Riem Geom
Riem Geom
Riem Geom
Notes
(*) means that the proof/proposition of the theorem is not required while (**) means
that the theorem/the argument is not required for the exam
Contents
1. Topological manifolds 1
2. Smooth manifolds 4
3. The partition of unity (*) 7
4. Tangent and cotangent bundle 13
5. Vector bundles 18
6. The cotangent bundle 22
7. Tensors 24
7.1. Differential forms and integrations 29
8. Integral curves and symmetry generators 32
8.1. Sumbmanifolds 36
9. A mini course on linear connections 39
9.1. Riemannian geometry 44
9.2. Riemannian distance 48
9.3. Metric connection 48
9.4. Geodesics in Riemannian geometry (**) 51
9.5. The Riemannian curvature 52
10. Principal bundle and gauge theories (**) 54
1. Topological manifolds
Informally we can think that manifolds are generalization of surfaces and curves to
higher dimensions. The dimension of the manifolds would be the number independent
parameters needed to specify a “point”; thus an n dimensional manifold is in some sense
an object modelled locally on Rn . The easiest example we can visualize is the sphere S 2
given by the equation x2 + y 2 + z 2 = 1; near the north pole (0, 0, 1) we can solve for z
as a function of x and y thus we would need two parameter to specify a point close to
the north pole and that’s why we will say that the sphere is a 2 dimensional manifold
in the sense the “locally” looks like R2 . But what do we mean with looks like??? the
main idea is that U ∈ Rk and V ∈ Rn are said to be homeomorphic if we can find a one
to one correspondence ϕ : U → V s.t. ϕ and its inverse are continuous map.
Let’s try to be a bit more “mathy” and abstract now. We want to come up with a
notion of “space” in which the notion continuous function makes sense. Let’s discuss the
case of Rn for simplicity the generalization to any metric space is obvious. A set U in Rn
is said to be open if for every p in U there is an open ball Br (p) such that Br (p) ⊂ U . A
neighborhood of p is an open set containing p. It is clear that the union of an arbitrary
collection of open sets is open, but NOTE the same is not true of the intersection of
infinitely many open sets. A map f from an open subset of Rn to Rm is continuous if
and only if the inverse image f −1 (V ) of any open set V in Rm is open in Rn . This shows
that continuity can be defined in terms of open sets only for Rn as for every metric
1
2
spaces 1. More in details first one defines open ball Br (p) = {x ∈ Rn s.t. d(x, p) < r}
where d(x, y) is the usual euclidean distance. We define axiomatically thus open sets
looking at what happens in Rn
Definition 1.1. Let X be a set; a topology on X is a collection τ of subsets of X, called
open sets,such that
• X, ∅ ∈ τ
• U1 , .., Un ∈ τ → U1 ∩ .. ∩ Un ∈ τ (intersection of finite
S number of opens)
• {Ui } finite or infinite collection of elements of τ → Ui ∈ τ
We call the pair (X, τ ) or simply X , a topological space
The idea of open sets it is somehow needed to recover the notion of nearness we would
have in a metric space. In this sense a neighborhood of a point p ∈ X is just an open set
U containing p. A natural example of topology of a set S is the collection of all subsets
of S. This is sometimes called the discrete topology.
Let’s resume an example: Consider the set S = {1, 2, 3}
• τ = {{1}, {1, 2}, {1, 2, 3}, {∅}} is a topology
• τ = {{S}, {∅}} is the trivial topology
• τ = {{1}, {2}, {3}{1, 2}, {1, 3}, {2, 3}{1, 2, 3}, {∅}} is the discrete topology
Once we have a top. space (S, τ ) a subset naturally have a top. space structure (A, τA )
given by
τA := {U ∩ A with U ∈ τ }
This is sometimes called subspace topology or relative topology.
Take now a topological space X and a set S and π : X → S a surjective map. We can
naturally define a topology on S by saying that U ∈ S is open if π −1 (U ) is open. This
is called quotient topology In general it’s hard to describe all the open sets in τ and we
introduce thus the notion of a base.
