) ⊂ Ui and that ϕi (U (p)) is open in Rm . Since ϕi
is a homeomorphism, we conclude that U (p) is open. Similarly, U (q) is open. This
proves that Pm (R) is indeed Hausdorff.
Proof of Theorem 3.3.2. We consider the quotient topology on M/G. First note
that π is an open map, i.e. π(V ) is open M/G for every open set V ⊂ M . This
holds because π −1 (π(V )) is the union of the open sets g · V , g ∈ G.
To see that M/G is second countable, let {Uk }k∈N be a countable basis for the
topology of M . Then the collection {π(Uk )}k∈N is a basis for the topology of M/G
(since π is an open map, these sets are indeed open). Let V ⊂ M/G be any open
set. Then π −1 (V ) = ∪k∈I Uk , for some subset I ⊂ N, and therefore V = ∪k∈I π(Uk ).
Next, we show that M/G is Hausdorff. Let [x], [y] ∈ M/G with [x] 6= [y]. Since
the action is proper, the points have neighborhoods Ux and Uy , respectively, such
that Ux and g · Uy intersect only for a finite number of elements g ∈ G; call these
elements g1 , . . . , gk . Since [x] 6= [y] it follows that gi y 6= x, for 1 ≤ i ≤ k; thus, since
M is Hausdorff, we can find open neighborhoods Ui of gi y and Ux0 ⊂ Ux of x such
−1
0 0
that Ui ∩ Ux = ∅. Let Uy := ∩i=1 gi Ui ∩ Uy . Then the open neighborhoods Ux0
k
and Uy0 of x and y, respectively, have the property that Ux0 ∩ gUy0 = ∅ for all g ∈ G.
This implies that gUx0 ∩ hUy0 = ∅ for all g, h ∈ G; or equivalently, π(Ux0 ) ∩ π(Uy0 ) = ∅.
Since π is an open map, the sets π(Ux0 ) and π(Uy0 ) are disjoint open neighborhoods
of [x], respectively of [y]. Thus M/G is Hausdorff.
Next, we construct charts on M/G. Consider [x] ∈ M/G. Since the action is
proper, there is an open neighborhood U of x such that g · U intersects U only for
a finite number of elements g ∈ G; call these elements g0 := e, g1 , . . . , gk . Since the
action is free, gi · x 6= x, for 1 ≤ i ≤ k; thus we can find open neighborhoods Ui of
gi · x and U 0 ⊂ U of x such that Ui ∩ U 0 = ∅. Let V := ∩ki=1 gi−1 Ui ∩ U 0 . This
Above we have constructed a chart around each [x] ∈ M/G; thus the open sets in
A cover M/G. Finally, we need to prove that the transition maps are smooth. Let
(V, ϕ) and (W, ψ) be two smooth charts on M such that π|V and π|W are injective.
We need to show that the following map is a diffeomorphism:
ϕ ◦ (π|V )−1 ◦ (π|W ) ◦ ψ −1 : ψ(V ∩ π −1 (π(W ))) −→ ϕ(π −1 (π(V )) ∩ W ).
Let p ∈ π(V ) ∩ π(W ), and let x ∈ V and y ∈ W such that [x] = p = [y]. We show
that the transition map is smooth in a neighborhood of ψ(y). Since [x] = [y], there
exists g ∈ G such that x = g · y. Let U := g −1 · V ∩ W . On U , we claim that the
following equality holds:
(π|V )−1 ◦ π|U = ag |U : U −→ V.
By composing on both sides with π|V , the equality clearly holds; and so the claim
follows by injectivity of π|V . Hence, on ψ(U ), the transition map is a composition
of diffeomorphisms
ϕ ◦ ag ◦ ψ −1 : ψ(U ) −→ ϕ(g · U );
32 IOAN MĂRCUT
, , MANIFOLDS
Σg
For g = 0, we obtain the 2-sphere S 2 and for g = 1 the 2-torus T 2 :
Σ0 ∼
= S2 Σ1 ∼
= T2
LECTURE 3. 33
x
Σ3
In this position, the group Z2 = {0̂, 1̂} acts freely on Σg :
0̂ · (x, y, z) = (x, y, z), 1̂ · (x, y, z) = (−x, −y, −z).
The quotient space Σg /Z2 inherits therefore a smooth structure. For g = 0, S 2 /Z2
is the projective space (see Exercise 3.7), and for g = 1, T 2 /Z2 is the Klein bottle
(described also in Exercise 3.9).
We are ready to state the classification theorem for compact surfaces:
Theorem ♣ 3.4.2. Any compact connected 2-dimensional manifold is diffeomor-
phic to either Σg or Σg /Z2 , for some g ≥ 0.
The manifolds in the list Σg , Σg /Z2 , for g ≥ 0, are non-diffeomorphic. Let us
mention that, the surfaces Σg are the orientable ones, and the surfaces Σg /Z2 are
the non-orientable ones; the notion of orientable manifold will be introduced in a
later lecture.
3.5. Exercises
In the following exercise we construct a non-example of a manifold; a non-Hausdorff
second countable topological space with a smooth 1-dimensional atlas.
Exercise 3.1 (Two intervals glued along an open subinterval). Consider the set:
M = (−1, 1) ∪ [2, 3).
Let U = (−1, 1) and V = (−1, 0) ∪ [2, 3). On these subsets of M define the maps
ϕ : U → R, ϕ(t) = t, and
t, t ∈ (−1, 0)
ψ : V → R, ψ(t) =
t − 2, t ∈ [2, 3)
Show that A = {(U, ϕ), (V, ψ)} satisfies the conditions (1)-(4) of Proposition 3.1.1.
Show that the topology TA on M is non-Hausdorff.
Exercise 3.2. Let L be the set of all lines in R2 . For a, b, c ∈ R, with a 6= 0 or
b 6= 0, denote by l(a, b, c) the line given by the equation
ax + by + c = 0.
Let U ⊂ L be the set of lines l(a, b, c) with a 6= 0, and let V ⊂ L be the set of lines
l(a, b, c) with b 6= 0. Define the maps
ϕ : U → R2 , ϕ(l(a, b, c)) = (b/a, c/a), ψ : V → R2 , ψ(l(a, b, c)) = (a/b, c/b).
34 IOAN MĂRCUT
, , MANIFOLDS
Prove that L is a smooth 2-dimensional manifold with atlas A = {(U, ϕ), (V, ψ)}.
Exercise 3.3. Let S 1 denote the unit circle in C
p
S 1 := {z = x + iy : |z| = x2 + y 2 = 1} ⊂ C.
Let M be the collection of subsets of S 1 with 2 elements:
M := {{z, w} : z, w ∈ S 1 , z 6= w}.
(Pay attention to the fact that {z, w} = {w, z}!!)
(a) Consider the set
V := {(α, β) : 0 < α < β < 2π} ⊂ R2 ,
and for k ∈ {0, 1, 2} consider the map ψk : V → M given by
2πk 2πk
ψk (α, β) := {ei(α+ 3 ) , ei(β+ 3 ) }.
Prove that ψk is injective.
(b) Let Uk := ψk (V ), and let ϕk := ψk−1 : Uk → V . Prove that M has the structure
of a smooth 2-dimensional manifold, for which
A := {(Uk , ϕk ) : k ∈ {0, 1, 2}}
is a C ∞ -atlas. (Recall that eiα = cos(α) + i sin(α)).
Exercise 3.4. Show that the manifolds L and M constructed in Exercise 3.2 and
3.3, respectively, are diffeomorphic. Note that both are a Möbius band:
D n−k
X E
χI (A) = Span eiu + au,v ejv : 1 ≤ u ≤ k , for A = (au,v )1≤u≤k,1≤v≤n−k ,
v=1
where Span denotes the linear span of a subset of Rn .
(a) Prove that χI is injective.
(b) Let UI := χI (Rk×(n−k) ) and let ϕI := χ−1
I : UI → R
k×(n−k)
.
Prove that Gr(k, n) is a smooth k × (n − k)-dimensional manifold with atlas:
A = {(UI , ϕI ) : I = {1 ≤ i1 < i2 < . . . < ik ≤ n}}.
(c) Prove that the manifolds Gr(k, n) and Gr(n − k, n) are diffeomorphic.
The following example is from [14].
Exercise 3.6. A continuous action of a group G on a manifold M is called wan-
dering if every point has a neighborhood U such that gU ∩ U = ∅, for all g 6= e.
Consider the action of (Z, +) on M := R2 \{0} given by
n · (x, y) = (2n x, 2−n y).
(a) Prove that this action is wandering, but not proper.
(b) Consider the quotient topology on M/Z. Prove that M/Z is second countable,
but not Hausdorff.
(c) Prove that M/Z has a smooth 2-dimensional atlas, i.e. a collection of charts A
satisfying conditions (1)-(4) from Proposition 3.1.1.
In the following exercise we construct the projective plane as a quotient of the
sphere.
Exercise 3.7. Consider the following action of Z2 = {0̂, 1̂} on S m :
0̂ · x = x, 1̂ · x = −x.
Prove that the action is smooth, proper and free. Prove that S m /Z2 is diffeomorphic
to Pm (R).
In the following exercise we construct the m-dimensional torus.
Exercise 3.8. Consider the action of the group (Zm , +) on Rm given by:
(k 1 , . . . , k m ) · (x1 , . . . , xm ) := (k 1 + x1 , . . . , k m + xm ).
Prove that the action is smooth, proper and free. The quotient manifold
T m := Rm /Zm = {x + Zm : x ∈ Rm }
is called the m-dimensional torus. Prove that T m is diffeomorphic to the product
of m-copies on S 1 (see Proposition 2.1.6 for the construction of the product).
By using actions on R2 , in the following exercise we construct several 2-dimensional
manifolds: the cylinder, the Möbius band, the torus and the Klein bottle.
Exercise 3.9. Consider the following two diffeomorphisms of R2 :
a, b : R2 −→ R2 , a(x, y) = (x, y + 1), b(x, y) = (x + 1/2, −y).
Let Diff(R2 ) be the group of diffeomorphisms of R2 . Let Ga , Gb , Ga,b2 and Ga,b ,
respectively, denote the subgroups of Diff(R2 ) generated by the sets {a}, {b}, {a, b2 }
and {a, b}, respectively.
36 IOAN MĂRCUT
, , MANIFOLDS
(a) Prove that the following relation holds: ba = a−1 b. Using this relation, show
that every element in Ga,b can be represented as an bm , for unique n, m ∈ Z.
Prove that the action of all four groups Ga , Gb , Ga,b2 and Ga,b on R2 is smooth,
free and proper.
(b) Prove that the quotient R2 /Ga is diffeomorphic to the cylinder S 1 × R.
(c) Prove that R2 /Ga,b2 is diffeomorphic to the 2-torus T 2 (from Exercise 3.8).
(d) Prove that R2 /Gb is diffeomorphic to the manifold from Exercise 3.2. (A man-
ifold diffeomorphic to R2 /Gb is called a Möbius band; see also Exercise 3.4.)
(e) Prove that Ga,b2 is an index two subgroup of Ga,b , i.e. Ga,b /Ga,b2 ∼= Z2 . Using
this, define a smooth, free and proper action of Z2 on T 2 such that T 2 /Z2 is
diffeomorphic to R2 /Ga,b . (A manifold diffeomorphic to R2 /Ga,b is called a
Klein bottle.)
Exercise 3.10. Let a : G × M → M be a proper smooth action, and assume that
M is compact. Prove that G is finite.
LECTURE 4
Note that χ is well-defined: on the overlap of the sets W and M \ϕ−1 (B ), we have:
θ ◦ ϕ|W \ϕ−1 (B ) = 0.
Since χ is smooth on these two open sets that cover M , it follows that χ is smooth
on M . The conclusion holds with V := ϕ−1 (B/2 ).
Next, we show next that the family {χj }j∈J is locally finite. Let p ∈ M , and
consider O an open neighborhood of p such that O intersects only a finite number
of the sets supp(ρi ); denote these indexes by i1 , . . . , in ∈ I. If q ∈ supp(χj ), then
by (*), q ∈ supp(ρi ) for some i satisfying f (i) = j. Therefore, O intersects only the
sets supp(χj ), for j = f (i1 ), . . . , f (ik ). Finally, it is clear that:
X X X X
χj = ρi = ρi = 1.
j∈J j∈J i∈f −1 (j) i∈I
The following two results are used in the proof of Lemma 4.3.4.
Lemma 4.3.2. The topology of a manifold has a countable basis all of whose ele-
ments have compact closures.
Proof. Let B be a countable basis for M , and let Bc be the collection of sets in B
with compact closure. Clearly, Bc is countable. We prove that Bc is a basis. Let
O ⊂ M be an open set. For p ∈ O, let Kp ⊂ O be a compact neighborhood (e.g. let
Kp be the preimage by a chart around p of a small closed ball). Since B is a basis,
there exists Up ∈ B such that p ∈ Up ⊂ int(Kp ) ⊂ Op . In particular: Up ∈ Bc and
Up ⊂ O. Thus, we can write O = ∪p∈O Up , which shows that Bc is a basis.
Lemma 4.3.3. Any manifold M has an open cover {Gk }∞
k=1 such that, for k ≥ 1,
Gk is compact and Gk ⊂ Gk+1 .
Proof. By Lemma 4.3.2, there exists a countable basis B = {Bn }n≥1 for the topol-
ogy of M such that B n is compact for all n ≥ 1. Let Om := ∪m n=1 Bn . Note that
Om ⊂ Om+1 , Om = ∪m n=1 B n is compact, and that ∪m≥1 Om = M . Therefore,
for any m ≥ 1 there is a smallest integer f (m) > m such that Om ⊂ Of (m) .
Let m1 := 1, and define inductively for k ≥ 2, mk := f (mk−1 ). The sequence
Gk := Omk satisfies all requirements.
40 IOAN MĂRCUT
, , MANIFOLDS
Theorem 4.2.4 can be used to extend smooth functions defined on closed em-
bedded submanifolds in Rn .
Corollary 4.4.2. Let M ⊂ Rn be an embedded manifold in Rn which is closed as
a subset. For every f ∈ C ∞ (M ) there exists fe ∈ C ∞ (Rn ) such that fe|M = f .
Proof. Recall that around every point in M there is a diffeomorphism ϕ : U → V ,
with V, U ⊂ Rn open, such that
ϕ(M ∩ U ) = (Rm × {0}) ∩ V.
We cover M by a family of such charts {(Ui , ϕi )}i∈I . Consider the following open
cover: {Rn \M } ∪ {Ui }i∈I , and let {ρ} ∪ {ρi }i∈I be a subordinate partition of unity.
For each i ∈ I, consider the following local extension of f :
fei := f ◦ (ϕ−1 n ∞
i ◦ prm ◦ ϕi ) ∈ C (Ui ),
4.5. Exercises
Exercise 4.1. Construct a sequence of functions ρk ∈ C ∞ (R), k = 1, 2, . . ., such
P x ∈ R, ρk (x) 6= 0 only for a finite number of k’s, but such that the
that at every
function k≥1 ρk is not smooth.
Exercise 4.2. Let U = {Ui }i∈I and V = {Vj }j∈J be two open covers of M . Show
that W := {Ui ∩ Vj }(i,j)∈I×J is an open cover of M . If {ρi }i∈I is a partition of
unity subordinate to U and {χj }j∈J is a partition of unity subordinate to V, prove
that {ρi χj }(i,j)∈I×J is a partition of unity subordinate to W.
The locally finite cover found in Lemma 4.3.4 is at most countable. However,
this fact is automatic.
Exercise 4.3. Let U be a locally finite open cover of a manifold M .
(a) If M is compact, prove that U has a finite number of elements.
(b) In general, prove that U is at most countable.
Hint: choose a second cover V, as in Lemma 4.3.4, and show that every element
in V hits only a finite number of elements of U.
Exercise 4.4. Let {ρi }i∈I be a locally finite family of smooth functions on M .
(a) If M is compact, prove that only a finite number of functions are not identically
zero.
(b) In general, prove that the set of functions which are not identically zero is at
most countable.
LECTURE 5
Remark. For simplicity, from now on we will usually suppress the adjective smooth,
by implicitly assuming smoothness of the objects involved: by manifold we mean
smooth manifold, by chart we mean smooth chart, etc.
With the canonical identification of the tangent space of Rm , note that the
differential of a chart (U, ϕ) around p ∈ M becomes
dp ϕ : Tp M → Tϕ(p) Rm , dp ϕ(v) = vϕ .
The chain rule also extends to the setting of smooth manifolds, and its proof
is again a straightforward application of the ‘classical’ chain rule:
Lemma 5.1.7. Let f : M → N and g : N → P be smooth maps between smooth
manifolds. The differentials satisfy the chain rule:
dp (g ◦ f ) = df (p) g ◦ dp f, ∀ p ∈ M.
in other words the differential of f on the standard basis is given by the partial
derivatives of the local expression f ◦ ϕ−1 of f .
More generally, let f : M → N be a smooth map, and consider a smooth chart
(V, ψ) around q = f (p), with local coordinates ψ = (yψ1 , . . . , yψn ).The matrix of dp f
in the bases
( )
∂ ∂
: 1≤i≤m and : 1≤j≤n
∂xiϕ p ∂yψj q
of Tp M and Tq N , respectively, is the Jacobian matrix of the local representation
ψ ◦ f ◦ ϕ−1 (x) = (f 1 (x), . . . , f n (x))
of f in these charts:
∂f 1 ∂f 1 ∂f 1
∂x1 ∂x2 ... ∂xm
∂f 2 ∂f 2 ∂f 2
dp f =
∂x1 ∂x2 ... ∂xm (ϕ(p));
... ... ... ...
∂f n ∂f n ∂f n
∂x1 ∂x2 ... ∂xm
n
∂ X ∂f j ∂
dp f = (ϕ(p)) j .
i
∂xϕ p
∂x i
j=1 ∂yψ q
46 IOAN MĂRCUT
, , MANIFOLDS
Note that the space of derivations is a vector space: for all λ, µ ∈ R and
D1 , D2 ∈ Derp (C ∞ (M )),
λD1 + µD2 : C ∞ (M ) → R, (λD1 + µD2 )(f ) := λD1 (f ) + µD2 (f )
is also derivation of C ∞ (M ) at p.
The following result shows that all derivations come from derivatives along
tangent vectors.
Theorem 5.4.4. Let M be a smooth manifold, and let p ∈ M . The map
Tp M → Derp (C ∞ (M )), v 7→ Dv
is a linear isomorphism.
To prove the theorem, we first show that derivations of C ∞ (M ) at p are local,
in the sense that they depend only on the behavior of the functions around p. The
following terminology will be therefore rather useful:
Definition 5.4.5. Let M be a manifold and p ∈ M . Two functions f, g ∈ C ∞ (M )
are said to have the same germ at p, if there is an open neighborhood W of p such
that f |W = g|W .
Lemma 5.4.6. Let f, g ∈ C ∞ (M ) have the same germ at p ∈ M , then for any
D ∈ Derp (C ∞ (M )), we have that D(f ) = D(g).
Proof. Let h := f − g; then h|W = 0, for some neighborhood W of p. We show
that D(h) = 0. Consider a bump function χ ∈ C ∞ (M ) such that χ(p) = 1 and
supp(χ) ⊂ W . Then, χh = 0. Therefore:
0 = D(0) = D(χh) = h(p)D(χ) + χ(p)D(h) = D(h).
For the proof, we will need the following version of Taylor’s theorem:
48 IOAN MĂRCUT
, , MANIFOLDS
Proof. Fix x ∈ B, and consider the function h : [0, 1] → R, h(t) = f (tx). Clearly,
h is smooth, therefore the fundamental theorem of calculus applies:
Z 1
h(1) − h(0) = h0 (t)dt.
0
This equation gives:
Z 1 Z 1Xm m
d ∂f X
f (x) − f (0) = f (tx1 , . . . , txm )dt = i
(tx)xi
dt = gi (x)xi ,
0 dt 0 i=1 ∂x i=1
R 1 ∂f ∞ ∂f
where gi (x) = 0 ∂xi (tx)dt. Clearly, gi ∈ C (B) and gi (0) = ∂xi (0).
Proof of Theorem 5.4.4. Consider a chart (U, ϕ) on M around p, with coordi-
nates ϕ = (x1ϕ , . . . , xm
ϕ ) such that ϕ(p) = 0 and ϕ(U ) = B is convex.
By Lemma 5.4.7, it suffices to show that the map
Tp M → Derp (C ∞ (U )), v 7→ Dv
is an isomorphism.
Let v ∈ Tp M and denote vϕ = (v 1 , . . . , v m ). By the formulas in Section 5.2,
m
X ∂xi
Dv (xiϕ ) = dp xiϕ (v) = vj (0) = v i ;
j=1
∂xj
thus, the map is injective.
Let D ∈ Derp (C ∞ (U )). We show that D vanishes on constant functions. Since
D(1) = D(1 · 1) = 1 · D(1) + D(1) · 1 = 2 · D(1),
we obtain that D(1) = 0. By linearity, D(c) = D(c · 1) = c · D(1) = 0, for all c ∈ R.
Let v ∈ Tp M be the vector with vϕ = (v 1 , . . . , v m ) ∈ Rm , where v i := D(xiϕ ).
We show that D = Dv . Let f ∈ C ∞ (U ). Applying Lemma 5.4.8, we find functions
gi ∈ C ∞ (U ), for 1 ≤ i ≤ m, such that
Xm ∂
f = f (p) + xiϕ gi , gi (p) = dp f .
∂xi p
i=1 ϕ
Applying D, we obtain:
m
X
D(f ) = D(f (p)) + D(xiϕ gi ) =
i=1
m
X
0 · D(gi ) + D(xiϕ ) · gi (p) =
=0+
i=1
m
X ∂
= vi dp f = dp f (v) = Dv (f ).
i=1
∂xiϕ p
Thus, D = Dv .
LECTURE 5. 49
5.5. Exercises
Exercise 5.1. Let M ⊂ Rn be a manifold embedded in Rn , as in Definition 2.4.1,
and let p ∈ M . Let Vp ⊂ Rn consist of all vectors γ 0 (0) ∈ Rn , where γ : (−, ) → M
is a smooth curve such that γ(0) = p. Show that Vp is a linear subspace of Rn , and
that there is a linear isomorphism Vp − ∼→ Tp M , such that γ 0 (0) ∈ Vp corresponds
d
to d0 γ( dt |t=0 ) ∈ Tp M .
The following exercise gives an alternative definition of the tangent space:
Exercise 5.2. Let M be a manifold and let p ∈ M . Denote by Ip ⊂ C ∞ (M ) the
ideal of functions vanishing at p:
Ip = {f ∈ C ∞ (M ) : f (p) = 0}.
Denote by Ip2 the square of this ideal, i.e. Ip2 consists of finite sums of the form:
k
X
f= gj hj , gj , hj ∈ Ip , 1 ≤ j ≤ k.
j=1
for unique u0 , . . . , uk ∈ R; the algebraic operations are defined as usually for poly-
nomials. Consider the following algebra:
R[] := R[X]/(X 2 ),
where (X 2 ) ⊂ R[X] is the ideal generated by X 2 , i.e. each element p ∈ R[] can be
uniquely written as
p = u0 + u1 ,
where := X + (X 2 ) ∈ R[X]/(X 2 ); and the multiplication is such that 2 = 0:
(u0 + u1 ) · (v0 + v1 ) = u0 v0 + (u0 v1 + u1 v0 ).
Exercise 5.3. (a) Let s : A → R be an algebra homomorphism, and let D ∈
Ders (A, R). Prove that the map
s + D : A −→ R[], (s + D)(u) := s(u) + D(u)
is an algebra homomorphism.
(b) Prove that any algebra homomorphism σ : A → R[] is of the form described
at (a).
Exercise 5.4. Let s : A → R be a homomorphism of algebras.
(a) Prove that a homomorphism of algebras t : B → A induces a linear map
between the derivation spaces, defined as follows:
t∗s : Ders (A, R) −→ Ders◦t (B, R), t∗s (D) := D ◦ t.
(b) If t is surjective, prove that t∗s is injective.
(c) Prove that the following “chain rule” holds: if t : B → A and r : C → B are
algebra homomorphisms, then
(t ◦ r)∗s = rs◦t
∗
◦ t∗s : Ders (A, R) −→ Ders◦t◦r (C, R).
Exercise 5.5. Let X be a topological space, and let C(X) be the algebra of
continuous maps X → R. Consider a subalgebra A ⊂ C(X), with unit 1A the
constant map 1A (p) = 1, and which satisfies the condition:
(*) (f ∈ A, f (x) 6= 0, ∀ x ∈ X) =⇒ 1/f ∈ A.