Definition 1.2. A subcollection β of a topology τ of a set S is a basis for τ if for every
U ∈ τ and p ∈ U we can find V ∈ β s.t. p ∈ V ⊂ U
An useful criterion for deciding if a collection of subsets is a basis for some topology
is given by the following:
Proposition 1.3. A collection β = {ui } of subset of S is a basis for some topology of
S if and only if
• S is the union of all ui
• given ui and uj and p ∈ ui ∩ uj there is uk ∈ β s.t. p ∈ uk ⊂ ui ∩ uj
Another point of view or if you prefer a consequence of the previous definition is to
say that β is a basis if every open set of S can be recovered by a union of sets in β.
Definition 1.4. A top space is named second countable if it admits a countable basis.
Fact: Rn is second countable and and subset of a second countable top space is second
countable One of the most important property of a second countable top space is that
any open cover admits a countable subcover.
Sometimes one may want that open sets separate points. More precisely
1For metric space we say that given (X, d ) and (Y, d ) be metric spaces a function f : X → Y
X Y
is continuous at p ∈ X if for every > 0 there exists δ > 0 such that dX (x, p) < δ implies that
dY (f (x), f (a)) < .
3
Definition 1.5. A topological space S is Hausdorff if given any two distinct points x
and y in S, there exist disjoint open sets U, V such that x ∈ U and y ∈ V .
We now go back to the concept of continuity of a function. Let f : X1 → X2 be a
function of topological spaces. Mimicking the definition from advanced calculus, we say
that f is continuous at a point p in X1 if for every neighborhood U2 of f (p) in X2 (that
is an open set containing f (p) , there is a neighborhood U1 of p such that f (U1 ) ⊂ U2 .
In the case of quotient topology one can prove that a function from the quotient space
is continuous IFF its composition with the quotient map is continuous.
The notion of equal in the “category” of top spaces is encoded in the definition of
homeomorphism
Definition 1.6. Two top space X and Y are said to be homeomorphic or topologically
equivalent if exists an homeomorphism between them that is a cont and bijective map
with cont inverse.
Definition 1.7. A topological space is disconnected if it is the union of 2 disjoint non
empty open sets. If it is not disconnected we will call it connected.
Definition 1.8. An open cover of a top space S is a collection of open subsets whose
union is S. A subcover is a subcollection still covering S. We will say that S is compact
if every open cover admits a finite subcover.
A subset A of a top space S is named closed if its complement is open. Note that
closed does not mean not open for example given the top space S the open set S is also
closed being its complement the empty set. We can define the closure of a set A in S by
Ā = ∩{B ⊂ S : A ⊂ B and B is closed}
It is time to translate to the math language the notion of “looks like Rn ”. A topologi-
cal space M is said to be locally Euclidean of dimension n if every point p ∈ M has a
neighborhood U that is homeomorphic to an open subset of Rn , that is we have an home-
omorphism ϕ : U → U e ⊂ Rn . Let’s call the pair (U, ϕ) the local coordinate chart and
we sometimes refer to U as coordinate open set. If the image of ϕ is an open ball of Rn
we will call U Euclidean open ball in M if ϕ(p) = 0 we say that the chart is centered in p.
Definition 1.9. A subset of a top space is called precompact if its closure is compact.
Definition 1.10. A topological manifold is a top space that is Hausdorff second count-
able and locally Euclidean.