We denote by A∨ the set of characters. For χ ∈ A∨ , let Iχ denote the kernel of χ:
Iχ := {f ∈ A : χ(f ) = 0}.
(a) Let p ∈ X. Prove that the following map gives a character of A:
χp : A −→ R, χp (f ) := f (p).
We denote Ip := Iχp .
(b) For χ ∈ A∨ , and p ∈ X, prove that Ip ⊂ Iχ , implies χ = χp .
(c) Prove that if f ∈ Iχ , then there exists p ∈ X such that f (p) = 0.
(d) Let χ ∈ A∨ be a general character, and assume that χ 6= χp for every p ∈ X.
Prove that for any point p ∈ X there exists f ∈ Iχ such that f (p) > 0.
(e) Prove that if X is compact, then every χ ∈ A∨ is of the form χ = χp for some
p ∈ X.
(f) Assume that there exists f ∈ A such that f −1 (λ) ⊂ X is compact for any
λ ∈ R. Show that every χ ∈ A∨ is of the form χ = χp for some p ∈ X.
LECTURE 5. 51
ψ ◦ f ◦ ϕ−1 = im n
n |ϕ(U ) : ϕ(U ) −→ R ,
ψ ◦ f ◦ ϕ−1 = prm n
n |ϕ(U ) : ϕ(U ) −→ R ,
The proofs of these two results are completely analogous, therefore, we only
give the details for the second result; the proof of the local immersion theorem is
left as an exercise.
ϕ := g −1 ◦ ϕ e−1 (O);
e : U −→ Rm , U := ϕ
e = (h, B) ◦ ϕ
hence, the charts (U, ϕ) and (V, ψ) satisfy the conditions from the statement.
Remark 6.2.4. In the local submersion theorem, the chart on the codomain (V, ψ)
can be chosen arbitrary; this follows from the proof. Similarly, in the local immer-
sion theorem the chart (U, ϕ) on the domain can be chosen arbitrary.
LECTURE 6. 55
h(q) = (xk+1 m
ϕ (q), . . . , xϕ (q)), for q ∈ U.
Thus, for q ∈ U : h(q) = 0 iff ϕ(q) ∈ Rk × {0}, and since U ∩ S = h−1 (0), we obtain:
ϕ(S ∩ U ) = (Rk × {0}) ∩ ϕ(U ). This concludes the proof.
Let us also mention a deep result of mathematical analysis, which states that
most points are regular values (note that points that are not in the image are always
regular values). Without introducing all notions involved, we state this result below
(for details, the reader is encouraged to consult [8] or [4]).
Theorem ♣ 6.3.7 (Sard’s Theorem). The set of critical values of a smooth map
f : M → N has Lebesgue measure zero. In particular, the set of regular values of f
is dense in N .
Remark 6.3.8. If dimM < dimN , then every point in M is a critical point of
f . In this situation, Sard’s Theorem says that f (M ) has measure zero in N ; and
therefore N \f (M ) is dense in N . In particular, one cannot have a surjective map
smooth f : M → N if dimM < dimN .
For this property, smoothness is essential. In the continuous setting, there are
the co-called space-filling curves (e.g. Peano’s curve [21]). For example, one can
prove that any connected, nonempty manifold is the image of a continuous curve!!
LECTURE 6. 57
f −1 (Q1 )
2
S
f −1 (Q2 )
p2
p
p0 Z p1
f (p2 )
f (p) Q1 Q2
R2
f (p0 ) f (Z) f (p1 )
6.5. Exercises
Exercise 6.1. Let k ≤ n, and let A : Rn → Rk be a surjective linear map, and let
C : Rk → Rn be an injective linear map.
(a) Prove that there exists a linear map B : Rn → Rn−k such that the map
(A, B) : Rn → Rk × Rn−k = ∼ Rn , (A, B)(x) := (A(x), B(x)),
is a linear isomorphism.
(b) Prove that there exists a linear isomorphism G : Rn → Rn such that
A ◦ G = prnk .
(c) Prove that there exists a linear map D : Rn−k → Rn such that the map
C + D : Rk × Rn−k ∼ = Rn → Rn , (C + D)(y, z) := C(y) + D(z),
is a linear isomorphism.
(d) Prove that there exists a linear isomorphism H : Rn → Rn such that
H ◦ C = ikn .
Exercise 6.2. Let a ∈ R. Show that the following map is an immersion:
fa : R −→ S 1 × S 1 , fa (t) := (eit , eita ).
√
Make a sketch of fa (R) for a = 0, 1/2, 1, 2, 3. For which a ∈ R is fa injective?
Exercise 6.3. Let f : P2 (R) → R3 be the map defined by
1
f ([x, y, z]) = 2 (yz, xz, xy).
x + y2 + z2
Show that f is smooth and show that it only fails to be an immersion at 6 points.
Make a sketch of the image of f .
Being an immersion/submersion is an open condition:
Exercise 6.4. If a smooth map f : M → N is an immersion (resp. a submersion)
at p ∈ M prove that there is an open neighborhood V ⊂ M of p such that f |V is
an immersion (resp. a submersion).
Do not use the local immersion/submersion theorem!
Exercise 6.5. Prove the local immersion theorem.
Exercise 6.6 (Alternative definition of immersions and submersions). Let f : M →
N be a smooth map, and let p ∈ M . Prove the following:
(a) f is an immersion at p iff it has a local left inverse, i.e. there exist open neigh-
borhoods U ⊂ M of p and V ⊂ N of f (p) and a smooth map σ : V → U such
that f (U ) ⊂ V and
σ(f (x)) = x, for all x ∈ U.
(b) f is a submersion at p iff it has a local right inverse, i.e. there exist open
neighborhoods U ⊂ M of p and V ⊂ N of f (p) and a smooth map j : V → U
such that j(f (p)) = p and
f (j(y)) = y, for all y ∈ V.
Exercise 6.7. Let f : M → N be a submersion. Prove that f is an open map (i.e.
for every open set U ⊂ M , f (U ) is open in N ).
60 IOAN MĂRCUT
, , MANIFOLDS
All these curves are immersed submanifolds in R2 ! Here are some of their properties:
a. The red line segment that Ca approaches is not part of Ca . An injective immer-
sion that parameterizes Ca (or a curve resembling Ca ) is:
endows Cb with the structure of an immersed submanifold. Note that the re-
sulting topologies are different! In fact, these are the only two differentiable
structures on Cb making Cb an immersed submanifold. Note that (0, 0) = j1 (0)
LECTURE 7. 63
and (0, 0) = j2 (π). At this point, the tangent spaces corresponding to the dif-
ferent smooth structures are different:
D ∂ ∂ E
T(0,0) Cb = d0 j1 (T0 R) = 2 (0,0) +
∂x ∂y (0,0)
D ∂ ∂ E
T(0,0) Cb = dπ j2 (Tπ R) = 2 (0,0) − .
∂x ∂y (0,0)
c. The curve Cc has 23 = 8 differentiable structures; at each intersection point there
are 2 options how to separate the curve. Also, the 8 topologies are different!
d. The curve Cd has a unique differentiable structure as an immersed submanifold;
but note that it is not an embedded submanifold!
e. The curve Ce is a closed version of the curve Ca ; the behavior at the two “ends”
is similar to that of the Topologist’s sine curve. This curve has a unique dif-
ferentiable structure as an immersed submanifold, but note that it is not an
embedded submanifold!
f. The curve Cf represents two curves that are “infinitely tangent”. There are two
differentiable structure on Cf which make it into an immersed submanifold. For
example, as a model one can take the following two decompositions of Cf :
Cf = {(x, 0) : x ∈ R} ∪ {(x, e−1/x ) : x > 0};
Cf = {(x, 0) : x > 0} ∪ {(x, f (x)) : x ∈ R},
where f is the smooth function: f (x) = 0, if x ≤ 0 and f (x) = e−1/x , if x > 0,
which give different decompositions of Cf into connected components. Note that
both smooth structures have the same tangent space at (0, 0):
D ∂ E
T(0,0) Cf = .
∂x (0,0)
The following is the main tool for comparing different immersed submanifolds
with the same underlying subset.
Lemma 7.1.5. Let f : N → M be an injective immersion. Consider a smooth
map ϕ : X → M such that ϕ(X) ⊂ f (N ), and denote by ϕ e : X → N the unique
map satisfying f ◦ ϕ
e = ϕ. If ϕ
e is continuous then it is smooth.
X
ϕ
ϕ
"
e
N
f
/M
7.3. Exercises
Exercise 7.1. If X is compact and Z is Hausdorff, prove that any continuous
bijection f : X → Z is a homeomorphism. Do not use Lemma 7.1.11!
Definition 7.3.1. An immersed submanifold S ⊂ M is called an initial subman-
ifold, if for every smooth map ϕ : X → M , such that ϕ(X) ⊂ S, we have that the
induced map ϕ e : X → S, ϕ(x)
e = ϕ(x) is smooth.
X
ϕ
ϕ
"
e
S
i /M
Exercise 7.2. Prove that embedded submanifolds are initial. Hint: Use Lemma
7.1.5.
Exercise 7.3. Prove that an initial submanifold has a unique differentiable struc-
ture such that the inclusion map is an immersion. Hint: Use Corollary 7.1.6.
Exercise 7.4. (a) Construct a smooth map χ : (−, 1 + ) → [0, 1] such that
χ|(−,0] = 0, χ|[1,1+) = 1, χ(0, 1) = (0, 1),
and χ : (0, 1) → (0, 1) is a diffeomorphism.
(b) Let M be a manifold and consider two smooth curves γ1 , γ2 : [0, 1] → M such
that γ1 (1) = γ2 (0). Show that the following curve is smooth:
γ1 (χ(t)), 0≤t≤1
γ : [0, 2] −→ M, γ(t) =
γ2 (χ(t − 1)), 1 ≤ t ≤ 2
Recall that a map from a closed interval f : [a, b] → M is smooth iff there is a
smooth map fe : (a − , b + ) → M , for some > 0, such that fe|[a,b] = f .
(c) Consider a subset in R2 which looks like the letter T :
S = {(0, t) : t ∈ R} ∪ {(t, 0) : t ∈ [0, ∞)}.
Prove that S is an immersed submanifold of R2 .
(d) Using item (b) prove that S is not an initial submanifold of R2 .
Exercise 7.5. Consider the six curves from Example 7.1.4. Show that Cb , Cc , Cd ,
and Cf are not initial submanifold. Explain why Ca and Ce are initial manifolds
(for Ce , you do not need to give a rigorous proof). Show that Ce is not an embedded
submanifold (note that Ce is a compact subset).
Exercise 7.6. Let M be a smooth manifold, and let X ⊂ M be a subset which is
at most countable. Prove that X is an initial submanifold of M . Prove that X is
an embedded submanifold iff it is discrete, i.e. for every point x ∈ X has an open
neighborhood U ⊂ M such that X ∩ U = {x}.
Exercise 7.7. For a ∈ R\Q, consider the map from Exercise 6.2:
fa : R −→ S 1 × S 1 , fa (t) := (eit , eita ).
Show that fa (R) is an initial submanifold of the 2-torus S 1 × S 1 .
LECTURE 7. 67
π −1 (U )
Φ / U × Rr pr1 ◦ Φ = π,
π pr1
U
idU
/U
where pr1 (x, v) = x, and such that for every p ∈ U , the restriction
pr2 ◦ Φ|Ep : Ep −→ Rr
is a linear isomorphism.
Here are some examples:
Example 8.2.2. The trivial vector bundle of rank r over a manifold M is
pr1 : M × Rr −→ M,
where the vector space structure on {p} × Rr is the obvious one.
Example 8.2.3. The tangent bundle T M of a manifold M is a vector bundle of
rank(T M ) = dim(M ).
LECTURE 8. 71
E
Φ /F πF ◦ Φ = ϕ ◦ πE
πE πF
M
ϕ
/N
The space of sections Γ(E) forms a module over the commutative algebra of
smooth functions C ∞ (M ), with the point-wise defined operations:
+ : Γ(E) × Γ(E) −→ Γ(E), (σ1 + σ2 )(p) := σ1 (p) + σ2 (p) ∈ Ep ,
· : C ∞ (M ) × Γ(E) −→ Γ(E), (f · σ)(p) := f (p)σ(p) ∈ Ep ,
where, on the right hand side, the vector space operations on Ep are used. The
zero-section plays the role of the zero-element of Γ(E).
Consider the case of a trivial bundle M × Rr → M . Its sections are smooth
maps of the form:
σ : M −→ M × Rr , σ(p) = (p, (f1 (p), . . . , fr (p))) ∈ M × Rr .
Thus, we have a one-to-one correspondence between Γ(M × Rr ) and C ∞ (M ; Rr ).
Using the structure of a C ∞ (M )-module of Γ(M × Rr ), we can decompose any
section σ uniquely as
σ = f1 σ1 + . . . + fr σr ,
where, for 1 ≤ i ≤ r, σi ∈ Γ(M × Rr ) represents the section
σi (p) := (p, (0, . . . , 1, . . . , 0)) ∈ M × Rr ,
with 1 is on the i-th position. In other words:
Γ(M × Rr ) is a free C ∞ (M )-module with basis σ1 , . . . , σr .
Since a general vector bundle π : E → M is locally trivializable, there is a cover
of M by open sets U such that Γ(E|U ) is a free C ∞ (U )-module of rank r.
8.4. Frames
Definition 8.4.1. Let π : E → M be a vector bundle. A local frame over U ⊂ M
is a smooth map
γ : E|U −→ Rr ,
such that, for all p ∈ U , γ|Ep : Ep → Rr is a linear isomorphism.
Lemma 8.4.2. Let π : E → M be a vector bundle, and let U ⊂ M be an open set.
There is a one-to-one correspondence between the following objects:
(1) local trivializations Φ : E|U − ∼→ U × Rr ;
r
(2) local frames γ : E|U → R ;
(3) sections σ1 , . . . , σr ∈ Γ(E|U ) such that, for every p ∈ U , σ1 (p), . . . , σr (p) forms
a basis of Ep .
This correspondence is such that, given a local frame γ, the corresponding local
trivialization is Φ := π × γ : E|U → U × Rr , and the corresponding set of sections
σ1 , . . ., σr is given by σi (p) := γp−1 (ei ), where e1 , . . . , er is the standard basis of Rr .
Proof. By passing to the restriction E|U , we may assume that U = M .
First, we explain the equivalence between (a) and (b). Consider a trivialization
Φ:E− ∼
→ M × Rr . Then, by the definition of Φ, we have that Φ = π × γ, where
γ : E → Rr is a frame. Conversely, let γ : E → Rr be a frame on E over M . We
need to show that Φ = π × γ : E → M × Rr is a diffeomorphism. Since Φ is a
bijection, and the manifolds have the same dimension, invoking the inverse function
theorem, it suffices to show that π × γ is an immersion. Let v ∈ Te E be such that
0 = de (π × γ) (v) = (de π(v), de γ(v)).
LECTURE 8. 73
You can check that this is a consequence of the following classical theorem:
Theorem ♣ 8.5.8. On the sphere S n the maximum number k = k(n) of vector
fields X1 , . . . , Xk such that at every p ∈ S n the vectors X1,p , . . . , Xk,p ∈ Tp S n are
linearly independent is given as follows: write n + 1 = (2a + 1)24b+c , with a, b ∈ N
and c ∈ {0, 1, 2, 3}, then k(n) = 8b + 2c − 1.
If n is even then k(n) = 0, so every vector field on S n has a zero. In general,
note that k(n) is quite small compared to n; in particular, k(n) ≤ 2 log2 (n + 1) + 1.
8.6. Exercises
Exercise 8.1. Prove that T S 1 is isomorphic to the trivial vector bundle S 1 × R.
Exercise 8.2. Let f : M → N be a smooth map. Prove that df : T M → T N is a
smooth vector bundle morphism.
Exercise 8.3. Let M be a smooth manifold. Prove that there exists a vector field
V on the manifold T M (i.e. V ∈ X(T M ) = Γ T (T M ) ) satisfying:
dv π(Vv ) = v, ∀ v ∈ T M.
Hint: construct V locally, and use a partition of unity.
Exercise 8.4. Recall that the m-dimensional real projective space is the space of
all lines through the origin in Rm+1 , with the manifold structure from Example
3.2.1:
Pm (R) := {l : l ⊂ Rm+1 is a line with 0 ∈ l}.
The tautological line bundle over Pm (R) is the rank 1 vector bundle π : L(m) →
Pm (R) whose fiber over a line l is the line l:
L(m) := {(l, v) : l ∈ Pm (R), v ∈ l}.
(a) Sketch a picture of L(1).
(b) Show that π : L(m) → Pm (R) is indeed a vector bundle of rank one, such that
the following map is morphism of vector bundles:
L(m)
Φ / Pm (R) × Rm+1 , Φ(l, v) := (l, v) ∈ Pm (R) × Rm+1 .
pr
π
1
m
P (R)
id / Pm (R)
L(1)
Ψ / L(m) ,
π
pr
1
P1 (R) / Pm (R)
= X 1 ϕ∗ (Y 2 (f )) − X 2 ϕ∗ (Y 1 (f )) =
= X 1 ◦ X 2 (ϕ∗ f ) − X 2 ◦ X 1 (ϕ∗ f ) =
= [X 1 , X 2 ](ϕ∗ f ).
Vector fields can always be transported along diffeomorphisms.
Definition 9.3.4. The pushforward of a vector field X ∈ X(M ) via a diffeomor-
∼
phism ϕ : M −→ N is the vector field
ϕ∗ X = dϕ ◦ X ◦ ϕ−1 ∈ X(N ),
more explicitly:
(ϕ∗ X)q := dp ϕ(Xp ), where p = ϕ−1 (q).
Note that ϕ∗ X is the unique vector field on N such that X and ϕ∗ X are ϕ-related.
The pullback of a vector field Y ∈ X(N ) via the diffeomorphism ϕ : M −∼
→N
is the vector field
ϕ∗ Y := (ϕ−1 )∗ Y = (dϕ)−1 ◦ Y ◦ ϕ ∈ X(M ).
Note that ϕ∗ Y is the unique vector field on M such that ϕ∗ Y and Y are ϕ-related.
Example 9.3.5. Let ι : M ,→ N denote the inclusion of the embedded submanifold
M of N . Then X ∈ X(M ) is ι-related to Y ∈ X(N ) iff
Yp = Xp ∈ Tp M ⊂ Tp N, ∀ p ∈ M
For example, consider ι : S 1 ,→ R2 , and the following vector field on R2
∂ ∂
Y := x −y ∈ X(R2 ).
∂y ∂x
First, note that Y is tangent to S 1 because, for (u, v) ∈ S 1 , we have that
n ∂ ∂ o
T(u,v) S 1 := ker(d(u,v) (x2 + y 2 − 1)) = a (u,v) + b (u,v) : ua + bv = 0 .
∂x ∂y
Consider the “angle coordinate” θ ∈ R/(2πZ) on S 1 (see Example 8.5.4),
θ 7→ (x(θ), y(θ)) := (cos(θ), sin(θ)).
LECTURE 9. 81
∂ ∂
We claim that Y is ι-related to X := ∂θ , i.e. Y |S 1 = ∂θ . We will use Lemma 9.3.2.
For any f ∈ C ∞ (R2 ), we have that
∂f ∂f
Y (f ) = x (x, y) − y (x, y),
∂y ∂x
and so
∂f ∂f
ι∗ (Y (f )) = cos(θ) (cos(θ), sin(θ)) − sin(θ) (cos(θ), sin(θ)).
∂y ∂x
On the other hand, by the chain rule, we have that:
∂
X(ι∗ f ) = (f (cos(θ), sin(θ))) =
∂θ
∂f ∂ cos(θ) ∂f ∂ sin(θ)
= (cos(θ), sin(θ)) + (cos(θ), sin(θ)) =
∂x ∂θ ∂y ∂θ
∂f ∂f
= cos(θ) (cos(θ), sin(θ)) − sin(θ) (cos(θ), sin(θ)).
∂y ∂x
Consider a second chart (V, ψ = (yψ1 , . . . , yψm )) on M , and denote the corre-
sponding local expression of X by
m
X ∂
X|V = Y i i , Y i ∈ C ∞ (V ).
i=1
∂yψ
j
m
X
i
∂yψj
Y = X .
i=1
∂xiϕ
This expression represents the rule of change of coordinates for vector fields, i.e. it
gives the local expression of X in the chart (V, ψ) in terms of the local expression
of X in the chart (U, ϕ).
9.6. Exercises
Exercise 9.1. Consider the polar coordinates (U, r, θ) on R2 , where
U = R2 \ (−∞, 0] × {0} , x = r cos(θ), y = r sin(θ), (r, θ) ∈ (0, ∞) × (−π, π).
We have that
Theorem 10.1.5. Let X be a vector field on M . Then the domain D(X) of X is
open in R × M and the flow of X is a smooth map
φX : D(X) → M, (t, p) 7→ φtX (p).
Proof. Theorem 9.5.2 implies that for every point p ∈ M there exists an open
neighborhood W ⊂ M and there exists > 0 such that (−, ) × W ⊂ D(X), and
φX is smooth on (−, ) × W .
Let (t, p) ∈ D(X). Denote the image of the integral curve from p and φtX (p) by
C := {φsX (p) : s ∈ [0, t] (or s ∈ [t, 0])}.
Since C is compact, the argument above implies that there is some open set C ⊂
W ⊂ M and some > 0 such that (−, ) × W ⊂ D(X) and φX is smooth on
t/n
(−, ) × W . Let n > 0 be such that t/n ∈ (−, ); then φX : W → M is smooth.
Consider the open sets {Wi }ni=0 defined inductively by
−1
t/n
Wn := W, Wi := φX |W (Wi+1 ), 0 ≤ i ≤ n − 1.
LECTURE 10. 87
Taking the derivative of this equality, for Z = Y , and using the first part, we obtain:
d d
((φtX )∗ Y )g = (φtX )∗ ◦ Y ◦ (φ−t ∗
X ) g =
dt dt
= (φtX )∗ ◦ (X ◦ Y − Y ◦ X) ◦ (φ−t ∗
X ) g =
= (φtX )∗ ◦ [X, Y ] ◦ (φ−t ∗ t ∗
X ) g = ((φX ) [X, Y ])g,
where in the last step we have used again (∗) for Z = [X, Y ]. Since, g is arbitrary, we
obtain the third equality dt d
(φtX )∗ Y = (φtX )∗ [X, Y ]. Since [X, X] = 0, this implies
that (φtX )∗ X = X. Next, note that the vector fields (φtX )∗ X = X, (φtX )∗ Y and
(φtX )∗ [X, Y ] are φtX -related to the vector fields X, Y and [X, Y ], respectively. Thus,
by Lemma 9.3.3, and by the fact that φtX is a local diffeomorphism, (φtX )∗ [X, Y ] =
[X, (φtX )∗ Y ]. This implies also the last equality.
σp (t, s) := φ−s −t s t
Y ◦ φX ◦ φY ◦ φX (p) ∈ M,
m
n
p
q
p0
The black and red curves represent the flow lines of X and Y , respectively, and
In general, the curved parallelogram does not close up, i.e. p0 = σp (t, s) 6= p.
90 IOAN MĂRCUT
, , MANIFOLDS
∂
On the other hand, since σ(t, 0) = idM , the derivative t 7→ ∂s σp (t, 0) is a
smooth vector field on M , which is given by:
d d −s
φY ◦ φ−t s t
σp (t, s)s=0 = X ◦ φY ◦ φX (p) s=0 =
ds ds
= −Yp + (dφ−t
X )(YφtX (p) ) =
t ∗
= (φX ) Y − Y p .
Applying Proposition 10.3.3, we obtain the following description of the Lie bracket:
Proposition 10.4.1. With the notation from above, we have that:
d d
t=0
σ(t, s) = [X, Y ].
dt ds s=0
10.5. Exercises
Exercise 10.1. Let ψ : M → N be a smooth map. Consider X ∈ X(M ) which is
ψ-related to Y ∈ X(N ). Prove that their flows are ψ-related in the following sense:
for every p ∈ M , Ip ⊂ Iψ(p) , and for all t ∈ Ip we have that
Exercise 10.2. Let X ∈ X(M ). For λ ∈ R, prove that the flow of λX is given by:
φtλX (p) = φλt
X (p),
(a) Show that there exists a smooth vector field X on S m which in the coordinates
i
y+ is given by
m
i ∂
X
X|U+ = y+ i
∈ X(Rm ),
i=1
∂y +
i
and write the expression of X in the coordinates y− .