Note that open subset of a top manifold is a top manifold. Let’s discuss now a couple
of examples:
• S n is for free Hausdorff and second countable bc it lives inside Rn . Thus we need
just to check that it’s locally Euclidean. To this aim consider:
Ui± = {x = (x1 , .., xn+1 ) ∈ S n t.c. ± xi > 0}
equipped with the following homomorphism
ϕ±
i : Ui± → U ei ⊂ Rn
(x1 , .., xn+1 ) 7→ (x1 , .., xi−1 , xi+1 , .., xn+1 )
4
2. Smooth manifolds
We now want to add more structure to our manifolds in order to define functions
derivative of functions or more in general the concept of smoothness, in a consistent
way. In particular if we consider a topological manifold M we may want to define a
function f : M → R smooth at p ∈ M if fˆ := f ◦ ϕ−1 : Rn → R is smooth at p̂ := ϕ(p)
for some coordinate chart (U, ϕ). The problem of this definition is that it depends too
much on the choice of the coordinate chart. Consider for example two charts (U, ϕU ),
(V, ϕV ) with U ∩ V 6= ∅ and a point p ∈ U ∩ V then we may obtain that in the coordinate
chart U
fˆ|U := f ◦ ϕ−1
U
is smooth at p while in the intersection U ∩ V in the V chart
(fˆ|V )|U ∩V := f ◦ ϕ−1 |U ∩V = f ◦ ϕ−1 ◦ (ϕU ◦ ϕ−1 )|U ∩V
V U V
If ϕU ◦ ϕ−1
V is not good enough we can find a contradictions, that is f is smooth in a
given coordinate chart but not in the other. Let’s solve this problem. Let’s consider
to chart (Ui , ϕi ) and (Uj , ϕj ) and let’s define Uij := Ui ∩ Uj and we construct the so
called transition function ϕij := ϕj ◦ ϕ−1 n n
i |ϕi (Uij ) : R ⊃ ϕi (Uij ) → ϕj (Uij ) ⊂ R (we will
5
often omit the domain of the transition function to simplify the notation). In the case
of topological manifolds the transition functions are obviously homomorphisms. But
we want something more. In teh previous formula we emphasized the domani of he
functions and the chart used, we will omit in the future those details in order not tobe
pedantic with th notation, it willbe clear for the context what we are doing.
Let’s remember to ourselves the following definition:
Definition 2.1. A smooth map from an open set V of Rn to an open set W of Rm
bijective and with a smooth inverse is called a diffeomorphism
That’s what we need.
Definition 2.2. Two coordinate charts (Ui , ϕi ) and (Uj , ϕj ) are named smoothly com-
patible if Uij = ∅ or ϕij is a diffeo.
Now we put all charts together in a clever way.
Definition 2.3. A smooth atlas, or simply an atlas, A, for a topological manifold M is
a collection of smoothly compatible charts covering M .
Note that it is enough to check that every transition function ϕij is smooth (thus for
all i, j) to prove that we have an Atlas. We thus successfully removed the ambiguity of
the definition of a smooth function on M once we fixed the Atlas; a function f : M → R
then is smooth if fˆ = f ◦ ϕ−1 is smooth for every coordinate chart in the atlas A. We
are tempted to call this a “smooth structure”. A minor problem appears now. Consider
for example on R2 the following atlases
A1 := {(R2 , id)}
and
A2 := {(B1 (x), id)}
they are obviously different but they determine the same set of smooth functions. The
way out is easy
Definition 2.4. A smooth structure is an equivalence class of atlases where we will
declare the atlas A1 equivalent to A2 if A1 ∪ A2 ia another smooth atlas
Alternatively one can say the following
Definition 2.5. A smooth structure is a maximal Atlas Amax , one thus that is not
contained in a larger atlas.
Lemma 2.6. (1) Every smooth atlas A for M is contained in a unique max atlas
(2) Two atlases determine the same max atlas IFF their union is an atlas
Proof. Define now Amax to be the set of all possible chart compatible with those in A;
note that this means that is maximal by construction. Is it an Atlas? Amax obviously
cover M thus we need to prove compatibility of the charts. Lets consider (V1 , ψ1 ) and
(V2 , ψ2 ) charts in Amax ans suppose V1 ∩ V2 6= ∅; are the compatible? that is ψ1 ◦ ψ2−1
is a diffeo? Take a point p ∈ V1 ∩ V2 , then we can find a chart (U, ϕ) in A with p ∈ U .
Observe now that
ψ1 ◦ ψ2−1 = ψ1 ◦ ( ϕ−1 ◦ ϕ ) ◦ ψ2 = (ψ1 ◦ ϕ−1 ) ◦ (ϕ ◦ ψ2 )
| {z } | {z } | {z }
we insert them here smooth smooth
6
and V, ϕ is our smooth structure. Suppose now we want to use another basis B̃
where vB̃ = (ṽ1 , .., ṽn ), we know we can always find an invertible matrix Aij so
that
ṽi = Aij vj
thus in general an invertible linear map
ϕBB̃ : Rn → Rn
vB → vB̃
that is our construction does not depend on the choice of the basis as expected
in the sense that the two chart are compatible and thus the smooth structure
induced is the same.