(b) Determine the flow of X in both charts.
e ∈ X(Rm+1 ) satisfying
(c) Prove that there exists a vector field X
X
ep = Xp , ∀ p ∈ Sm.
∂
Exercise 10.5. (a) For any vector field X = f (x) ∂x ∈ X(R), prove that the vector
∂
field sin(x)f (x) ∂x ∈ X(R) is complete (Hint: use Lemma 10.2.3).
(b) Prove that any vector field on R can be written as the sum of two complete
vector fields on R (Hint: use part (a)).
Pm ∂
Exercise 10.6. Let X = i=1 X i (x) ∂x i ∈ X(R
m
) be a bounded vector field, i.e.
there is a constant C > 0 such that |X (x)| < C for all x ∈ Rm and all 1 ≤ i ≤ m.
i
where C ∞ (M )[[T ]] denotes the ring of formal power series in T with coefficients in
C ∞ (M ).
Exercise 10.8. Let A be a real vector space endowed with a bilinear map
∗ : A × A −→ A.
We are not assuming any other axiom for the operation ∗ (like associativity, com-
mutativity, Jacobi identity etc.).
An automorphism of (A, ∗) is a linear isomorphism g : A → A satisfying
g(a ∗ b) = g(a) ∗ g(b).
Denote the space of all automorphisms by Aut(A, ∗).
A derivation of (A, ∗) is a linear map D : A → A satisfying:
D(a ∗ b) = D(a) ∗ b + a ∗ D(b), for all a, b ∈ A.
Denote the space of all derivations by Der(A, ∗).
(a) Show that (Aut(A, ∗), ◦) is a group.
(b) Show that, if D1 , D2 ∈ Der(A, ∗), then
[D1 , D2 ] := D1 ◦ D2 − D2 ◦ D1 ∈ Der(A, ∗).
(c) Prove that the operation [·, ·] on Der(A, ∗) satisfies the condition (1),(2) and
(3) of Proposition 9.2.1; i.e. (Der(A, ∗), [·, ·]) is a Lie algebra.
92 IOAN MĂRCUT
, , MANIFOLDS
For (d) and (e), assume A to be finite dimensional1 Hence, we can talk about
smooth curves g : (−, ) → LinR (A, A), where LinR (A, A) is the (finite dimensional)
vector space of linear maps from A to A. Identifying Tg(t) LinR (A, A) ∼= LinR (A, A),
we regard dg
dt (t) ∈ LinR (A, A).
(d) Consider a smooth curve g : (−, ) → LinR (A, A), such that g(0) = idA and
g(t) ∈ Aut(A, ∗), for all t ∈ (−, ).
dg
Prove that ∈ Der(A, ∗).
dt (0)
(e) Let D ∈ Der(A, ∗). Show that the following series converges absolutely to an
automorphism of (A, ∗)
X 1
exp(D) = Dn ∈ Aut(A, ∗).
n!
n≥0
For (f) and (g) assume (A, ∗) to be associative and commutative, however,
A can be infinite dimensional.
(f) For a ∈ A and D ∈ Der(A, ∗), prove that a ∗ D ∈ Der(A, ∗), where a ∗ D is
defined in the obvious way:
(a ∗ D)(b) := a ∗ (D(b)).
(g) Prove that the following Leibniz rule holds:
[D1 , a ∗ D2 ] = D1 (a) ∗ D2 + a ∗ [D1 , D2 ].
1This assumption simplifies the discussion; however, the results hold also in an infinite dimensional
setting, e.g. for bounded operators on a Banach space.
LECTURE 11
ι : G −→ G, g 7→ g −1 .
A Lie subgroup of G is a subgroup H ⊂ G which is an immersed submanifold.
Examples 11.1.2. (1) Every at most countable group is a 0-dimensional Lie group.
(2) The additive group (Rn , +) is an n-dimensional Lie group.
(3) The multiplicative group (C∗ , ·) is a 2-dimensional Lie group. The following
are 1-dimensional subgroups R∗ ⊂ C∗ and S 1 ⊂ C∗ .
(4) The general linear group. The group GL(n) of linear automorphisms of
Rn is an n2 -dimensional Lie group. Its smooth structure comes by regarding
GL(n) as an open subset in the n2 -dimensional vector space M (n) of linear
maps from Rn to Rn .
(5) The special linear group. The group SL(n) is the subgroup of GL(n) con-
sists of linear automorphisms of Rn which have determinant 1.
(6) The orthogonal group. The group O(n) of orthogonal matrices (i.e. A ∈
M (n) such that AAt = I) is a compact Lie group of dimension n(n − 1)/2 (see
Exercise 6.11).
(7) There are complex versions of the groups above: GL(n, C) the group of in-
vertible complex matrices, with subgroup SL(n, C) consisting of matices with
determinant 1; O(n, C) the group of complex matrices satisfying AAt = I.
(8) The product G1 × . . . × Gk of Lie groups is a Lie group. In particular, the
k-torus T k := S 1 × . . . × S 1 is a k-dimensional compact Lie group.
(9) If G is any Lie group, then the connected component of the identity, denoted
by G◦ ⊂ G, is a Lie subgroup (which is both open and closed). In fact, G◦ is
a normal subgroup.
(10) The group GL(n)◦ consists of matrices with positive determinant.
(11) The special orthogonal group is the Lie group SO(n) := O(n)◦ .
(12) If G is a Lie group, then T G is a Lie group. Its structure maps are the differ-
entials of the structure maps of G.
93
94 IOAN MĂRCUT
, , MANIFOLDS
Definition 11.1.8. The Lie algebra of a Lie group G is the vector space
g := Te G
with Lie bracket:
[·, ·] : g × g → g, [v, w] := [v L , wL ]e .
By Proposition 9.2.1, (g, [·, ·]) satisfies indeed the axioms of a Lie algebra.
Remarks 11.1.9. (1) The Lie algebra g of a Lie group G plays the role of a
“linearized version” of G; it encodes infinitesimal information about G. In fact,
if G is simply connected, it can be fully recovered as a Lie group from the
structure of g (see for example [2]).
(2) In general Lie groups are denoted by capital Latin letters: G, H, K, L, P ,
Q, etc., while the corresponding Lie algebras are denoted by the corresponding
small Fraktur letters (also called Gothic letters): g, h, k, l, p, q etc. (in LaTeX
use \mathf rak{..}).
(3) One can define the Lie algebra of G also using right invariant vector fields;
nevertheless, one obtains an isomorphic object. The push-forward along the
inversion map gives an isomorphism between left invariant and right invariant
vector fields. Note however that, if both Lie algebras are identified with Te G
(by evaluating vector fields at e), one obtains opposite Lie brackets on Te G,
because the differential of the inversion map at e is −IdTe G .
Proposition 11.1.10. Left invariant vector fields are complete. Moreover,
φtX (gh) = gφtX (h), for all X ∈ XL (G), g, h ∈ G, t ∈ R.
Proof. Let X ∈ XL (G) and let g, h ∈ G. As in Lecture 10, let Ih denote the
maximal interval on which the flow line starting at h is defined. For t ∈ Ih , we
have:
d d
gφtX (h) = dφtX (h) λg φtX (h) = dφtX (h) λg XφtX (h) = XgφtX (h) ,
dt dt
where we have used that X is left invariant. This shows that
Ih 3 t 7→ gφtX (h)
is an integral curve of X. Since the curve starts at gh, by Lemma 10.1.2, we
conclude that Ih ⊂ Igh and that gφtX (h) = φtX (gh). In particular, Ie ⊂ Ig for all
g ∈ G, and so, by Lemma 10.2.2, X is complete.
The Proposition above allows us to define:
Definition 11.1.11. The exponential map of a Lie group G with Lie algebra g
is the map:
exp : g −→ G, v 7→ φ1vL (e).
Exercise 11.1 guides you through a possible proof that the exponential map is
indeed smooth. In fact, the exponential map is a local diffeomorphism around 0.
Proposition 11.1.12. The exponential map exp : g → G satisfies:
d0 exp = Idg ,
where we identify T0 g = g = Te G. Hence, exp is a diffeomorphism between a
neighborhood of 0 in g and a neighborhood of e in G.
96 IOAN MĂRCUT
, , MANIFOLDS
Proof. The result follows from the formula φtsvL (g) = φst
v L (g) of Exercise 10.2:
d d d
d0 exp(v) = =0
exp(v) = =0 φ1vL (e) = =0 φvL (e) = veL = v.
d d d
GO
F /H
O
exp exp
g
f
/h
So, for any w ∈ g, wL and f (w)L are also F -related, and so, by Lemma 9.3.3,
[v L , wL ] is F -related to [f (v)L , f (w)L ]. Evaluating this identity at the unit, implies
that f is a Lie algebra homomorphism:
[f (v), f (w)] = [f (v)L , f (w)L ]e = de F ([v L , wL ]e ) = f ([v, w]).
Finally, using again that v L is F -related to f (v)L , and Exercise 10.1, we obtain the
commutativity of the diagram:
F (exp(v)) = F (φ1vL (e)) = φ1f (v)L (F (e)) = exp(f (v)).
LECTURE 11. 97
11.2. Exercises
Exercise 11.1. Let G be a Lie group with Lie algebra g. Denote by v L ∈ XL (G)
the left invariant extension of a vector v ∈ g.
(a) Show that the following formula defines a smooth vector field on G × g:
ξ ∈ X(G × g), ξ(g,v) := (vgL , 0) ∈ Tg G ⊕ Tv g ∼
= T(g,v) (G × g).
(b) Prove that the flow of ξ is given by
φtξ (g, v) = (φtvL (g), v),
and conclude that ξ is complete.
(c) Use (b) to prove that the exponential map exp : g → G is smooth.
Exercise 11.2. Prove that left invariant vector fields commute with right invariant
vector fields:
X ∈ XL (G), Y ∈ XR (G) =⇒ [X, Y ] = 0.
Exercise 11.3. For k ≥ 1, denote by T k the k-dimensional torus, i.e.
T k = (S 1 )k = {(z1 , . . . , zk ) ∈ Ck : |z1 | = . . . = |zk | = 1}.
(a) Prove that left invariant vector fields on the k-torus are the same as right
invariant vector fields:
XL (T k ) = XR (T k ).
Moreover, if θi , 1 ≤ i ≤ k denote the “angle coordinates” on T k (see Example
8.5.5), prove that the following is a basis of XL (T k ):
∂ ∂
,..., ∈ X(T k ).
∂θ1 ∂θk
(b) Calculate the exponential map:
exp : Te T k −→ T k .
Hint: use the basis from (a) for Te T k .
(c) Consider an n × m matrix with integer coefficients A = {ai,j ∈ Z}1≤i≤n,1≤j≤m .
Show that the map
FA : (T m , ·) −→ (T n , ·),
a a1,m a an,m
FA (z1 , . . . , zm ) = (z1 1,1 . . . zm , . . . , z1 n,1 . . . zm )
is a Lie group homomorphism.
(d) Show that any Lie group homomorphism F : T m → T n is of the form described
in (c). Hint: use Proposition 11.1.14.
Exercise 11.4. Consider the Lie group GL(n) of invertible n×n-matrices. Denote
the space of all n × n-matrices by gl(n). Since GL(n) is an open subset of gl(n),
we identify:
Tg GL(n) = gl(n), for all g ∈ GL(n).
(a) Prove that, with the above identification, the differential of left translation is
simply given by matrix multiplication:
dh λg : Th GL(n) = gl(n) −→ Tgh GL(n) = gl(n),
dh λg (v) = g · v.
98 IOAN MĂRCUT
, , MANIFOLDS
(b) Prove that the left invariant extension of v ∈ TI GL(n) = gl(n) is the vector
field
v L ∈ XL (GL(n)), vgL = g · v ∈ Tg GL(n) = gl(n).
(c) Prove that the flow of the left invariant extension v L ∈ XL (GL(n)) of v ∈
TI GL(n) = gl(n) is given by
φtvL (g) = g · etv ,
n
where ew = n≥0 wn! is the matrix exponential.
P
(d) Using Proposition 10.4.1 and (c) prove that the Lie bracket on gl(n) is the
commutator of matrices:
[v, w] = v · w − w · v, v, w ∈ gl(n).
Exercise 11.5. The quaternions are the real 4-dimensional associative algebra
H with basis 1, i, j, k, where the multiplication is determined by the rules:
i2 = j 2 = k 2 = −1, ij = k, jk = i, ki = j, ji = −k, kj = −i, ik = −j.
One can also represent H as the space of all 2 × 2-matrices with complex entries of
the form:
z w
, z, w ∈ C.
−w z
The correspondence between the two descriptions is given by
i 0 0 1 0 i
i ←→ , j ←→ , k ←→ .
0 −i −1 0 i 0
(a) Prove that every element in H∗ = H\{0} is invertible, and conclude the H∗ is
a 4-dimensional Lie group.
(b) Prove that the map following map is a Lie group homomorphism:
F : (H∗ , ·) −→ (R>0 , ·), F (a + bi + cj + dk) = a2 + b2 + c2 + d2 .
(c) Prove that G := F −1 (1) is a Lie subgroup of H∗ which as a manifold is diffeo-
morphic to S 3 .
(d) Identifying T1 H∗ = H, prove that the exponential map of H∗ is given by the
usual exponential series:
X An
exp : H −→ H∗ , A 7→ .
n!
n≥0
(j) On g we consider the metric such that Σ is the unit sphere around the origin,
and denote by Br ⊂ g the open ball of radius r > 0. Using the previous items,
prove the following about the exponential map exp : g → G:
(i) exp is a diffeomorphism between Bπ and G\{−1};
(ii) exp sends the spheres of radii 2kπ to 1 ∈ G, and the spheres of radii
(2k + 1)π to −1 ∈ G;
(iii) exp restricts to a diffeomorphism between B(k+1)π \B kπ and G\{1, −1}.
(k) Show that every flow line of a left invariant vector field on G is a circle.
LECTURE 12
Vk ∗
forms a basis of V . Hence
k n = n!
k!(n−k)! , 0 ≤ k ≤ n
^
∗ k
dim V =
0, n < k.
Proof. First we prove (1). Note that the set of (k, l)-shuffles admits the disjoint
decomposition S(k, l) = S 0 (k, l) t S 00 (k, l), where
S 0 (k, l) = {σ ∈ S(k, l) : σ(1) = 1} S 00 (k, l) = {σ ∈ S(k, l) : σ(k + 1) = 1}.
Therefore,
(ιv1 (α ∧ β))(v2 , . . . , vk+l ) = (α ∧ β)(v1 , . . . , vk+l ) =
X
= sign(σ)α(vσ(1) , . . . , vσ(k) )β(vσ(k+1) , . . . , vσ(k+l) ) =
σ∈S(k,l)
X
= sign(σ 0 )(ιv1 α)(vσ0 (2) , . . . , vσ0 (k) )β(vσ0 (k+1) , . . . , vσ0 (k+l) )+
σ 0 ∈S 0 (k,l)
X
+ sign(σ 00 )α(vσ00 (1) , . . . , vσ00 (k) )(ιv1 β)(vσ00 (k+2) , . . . , vσ00 (k+l) ).
σ 00 ∈S 00 (k,l)
and moreover, that sign(σ 0 ) = sign(θ) and sign(σ 00 ) = (−1)k sign(η). Using this
fact, the above formula becomes:
(ιv1 (α ∧ β))(v2 , . . . , vk+l ) =
X
= sign(θ)(ιv1 α)(vθ(1)+1 , . . . , vθ(k)+1 )β(vθ(k+1)+1 , . . . , vθ(k+l)+1 )+
θ∈S(k−1,l)
X
(−1)k sign(η)α(vη(1)+1 , . . . , vη(k)+1 )(ιv1 β)(vη(k+1)+1 , . . . , vη(k+l−1)+1 ) =
η∈S(k,l−1)
= (−1)kl ιv (β ∧ α),
where we have used the induction hypothesis in the second line. Since this holds
for all v, (2) follows.
For the last part, let e1 , . . . , en ∈ V be the dual basis, i.e. these elements are
defined by ei (ej ) = δji , where δji = 1 if i = j and δji = 0 if i 6= j. By repeatedly
applying the derivation rule, one can easily show that, for 1 ≤ i1 < . . . < ik ≤ n
and 1 ≤ j1 < . . . < jk ≤ n, the following holds:
ei1 ∧ . . . ∧ eik (ej1 , . . . , ejk ) = δji11 . . . δjikk ,
LECTURE 12. 105
in other words the result is zero unless ip = jp , for all 1 ≤ p ≤ k, and then the
result is 1.
Consider now a linear combination:
X k
^
ω= ai1 ...ik ei1 ∧ . . . ∧ eik ∈ V ∗.
i1 <...<ik
Then for 1 ≤ j1 < . . . < jk ≤ n, we have that ω(ej1 , . . . , ejk ) = aj1 ...jk . Thus if
ω = 0, then all coefficients aj1 ...jk = 0. This shows that the k-forms ei1 ∧ . . . ∧ eik
are linearly independent.
Vk ∗ Vk ∗
To show that these forms span V , let ω ∈ V . Consider the k-form:
X k
^
η := ω − ω(ei1 , . . . , eik )ei1 ∧ . . . ∧ eik ∈ V ∗.
i1 <...<ik
We conclude that the set {ei1 ∧ . . . ∧ eik : 1 ≤ i1 < . . . < ik ≤ n} forms a basis of
Vk ∗
V . Note that sequences of the form 1 ≤ i1 < . . . < ik ≤ n are in one-to-one
correspondence with subset I = {i1 , . . . , ik } of {1, . . . , n} with exactly k elements;
Vk
this implies that dim( V ∗ ) = nk .
As a consequence of the graded commutativity, we obtain:
Vk ∗
Corollary 12.1.6. If k is odd and ω ∈ V then ω 2 = ω ∧ ω = 0.
Putting the exterior powers of V ∗ together, one obtains the exterior algebra:
Definition
V ∗ 12.1.7. Let V be a vector space. The exterior algebra of V ∗ , denoted
∗
by ( V , +, ∧), is the direct sum of the exterior products of V
^ k
M^
V ∗ := V ∗,
k≥0
j
^
β = β 0 + β 1 + . . . + βl , βj ∈ V ∗,
then their wedge product is given by:
X s
^
α ∧ β = γ0 + γ1 + . . . + γk+l , γs := αi ∧ βj ∈ V ∗.
i+j=s
defined by
k
^ k
^
A∗ : W ∗ −→ V ∗ , (A∗ ω)(v1 , . . . , vk ) := ω(Av1 , . . . , Avk ),
V0 V0
for k ≥ 1, and for a ∈ W ∗ = R, A∗ a = a ∈ V ∗ = R.
Remark 12.1.11. Note that (k, l)-shuffles are in one-to-one correspondence with
pairs of subsets (A, B) of the set {1, . . . , k + l}, such that:
{1, . . . , k + l} = A ∪ B, |A| = k, |B| = l.
The element σ ∈ S(k, l) corresponds to the pair
A = {σ(1), . . . , σ(k)}, B = {σ(k + 1), . . . , σ(k + l)}.
Conversely, given such a pair (A, B), one first orders the elements of the two sets:
A = {a1 < a2 < . . . < ak }, B = {b1 < b2 < . . . < bl },
and then the corresponding σ is given by:
σ(i) := ai , 1 ≤ i ≤ k; σ(k + j) := bj , 1 ≤ j ≤ l.
In particular, we have that
k+l (k + l)!
|S(k, l)| = = .
k k! l!
LECTURE 12. 107
1From this lecture on, we use simplified notation: instead of xi we denote the coordinates by xi .
ϕ
108 IOAN MĂRCUT
, , MANIFOLDS
where the sum runs over all 1 ≤ l1 , . . . , lk ≤ m. After ordering the indices l1 , . . . , lk
increasingly, and dropping the terms where some indexes are repeated (are therefore
are zero), we can rewrite:
X
dxi1 ∧ . . . ∧ dxik = cji11,...,i j1 jk
,...,jk dy ∧ . . . ∧ dy ,
k
j1 <...<jk
where the notation on the right is the standard notation for the minor consisting of
∂xi m
the rows i1 , . . . , ik and columns j1 , . . . , jk in the Jacobian matrix ( ∂y j )i,j=1 of the
with smooth coefficients ωi1 ...ik ∈ C ∞ (U ). The notation where the indexes are not
ordered is also useful:
X 1
ω|U = ωi1 ...ik dxi1 ∧ . . . ∧ dxik ,
i ...i
k!
1 k
X 1
(ιX ω)|U = X i1 ωi1 ...ik dxi2 ∧ . . . ∧ dxik .
i1 ...ik
(k − 1)!
12.6. Exercises
Exercise 12.1. Prove the following formulas for the wedge product:
1 X
(α ∧ β)(v1 , . . . , vk+l ) = sign(σ)α(vσ(1) , . . . , vσ(k) )β(vσ(k+1) , . . . , vσ(k+l) ).
k!l!
σ∈Sk+l
dimension n. An element α ∈ V ∗
V
Exercise 12.8. Let V be a vector space V of
is called invertible if there exists β ∈ V ∗ such that α ∧ β = 1. Prove that an
element
^ i
^
α = α0 + α1 + . . . + αn ∈ V ∗ , αi ∈ V∗
1
is invertible iff α0 6= 0 (Hint: think about the geometric series 1−q = 1+q +q 2 +. . .).
Exercise 12.9. Consider the differential forms, the vector field and the map below:
2
α = cos(y)dx + dz − e−w dw ∈ Ω1 (R4 ),
β = x2 dx ∧ dy + w7 dw ∧ dz ∈ Ω2 (R4 ),
γ = w2 dz + dz ∧ dy ∧ (dw + ydw ∧ dx) ∈ Ω(R4 ),
∂ ∂ ∂
X=x −y + y3 ,
∂y ∂z ∂w
f : R2 → R4 , f (u, v) = (u2 , euv , u − v 4 , sin(v)).
(a) Calculate the nine exterior products:
α∧α α∧β α∧γ
β∧α β∧β β∧γ .
γ∧α γ∧β γ∧γ
(b) Calculate the three interior products: ιX α, ιX β, ιX γ.
(c) Calculate the pullback along f of the three forms: f ∗ (α), f ∗ (β), f ∗ (γ).
(d) Construct, or prove that it is impossible to do so, two one-forms η, θ ∈ Ω1 (R4 )
such that β = η ∧ θ.
LECTURE 13
where by (1) the first term is determined since dai1 ...ik is the usual differential of
the smooth map ai1 ...ik . For the second term, by applying (3), we obtain:
d dxi1 ∧ . . . ∧ dxik = d dxi1 ∧ dxi2 ∧ . . . ∧ dxik − dxi1 ∧ d dxi2 ∧ . . . ∧ dxik .
Since d extends the usual derivation (1), and since d2 = 0 (2), the first term is zero.
By repeatedly applying this argument, the entire expression is zero.
We conclude that there is a unique linear operator d : Ω(M ) → Ω(M ) satisfying
the conditions (1)-(3), and that on a form α ∈ Ωk (M ) as in (*) it acts by:
X
(**) dα = dai1 ...ik ∧ dxi1 ∧ . . . ∧ dxik =
i1 <...<ik
X ∂ai
1 ...ik
X
= dxi ∧ dxi1 ∧ . . . ∧ dxik .
i1 <...<ik i
∂xi
Next, we check that this linear operator satisfies indeed the properties from
Theorem 13.1.1. Property (1) clearly holds.
We check (2)-(4) first on smooth functions. For a ∈ C ∞ (M ), we have that:
n
! n n X n
∂2a
2
X ∂a i X ∂a i
X
d a=d i
dx = d i
∧ dx = i ∂xj
dxj ∧ dxi =
i=1
∂x i=1
∂x i=1 j=1
∂x
n
X ∂2a X ∂2a
dxi ∧ dxi + dxj ∧ dxi + dxi ∧ dxj = 0,
= i 2 i j
i=1
(∂x ) i<j
∂x ∂x
where we have used that the partial derivatives commute, and the following alge-
braic properties of the exterior algebra:
dxi ∧ dxi = 0 and dxi ∧ dxj = −dxj ∧ dxi .
The derivation rule (3), for a, b ∈ C ∞ (M ), is immediate:
d(ab)(v) = v(ab) = v(a)b + av(b) = (da · b + a · db)(v), ∀ v ∈ T M.
Similarly, for f : N → M and a ∈ C ∞ (M ), (4) holds by the chain rule:
f ∗ (da)(v) = da(df (v)) = d(a ◦ f )(v) = d(f ∗ a)(v), ∀ v ∈ T N.