• Any open subset U of Rn is a topological manifold and the restriction of every
single chart of the given atlas for Rn defines a smooth structure on U . this
generalize to any open subset of a smooth manifold.
• The set of real invertible matrices GLn (R) can be defined by as det−1 (R \ 0)
where
2
det : Rn → R
2
where we have identified square matrices with Rn . Being the determinant a
continuous function and (R \ 0 an open set we can conclude that GLn (R) is open
2
inside Rn and thus a smooth manifold
• More example in the exercise sheet 1
Now we want generalize a bit the construction and we look at maps between smooth
manifolds M and N , F : M → N . We consider the smooth atlases (Ui , ϕi ) and (Va , ψa );
we say thaty F is smooth if ψa ◦F ◦ϕ−1 n
i : R ⊃ ϕi (Ui ∩F
−1 (V )) → ψ (V ) ⊂ Rm for every
a a a
(Ui , ϕi ) and (Va , ψa ). Sometimes we we just write F̂ for the coordinate representation
of F in some coordinate charts.
Definition 2.10. A map between smooth mfds F : M → N is a diffeomorphism if it is
smooth with a smooth inverse. In this case we would say that M and N are diffeomorphic
Note that the notion of being diffeomorphic depends on the smooth structure chosen
We have discuss previously that one can equip a top manifold with different atlases.In
case they are compatible we thus work with the same smooth structure, but in case not
we are temped to say that the manifolds are different from the “smooth” point of view.
This sentence is probably too strong. In the end in fact we may ad different smooth
structures to a top manifold but end up with diffeomorphic smooth mfds. Let s see it with
an example. Consider R with the standard atlas ϕ = id or with ψ(x) = x3 and denote
for simplicity Rϕ and Rψ the smooth mfds arising form those smooth structures. They
1
determine different smooth structures on R in fact the transition function ϕ◦ψ −1 (x) = x 3
1
that is not smooth at the origin. Anyhow defining F : Rϕ → Rψ F (x) = x 3 we can
conclude that they are diffeomorphic since F̂ = id. For this point of view the right
quastion one could try to ask is how many smooth structure up to diffeomorphism one
can add to a top manifold.
Lemma 3.1. Take M and N smooth mfds and {Ui } an open cover of M . Suppose then
that we have a collection of smooth functions Fi : Ui → N agreeing on the overlap then
exists a unique smooth map F : M → N such that F |Ui = Fi .
What about closed subsets?? Take as a counterexample M = N = R with the
standard smooth structure, U± = [±1, 0] and f± = ±x. They agree on the overlap
obviously but their ”union” |x| is not smooth at the origin.
Question: is there a way to have a weaker version of the gluing Lemma for smooth
manifolds, namely blend together local smooth objects without assuming that they agree
on “too big” overlaps ??
Answer: Yes it is called the partition of unity, and it is a crucial tool in differential
geometry. The idea is to construct functions that are identically vanishing in specified
parts of a manifolds. A partition of unity is used in two ways:
• to decompose a global object on a manifold into a locally finite sum of local
objects on the open sets {Ui } of an open cover
• to patch together local objects on the open sets {Ui } into a global object on the
manifold.
Thus, a partition of unity serves as a bridge between global and local analysis on a
manifold. This is useful because while there are always local coordinates on a manifold,
there may be no global coordinates. In subsequent sections. It is the single feature
that makes the behaviour of smooth manifolds so different from that of real-analytic or
complex manifolds.
One of the main tool needed will be the bumb function that we will now introduce.
Let us briefly remind ourselves that given a smooth function f : M → R on M , we can
define its support as follows:
supp(f ) = closure{p ∈ M s.t. f (p) 6= 0}
and we say f supported in some subset U if supp(f ) ⊆ U
Definition 3.2. A smooth function ρ : Rn → R is called a BUMP function for A
supported in U ⊂ Rn if it is equal to 1 in a specified closed set A and supp(f ) ⊂ U .