Next, we show that (3) holds in all degrees. Since the wedge product is bilinear,
and d is linear, it suffices to show (3) for elements of the form α = adxI ∈ Ωk (M )
LECTURE 13. 115
Using this, and again (4) for functions, we obtain that (4) holds in general:
df ∗ (α) = d f ∗ (a)f ∗ (dxI ) = d(f ∗ a) ∧ f ∗ (dxI ) + f ∗ (a)df ∗ (dxI ) =
By Lemma 4.1.2, there are function x ai1 ...ik ∈ C ∞ (M ) and an open set p ∈ V ⊂ U
ei , e
i i
e |V = x |V and e
such that x ai1 ...ik |V = ai1 ...ik |V . Consider the form:
X
αe := ai1 ...ik de
e xi1 ∧ . . . ∧ de
x ik .
i1 <...<ik
By the above, de
α = De e|V , Lemma 13.1.6 gives:
α. Since α|V = α
(Dα)|V = (De
α)|V = (de
α)|V = (dα)|V .
Thus, Dα and dα are equal in a neighborhood of p. We conclude that D = d.
Proof. By item (4) of Theorem 13.1.1, we have that f ∗ : Ω(N ) → Ω(M ) sends
closed forms to closed forms, and exact forms to exact forms; thus it induces indeed
a map f ∗ : HdR (N ) → HdR (M ). The fact that f ∗ is an algebra homomorphism on
differential forms, implies that it is an algebra homomorphism in cohomology, and
similarly, the equality f ∗ ◦ g ∗ = (g ◦ f )∗ holds already at the level of forms.
This operator is called a homotopy operator, and relation (*) is all what is needed
k
to conclude the proof. Namely, if [ω] ∈ HdR (M ), then dω = 0, and so
f1∗ ([ω]) = [f1∗ (ω)] = [f0∗ (ω) + d ◦ h(ω)] = f0∗ ([ω]).
Thus f1 and f0 induce the same map in cohomology.
Remark 13.2.13. The operator Ht : Ωk (N ) → Ωk−1 (M ) can be given explicitly:
Ht (ω) = ft∗ (ι d ft ω).
dt
follows by applying the derivation rule, that both commute with d, and that both
act the same on functions.
Formula (5) follows by using (2)-(4):
L[X,Y ] = d ◦ ι[X,Y ] + ι[X,Y ] ◦ d =
= d ◦ (LX ◦ ιY − ιY ◦ LX ) + (LX ◦ ιY − ιY ◦ LX ) ◦ d =
= LX ◦ d ◦ ιY − d ◦ ιY ◦ LX + LX ◦ ιY ◦ d − ιY ◦ d ◦ LX =
= LX ◦ (d ◦ ιY + ιY ◦ d) − (d ◦ ιY + ιY ◦ d) ◦ LX =
= LX ◦ LY − LY ◦ LX .
A differential k-form ω ∈ Ωk (M ) can be viewed also as a C ∞ (M )-multilinear
k-form on the C ∞ (M )-module X(M ) with values in C ∞ (M ):
ω : X(M ) × . . . × X(M ) −→ C ∞ (M )
| {z }
k × X(M )
ω(X1 , . . . , Xk )(p) := ωp ((X1 )p , . . . , (Xk )p ),
for X1 , . . . , Xk ∈ X(M ). This allows for a coordinate-free definition of the exterior
derivative:
Theorem 13.3.3. The exterior derivative satisfies:
k
X
dω(X0 , . . . , Xk ) = (−1)i LXi (ω(X0 , . . . , X
bi , . . . , Xk ))
i=0
X
+ (−1)i+j ω([Xi , Xj ], X0 , . . . , X
bi , . . . , X
bj , . . . , Xk ),
0≤i<j≤k
k
for all ω ∈ Ω (M ) and all X0 , . . . , Xk ∈ X(M ), where the notation X
bi means that
the vector field Xi is being skipped.
The proof is left as an exercise. In low degrees the formula gives:
da(X0 ) =LX0 (a),
dθ(X0 , X1 ) =LX0 (θ(X1 )) − LX1 (θ(X0 )) − θ([X0 , X1 ]),
dω(X0 , X1 , X2 ) =LX0 (ω(X1 , X2 )) − LX1 (ω(X0 , X2 )) + LX2 ω(X0 , X1 )
− ω([X0 , X1 ], X2 ) + ω([X0 , X2 ], X1 ) − ω([X1 , X2 ], X0 ),
for a ∈ C ∞ (M ), θ ∈ Ω1 (M ) and ω ∈ Ω2 (M ).
This vector space endowed with the bilinear extension of the commutator is the
graded algebra of derivations of A:
(Der(A), +, [·, ·]c ).
The commutator satisfies the following relations:
[D1 , D2 ]c = −(−1)pq [D2 , D1 ]c ,
[D1 , [D2 , D3 ]c ]c = [[D1 , D2 ]c , D3 ]c + (−1)pq [D2 , [D1 , D3 ]c ]c ,
for all D1 ∈ Derp (A), D2 ∈ Derq (A) and D3 ∈ Der(A). These relations are called
graded commutativity and the graded Jacobi identity, respectively. Note
that the second relation is equivalent to the operator [D1 , ·]c being a derivation of
degree p of Der(A), i.e. [D1 , ·]c ∈ Derp (Der(A)). A graded algebra which satisfies
these two axioms is called a graded Lie algebra.
In our case, A = Ω(M ), with multiplication the usual wedge product. In the
language of graded algebras, the derivation rule for the interior product and the
exterior derivative become:
ιX ∈ Der−1 (Ω(M )) and d ∈ Der1 (Ω(M )).
Note that we have the relations d2 = 0 and ιX ◦ ιY = −ιY ◦ ιX can be expressed
using the graded commutator:
[d, d]c = 2d2 = 0 and [ιX , ιY ]c = ιX ◦ ιY + ιY ◦ ιX = 0.
Moreover, Theorem 13.3.2 can be restated as follows:
(1) LX ∈ Der0 (Ω(M ))
(2) [LX , d]c = 0
(3) [LX , ιY ]c = ι[X,Y ]
(4) LX = [ιX , d]c
(5) [LX , LY ]c = L[X,Y ]
13.5. Exercises
The exterior derivative of a function a ∈ C ∞ (M ) can be calculated as follows:
d
da(v) = a(γ(t))t=0 ,
dt
dγ
where t 7→ γ(t) is a curve on M , with dt (0) = v. The following exercise extends
this result to higher degrees:
124 IOAN MĂRCUT
, , MANIFOLDS
Hint: Calculate dγ ∗ (ω), and use item (4) from Theorem 13.1.1.
Exercise 13.2. Let U ⊂ R3 be an open set. Recall the standard operators from
Calculus: the gradient, the curl and the divergence:
grad = ∇ : C ∞ (U ) −→ X(U ),
curl = ∇× : X(U ) −→ X(U ),
div = ∇· : X(U ) −→ C ∞ (U ),
C ∞ (U )
∇ / X(U ) ∇×
/ X(U ) ∇· / C ∞ (U ) .
id A B C
C ∞ (U )
d / Ω1 (U ) d / Ω2 (U ) d / Ω3 (U )
We will show that smooth densities, i.e. smooth section of the density bundle, can
be canonically integrated. The discussion below follows [10].
where we used the change of coordinate formula for multiple integrals. More gen-
erally, the same argument applies to the following situation: if Q is any other
n-dimensional parallelotope spanned by w1 , . . . , wn , and B : Rn → Rn is linear
isomorphism satisfying B(vi ) = wi , for 1 ≤ i ≤ n (i.e. B is the transition matrix
between the two bases) then we have that:
Vol(Q) = |det(B)|Vol(P ).
The following is the abstract notion of a volume of parallelotopes:
127
128 IOAN MĂRCUT
, , MANIFOLDS
14.3. Densities
Definition 14.3.1. The space of sections of the density bundle will be denoted by
Sections of the density bundle are called smooth densities (or 1-densities, or just
densities) on M . A density µ ∈ D(M ) is said to be positive, if µp (v1 , . . . , vm ) > 0
for every p ∈ M and every basis v1 , . . . , vm of Tp M .
µp = f (p)|dx1 ∧ . . . ∧ dxm |p ,
Proof. Consider an atlas on M , {(Ui , ϕi )}i∈I , and let {ρi }i∈I be a subordinate
partition of unity. Let µi = |dx1ϕi ∧ . . . ∧ dxm ϕi | ∈ D(Ui )Pbe the local density
corresponding to the chart ϕi . It is easy to check that µ := i∈I ρi µi is a positive
density on M .
130 IOAN MĂRCUT
, , MANIFOLDS
(2) It is local in the sense that: if µ ∈ Dc (M ) has support inside the open U , then
Z Z
µ= µ.
U M
∼
(3) It is invariant under diffeomorphisms: if ϕ : M −→ N is diffeomorphism, and
µ ∈ Dc (N ), then Z Z
µ= ϕ∗ (µ).
N M
1This property looks quite tautological. However, note that the right hand side of the equation is
the integral of the smooth compactly supported function f on U (the Riemann/Lebesgue integral),
and dx1 . . . dxm plays a formal notational role (the “volume element” which one often forgets to
write), whereas µ = f (x1 , . . . , xm )|dx1 ∧ . . . ∧ dxm | is an “actual volume element”, as defined in
Definition 14.1.1.
LECTURE 14. 131
Note that this definition is forced by conditions (1)-(3). We prove that this def-
inition does not depend on the chart. Consider a second chart (V, ψ) such that
supp(µ) ⊂ V . Then, by applying Step 2 to the change of coordinates diffeomor-
phism θ = ϕ ◦ ψ −1 : ψ(U ∩ V ) → ϕ(U ∩ V ), we obtain
Z Z Z
(ϕ−1 )∗ (µ) = θ∗ (ϕ−1 )∗ (µ) = (ψ −1 )∗ (µ).
ϕ(U ∩V ) ψ(U ∩V ) ψ(U ∩V )
Step 4. Consider a cover of M by coordinate charts {(Ui , ϕi )}i∈I , and let {ρi }i∈I
be a partition of unity subordinate to it. For µ ∈ Dc (M ), define
Z XZ
µ := ρi µ.
M i∈I M
R
Since ρi µ is supported in the domain Ui of the chart ϕi , it follows that M ρi µ is
well-defined; since the support of µ is compact, and the family {ρi }i∈I is locally
finite, only for a finite Rnumber of i ∈ I we have that ρi µ 6= 0, therefore the sum
above is finite. Thus, M µ is well-defined. Moreover, the definition is forced by
linearity of the integral and by uniqueness of the integral on the domains of charts
(see Step 3).
Step 5. Next, we prove that the definition is independent on the atlas and on
the partition of unity. Consider a second atlas {(Vj , ψj )}j∈JP , with subordinate
partition of unity {σj }j∈J . For each i ∈ I, we have ρi µ = j∈J ρi σj µ is a finite
R
sum of densities supported in Ui . Since M : Dc (Ui ) → R is linear,
Z XZ
ρi µ = ρi σj µ.
M j∈J M
Since all sums involved are finite, we can change the summation order on the
right hand side. Therefore, by applying the same argument with the two atlases
interchanged, we obtain:
XZ XZ
ρi µ = σj µ.
i∈I M j∈J M
This shows that the integral is independent on the choice of the atlas.
We have proven existence of the integral, and we have proven that the condi-
tions (1)-(3) imply uniqueness of the integral. We leave it to the reader to check
that conditions (1)-(3) are indeed satisfied for the integral we have constructed.
14.5. Orientations
Definition 14.5.1. Let V be an n-dimensional vector space V , and let B(V ) denote
the space of all bases of V . An orientation on V is a map
o : B(V ) −→ {−1, 1},
such that, for all linear isomorphisms A : V −∼
→ A, it satisfies:
o(Av1 , . . . , Avn ) = sign(det(A))o(v1 , . . . , vn ).
We denote by Or(V ) the set of orientations on V . Given an orientation o on V a
basis v1 , . . . , vn of V is said to be a positive basis for o (resp. negative basis), if
o(v1 , . . . , vn ) = 1 (resp. o(v1 , . . . , vn ) = −1).
Clearly, every vector space has precisely two orientations. This is true even for
a 0-dimensional vector space, which has a unique basis, namely the empty set ∅.
The two orientations are o(∅) = 1 and o(∅) = −1; we simply write o = 1 or o = −1.