Sometimes we will we consider A a neighbourhood of some point p ∈ M and we sat
bumb functions at p supported in U meaning that it is 1 in some neighbourhood of p.
We now construct an example of useful and non trivial bump function.
Fact: The function
− 1t
f (t) := e t>0
0 t≤0
is smooth (BUT not analytic). Moreover the following function we can construct out of
the previous one
f (2 − t)
r(t) :=
f (2 − t) + f (t − 1)
is such that r(t) = 1 when t ≤ 1, 0 < r(t) < 1 when 1 < t < 2 and zero otherwise.
Fact: The function ρ(x) : Rn → [0, 1] defined by ρ(x) = r(|x|) is a bump function
supported in any open set containing B̄2 (0) and identically equal to 1 in B̄1 (0). The
next goal is tu use bump function efficiently. For example:
9
Proof. See for example Lee book “introduction to smooth manifolds”. The compact
manifold case is some how simpler but less instructive and can be found in W. Tu book
“introduction to manifolds”.
Let C = {Ci } an open cover for M . Consider {Vj } a countable locally finite cover for
11
M by precompact open sets that we know exists by the previous Lemma. Then for each
p ∈ M construct (Wp , ψp ) a coordinate chart centred at p so that
• ψp (Wp ) = B3 (0)
• for every p we can find a Ci so that Wp ⊂ Ci
• When p ∈ Vj then Wp ⊂ Vj (comment this is possible because the local finite-
ness of {Vj } that is each p ∈ M has a neighbourhood intersecting finitely many
of the Vj ).
Define now Up = ψp−1 (B1 (0)), and observe that for every p ∈ Vj the set {Up } is an
open cover for V̄j and by paracompactness of V̄j we have taht I can a finite subset of
(j) (j) (j) (j)
those covering it that we call {U1 , .., Un(j) } with corresponding {W1 , .., Wn(j) }. The
(j) (j)
set (Wi , ψi ) refines C and satisfies b) and c) by construction. We need to prove a)
.It is obviously countable, is it also locally finite??? This is a direct consequence of the
fact that {Vj } is a countable locally finite cover of M . We refer to the afro mentioned
books for more details.
Theorem 3.9. Given a smooth mfd M and an open cover C = {Ci } we can always find
a partition of unity subordinate to C
Proof. Consider a regular refinement {Wa }|a∈A of {Ci } where A is just a set (we need
to specify it now). Define then
Ua = ψa−1 (B1 (0)) , Oa := ψa−1 (B2 (0))
Note that by hp {Wa } still covers M . We then use the bumb function ρ constructed
previously to construct:
ρ ◦ ψa on Wa
fa :=
0 on M \ Ōa
Note that suppfa ⊂ Wa . Define then
12
fa (p)
ga (p) = P
a fa (p)
It is here that the regular refinement property plays a crucial role: the denominator
contains in fact finitely many non vanishing terms being {Wa } locally finite. Observe
then that fa = 1 on Ua by construction, and since {Ua } cover M by the regularity
of the refinement every point is P contained in some Ua and we can conclude that the
denominator never vanishes and a ga = 1 and 0 ≤ ga (p) ≤ 1 for all p in M . Now we
need to reindex to go back to the given cover {Ci }. Given Ci we can find a subset A(i)
of A so that Ua ∈ ci for all a ∈ A(i) and define
X
φi = ga , a ∈ A(i)
a
Now one can check this function satisfies the desired request. In particular note that
{supp(φi )} is locally finite because {Ua } is locally finite and supp(fa ) ⊂ Ua thus supp(φi ) ⊂
Ci
We can now use this theorem to define the bump function on a manifold:
Corollary 3.10. Let M be a smooth mfd then for any closed set A ⊂ M and open set
U containing A we can find a smooth function β : M → R such that:
β(p) = 1 , ∀p ∈ A
and supp(β) ⊂ U
Proof. Consider the open cover given by {U1 := U, U2 := M \ A}. We know we can
construct a partition of unity {φ1 , φ2 } subordinate to this cover.The function
P φ2 by
construction vanishes on A because supp(φ2 ) ⊂ M \ A thus on A we have φi = φ1 = 1
and supp(φ1 ) ⊂ U1 = U thus φ1 is our β
this is a powerful object that we call again BUMP function for A supported on U .