The standard orientation on Rn is the orientation for which the standard
basis e1 , . . . , en is positive. In general, a basis v1 , . . . , vn of Rn is positive with
respect to the standard orientation iff the matrix A = (vij )1≤i,j≤n has positive de-
terminant, where we have denoted vi = (vi1 , . . . , vin ).
14.6. Exercises
Exercise 14.1. (a) Let (V, h·, ·i) be an n-dimensional vector space together with
an inner-product. Prove that the following expression defines a positive volume
element on V : q
µ(v1 , . . . , vn ) := det (hvi , vj ii,j ).
(b) Let M ⊂ Rn be an embedded submanifold. For each p ∈ M , we regard Tp M as
a linear subspace of Tp Rn = Rn . The restriction of the standard dot-product on
Rn gives an inner-product on Tp M . Show that the following defines a smooth
density on M :
q
µMp (v 1 , . . . , v m ) := det (hvi , vj ii,j ),
where v1 , . . . , vm is a basis of Tp M .
n
(c) With the notation from (b), prove that µR = |dx1 ∧ . . . ∧ dxn |.
(d) For n = 3, show that the construction from (b) recovers the usual notions of
length of a curve and area of a surface, more precisely:
• Let γ : (a, b) ,→ R3 be a smooth embedding, and let C be the image of
γ. Prove that µC = ds, where, at p = γ(t), dsp denotes the “arc-length
134 IOAN MĂRCUT
, , MANIFOLDS
element”, dγ
dt (t) dt. Conclude that
Z Z b
C
dγ
µ = dt (t) dt,
C a
assuming that both sides are bounded (note that the right hand side is
the length of C).
• Let U ⊂ R2 be an open set, and let σ : U ,→ R3 be a smooth embedding
parameterizing the curve S := σ(U ). Prove that µSp = dSp , where dSp is
the “area form”:
∂σ ∂σ
dSp := (u, v) ×
(u, v) dudv, at p = σ(u, v).
∂u ∂v
Conclude that the following holds:
Z Z
S
∂σ ∂σ
µ = ∂u (u, v) × ∂v (u, v) dudv,
S U
assuming that both sides are bounded (note that the right hand side is
the area of S).
Exercise 14.2. Let G be a Lie group with Lie algebra g := Te G. A smooth density
µ ∈ D(G) is called left invariant, if λ∗g (µ) = µ for all g ∈ G (recall that λg denotes
left translation by g). Let DL (G) denote the space of left invariant densities.
(a) Prove that evaluation at the identity
DL (G) −→ Ve(g), µ 7→ µ|p
is a linear isomorphism.
(b) If G is compact, prove
R that there exists a unique left invariant density µG ∈
DL (G) such that G µG = 1 (The density µG is called the Haar measure of
G).
(c) For g ∈ G and µ ∈ DL (G), prove that ρ∗g (µ) ∈ DL (G) (recall that ρg denotes
right translation by g). Moreover, show that ρ∗g (µ) = u(g)µ, where u(g) ∈
R\{0} does not depend on µ 6= 0. Prove that the map u : G → R\{0},
g 7→ u(g) is a Lie group homomorphism.
(d) If G is compact, prove that u(g) = 1, for all g ∈ G.
(e) Consider the Lie group G = Aff(1) consisting of affine transformation of the real
line R, i.e. the elements of Aff(1) are maps f : R → R of the form f (x) = ax+b,
with a 6= 0, and the group structure is the composition of maps. Calculate the
map u : Aff(1) → R\{0}.
Exercise 14.3. Prove that a manifold M is orientable if and only if it admits
an atlas {(Ua , ϕa )}a∈I for which all transition maps ϕa ◦ (ϕb )−1 : ϕb (Ua ∩ Ub ) →
ϕa (Ua ∩ Ub ), for a, b ∈ I, have positive Jacobian determinant:
!
∂xia
det > 0,
∂xjb
where ϕa = (x1a , . . . , xm
a ).
Clearly the composition of smooth maps between open subsets of upper half-
spaces is again smooth (its extension is the composition of the extensions), and the
usual chain rule holds as well (it follows by the chain rule for the extensions).
The notion of a diffeomorphism generalizes accordingly: if U, V ⊂ Hn are
open subsets, then f : U → V is a diffeomorphism if and only if f is smooth,
bijective, and f −1 : V → U is also smooth.
The boundary and the interior are preserved under diffeomorphisms:
Lemma 15.1.1. Let U, V ⊂ Hn be open sets, and let f : U → V be a diffeomor-
phism. Then
f (∂U ) = ∂V f (int(U )) = int(V ).
Proof. Note that f |int(U ) : int(U ) → Rn is an injective local diffeomorphism;
hence f (int(U )) is an open set (as a subset of Rn ). This implies that f (int(U )) ⊂
int(V ). The same argument for f −1 implies that f −1 (int(V )) ⊂ int(U ). Since f is a
bijection, we conclude that f (int(U )) = int(V ). But then f restricts to a bijection
between the complements of these sets; thus, f (∂U ) = ∂V .
For any p ∈ ∂M , we have that ψ(0, p) = p, hence d(0,p) ψ(0, v) = v for any v ∈
Tp ∂M . On the other hand, since the curve t 7→ ψ(t, p) is a flow line of X,
∂
d(0,p) ψ
, 0 = Xp .
∂t
Since Tp M = Tp ∂M ⊕ RXp , we conclude that d(0,p) ψ is a linear isomorphism. The
Inverse Function Theorem 6.1.1, implies that ψ is a diffeomorphism in a neigh-
borhood of p. On the other hand, since ∂M is compact and ψ(0, p) = p for all
p ∈ ∂M , Exercise 6.8 implies that ψ is a diffeomorphism between a neighborhood
V ⊂ U of {0} × ∂M and a neighborhood O ⊂ M of ∂M . On the other hand,
since ∂M is compact, there exists > 0 such that [0, ) × ∂M ⊂ V . Finally, define
ϕ : [0, 1) × ∂M → M , ϕ(t, p) = ψ(t, p).
Remark 15.2.3. If the orientation changes, the sign of the integral changes:
Z Z
ω=− ω.
(M,−o) (M,o)
where the coefficients fi ∈ Cc∞ (Hm ) are compactly supported function on Hm , and
di denotes that dxi is missing in the product. The differential of ω is given by:
dx
m
!
i−1 ∂fi
X
dω = (−1) i
(x) dx1 ∧ . . . ∧ dxm .
i=1
∂x
Note that, under the standard orientation on Hm , we have that dx1 ∧ . . . ∧ dxm
corresponds to the density |dx1 ∧ . . . ∧ dxm |, and therefore:
Z m Z
X ∂fi
dω = (−1)i−1 i
(x)dx1 . . . dxm .
Hm
i=1 Hm ∂x
Choosing a large enough number N > 0 such that
supp(ω) ⊂ [−N, N ]m−1 × [0, N ].
Consider a term in the sum above with 1 ≤ i < m. By interchanging the order of
integration, we first calculate the integral with respect to dxi :
Z N
∂fi
i
(x)dxi = fi (x1 , . . . , xi−1 , N, xi+1 , . . . , xm )
−N ∂x
− fi (x1 , . . . , xi−1 , −N, xi+1 , . . . , xm ) = 0,
where we have used the Fundamental Theorem of Calculus and that supp(fi ) ⊂
[−N, N ]m−1 × [0, N ]. The term with i = m gives:
Z N
∂fm
m
(x)dxm = fm (x1 , . . . , xm−1 , N ) − fm (x1 , . . . , xm−1 , 0) =
0 ∂x
= −fm (x1 , . . . , xm−1 , 0).
We conclude that
Z Z
m
(*) dω = (−1) fm (x1 , . . . , xm−1 , 0)dx1 . . . dxm−1 .
Hm Rm−1
m
On the other hand, if i : ∂H → Hm denotes the inclusion, we have that:
i∗ (ω) = fm (x1 , . . . , xm−1 , 0)dx1 ∧ . . . ∧ dxm−1 .
144 IOAN MĂRCUT
, , MANIFOLDS
Note that the vector − ∂x∂m is pointing outwards, therefore, the boundary orienta-
tion on ∂Hm is such that
∂ ∂ ∂ ∂ ∂
(∂ost ) , . . . , = o st − , , . . . , = (−1)m .
∂x1 ∂xm−1 ∂xm ∂x1 ∂xm−1
We obtain that
(∂ost )dx1 ∧ . . . ∧ dxm−1 = (−1)m |dx1 ∧ . . . ∧ dxm−1 |,
and therefore
Z Z Z
∗ m
ω= (∂ost ) i (ω) = (−1) fm (x1 , . . . , xm−1 , 0)dx1 . . . dxm−1 .
∂Hm ∂Hm Rm−1
m
This and (∗) prove Stokes’ Theorem for H with the standard orientation.
Consider an oriented manifold (M, o) with boundary. Let ω ∈ Ωm−1 c (M ).
Assume first that the support of ω lies inside the domain of a coordinate chart
ϕ : U → Hm , such that U and ∂U := U ∩ ∂M are connected. By Proposition 15.2.4
and Stokes’ Theorem for Hm , we obtain that
Z Z Z Z Z
−1 ∗ −1 ∗
dω = dω = (ϕ ) (dω) = d(ϕ ) (ω) = d(ϕ−1 )∗ (ω) =
M U ϕ(U ) ϕ(U ) Hm
Z Z
−1 ∗
= (ϕ ) (ω) = (ϕ−1 )∗ (ω),
∂Hm ∂ϕ(U )
15.4. Exercises
Exercise 15.1. Let M be a manifold without boundary, and let f : M → R be a
smooth function. If 0 ∈ R is a regular value of f , prove that N := f −1 ([0, ∞)) can
be given the structure of a manifold with boundary ∂N = f −1 (0) such that the
inclusion N → M is smooth. Hint: use the local submersion Theorem 6.2.3.
Exercise 15.2. Show that Stokes’ Theorem implies the following classical results
from Calculus:
(a) Green’s Theorem. Let Ω ⊂ R2 be an open set bounded by the simple closed
smooth curve C. For every smooth vector field F = (f1 , f2 ) on R2 , we have
that: I ZZ
∂f2 ∂f1
f1 dx + f2 dy = − dxdy.
C Ω ∂x ∂y
(b) Divergence Theorem. Let Ω ⊂ R3 be a bounded open set with smooth boundary.
Then for every smooth vector field F on R3 , we have that:
ZZZ ZZ
divF dv = F · n dA.
D ∂D
3
(c) Classical Stokes’ Theorem. Let S ⊂ R be a smooth bounded surface with
smooth boundary. Then for every smooth vector field F = (f1 , f2 , f3 ) on R3 ,
we have that:
I ZZ
f1 dx + f2 dy + f3 dz = curl(F) · n dA.
∂S S
Use Exercise 13.2 and a Calculus book (or Wikipedia) to make sense of all terms.
Exercise 15.3. Let U ⊂ C be an open set, and let f : U → C be a holomorphic
function. Consider a closed disk D ⊂ U with center a and boundary C. Prove
Cauchy’s integral formula:
I
1 f (z)
f (a) = dz.
2πi C z − a
Hint: Use Exercise 13.4 and Stokes’ Theorem.
BIBLIOGRAPHY
17. https://en.wikipedia.org/wiki/Long_line_(topology)
18. http://pages.uoregon.edu/koch/math431/LongLine.pdf
19. https://commons.wikimedia.org/wiki/File:Orientation_cover_of_
Mobius_strip.webm
20. https://en.wikipedia.org/wiki/Prfer_manifold
21. https://en.wikipedia.org/wiki/Space-filling_curve
22. http://earth.nullschool.net/
EXERCISES AND ASSIGNMENTS
Homework assignments
• Homework for 15/9: 1.2, 1.7.
• Homework for 22/9: 2.3.
• Homework for 29/9: 3.9.
• Homework for 6/10: 4.3.
• Homework for 13/10: 5.1.
• Homework for 20/10: 6.3, 6.6.
• Homework for 27/10: 6.5, 6.11.
• Homework for 17/11: 8.1, 8.2, 8.3.
• Homework for 24/11: 9.3, 9.7, 9.8.
• Homework for 1/12: 10.2,10.3,10.4.
• Homework for 8/12: 11.3.
• Homework for 15/12:12.3,12.9.
• Homework for 22/12: 13.4, 13.5, 13.6.
• Homework for 12/1: 14.1, 14.6.
• Homework for 19/1: 15.2.
149