Let’s use it:
Lemma 3.11. Let f be a smooth function defined on a closed subset A of M . For any
open U containing A we can find f ext a smooth function on M such that it coincide with
f on A and supp(f ext ) ⊂ U
13
for some smooth functions gi vanishing in p. We now use the Leibniz rule and we get
Xp · f = Xp · f (p) +Xp · (∂i f (p)(xi − xi0 )) + Xp · gi (x)(xi − xi0 ) = (∂i f )(p)Xp (xi − xi0 )
| {z } | {z }
0 by prop. 1 0 bc of property 2
Fact: The tangent space at a point p of a finite dimensional vector space V , that is
Tp V , is isomorphic to V , namely
Tp V ∼ V , ∀p ∈ V
We were able so far to write functions representation in a chart, we will do the same for
derivations. Consider the chart (U, ϕ) for M . We know we have a canonical basis for
Tp̂ Rn given by ∂i |p̂ and that, being ϕ a diffeomorphism we know by the second property
of the push forward that Tp̂ Rn ∼ Tp M thus the set {∂i |p } defined by
∂i |p := (ϕ−1 )∗ ∂i |p̂
is a basis for Tp M . What’s the meaning of this object?? Consider a functionf : U → R
then
∂ fˆ
∂i |p f = (ϕ−1 )∗ ∂i |p̂ f = ∂i |p̂ (f ◦ ϕ−1 ) = ∂i |p̂ fˆ = (p̂)
∂xi
Let ‘s see now how to pushforward this basis. Consider two charts (U, ϕ) and (V, ψ) for
M and N and denote the coordinates in the domain by p̂ = ϕ(p) = (x1 , ..., xn ) = xi and
q̂ = ψ(q) = (z 1 , ..., z m ) = z a , take then consider f ∈ C ∞ (N ):
(F∗ (∂xi |p ))f = (∂xi |p ))(F ◦ f )
∂ F̂ a
= ∂xi
(p̂) ∂z a |F (p) f
from which we get
∂ F̂ a
(F∗ (∂xi |p )) =(p̂) ∂z a |F (p)
∂xi
We have seen that in a given coordinate chart (U, ϕ) with coordinates function of ϕ
16
given by (x1 , .., xn ), we have a natural basis for Tp M thus every tangent vector at p can
be written as
∂
Xp = ai ∂i |p = ai i
∂x p
What if we change chart (V, ψ) with coordinate function (x̃1 , .., x̃n )? (note we take
p ∈ V ∩ U ). In this case we can use the basis {∂˜i |p } with ∂˜i |p := ∂∂x̃i |p and we have
Xp = ãi ∂˜i |p
How the tilde and non tilde are related?? First of all let’s write the transition map:
ψ ◦ ϕ−1 (x) = (x̃1 (x), .., x̃n (x)) = x̃(x)
Let’s compute it by acting on a general smooth function f :
def
∂i |p f = ((ϕ−1 )∗ ∂i |ϕ(p) ) f
= ( ∂i |ϕ(p) ) f ◦ ϕ−1
| {z }
fˆ(x)
= ( ∂i |ϕ(p) ) f ◦ ψ ◦ ψ ◦ ϕ−1
−1
| {z } | {z }
fˆ(x̃) x̃(x)
∂ x̃j
chain rule = ∂xi ϕ(p) ∂˜j |ψ(p) fˆ(x̃)
def ∂ x̃j ˜
= ∂xi ϕ(p) ∂j |p f
There is a more geometrical interpretation of tangent vectors we are now able to discuss
properly
Definition 4.6. A smooth curve is a smooth map γ : J ⊂ R → M ; we will often denote
it by γ(t)
Take t0 ∈ J, define p0 = γ(t0 ) and f smooth function on M ; consider then the function
restricted to the curve, that is f (γ(t)), let’s see how it changes
d
(f ◦ γ) (t0 ) = (γ∗ ∂t |t0 )f
dt
In a given coordinate chart containing γ(t0 ) we know how to handle this object and we
get
d d d
(γ∗ ∂t |t0 )γ(t0 ) f = f ◦γ = (f ◦ ϕ−1 ◦ ϕ ◦ γ ) = fˆ(γ 1 (t), .., γ n (t))
dt t0 dt t0 | {z } | {z } dt t0
fˆ γ̂
where γ̂ = (γ 1 (t), .., γ n (t)) simply means we are taking the coordinate representation of
every point along the curve. In conclusion we obtain:
γ̇ i (t0 )∂i |γ̂(t ) fˆ = γ̇ i (t0 )∂i |γ(t ) f
0 0
with γ̇ i
:= ∂t γi.
We will sometimes denote this vector by γ̇γ(t0 ) In conclusion given a
curve γ we can produce a tangent vector at p0 = γ(t0 ) that is γ̇ i (t0 )∂i |p0 .
Natural question: every tangent vector arise from a curve?
Lemma 4.7. Let p ∈ M then every Xp ∈ Tp M is the tangent vector to some curve
Proof. Consider a coordinate chart (U, ϕ) centered at p (that is ϕ(p) = 0) the vector
Xp = ai ∂i |p . We want to construct a curve γ : (−, ) → U such that γ̇ i (0) = ai and
γ(0) = p. To this aim we construct γ̂ = (ta1 , .., tan ) from which we have
γ(t = 0) = ϕ−1 (γ̂(t = 0)) = ϕ−1 (0) = p
and
γ̇γ(0) = γ̇ i (0)∂i |γ(0) = ai ∂i |γ(0)=p
18
5. Vector bundles
We want now to extend the notion of vector field on Rn to a general smooth manifold
M , namely something that evaluated at each point p produces an element of Tp M
“smoothly”. This goal can be achieved using the general notion of vector bundle that
we introduce int he following:
Definition 5.1. Consider M and E smooth manifolds a a smooth surjective map π :
E → M . If
def
• Ep = π −1 (p) named the fiber over p, is a real vector space of dimension k
• for every p ∈ M we can find a neighborhood of p and a diffeomorphism
Φ : π −1 (U ) → U × Rk
s.t. Φ|π−1 (p) : Ep → {p} × Rk ∼ Rk is a linear isomorphims and pr1 ◦ Φ = π. The
diffeo φ satisfying this requirements is called local trivialization
we will call the set of data (E, M, π, {(U, φ)}) a (smooth) vector bundle of rank k (we
will just denote it by π : E → M or jut E sometimes in the future).
What happen when we change trivialization?
Lemma 5.2. Given two trivialization (U, Φ) and (V, Ψ) of a vector bundle of rank k
with U ∩ V 6= ∅, then we can find a smooth map g : U ∩ V → GLk (R) s.t. for every
p ∈ U ∩ V we have
Φ ◦ Ψ−1 (p, v) = (p, g(p) · v)
where we denoted by g(p) · v the action of an element of GLk (R) on an arbitrary element
v ∈ Rk
Proof. We know that pr1 ◦ Φ = π = pr1 ◦ Ψ thus the following diagram commutes
Ψ−1 Φ
(U ∩ V ) × Rk π −1 (U ∩ V ) (U ∩ V ) × Rk
pr1
π
pr1
U ∩V
19
thus
up,a ⊂ U
which shows that {up,a } is a basis for a topology on M .
Proof. Consider a countable basis of coordinate opens sets {Ui }. Being T Ui homeomor-
phic to R2n is second countable. For each T Ui chooose a countable basis, and thus also
T M will be second countable being the basis for M choosen countable.
It is then easy to prove that T M is Hausdorff. If (p, Xp ) and (q, Xq ) live on two
different fibers then we can find two neighborhoods for p and q we call U1 and U2
separating the points so that T U1 and T U2 separates (p, Xp ) and (q, Xq ). If we consider
points on the same fiber we can easily use the isomorphism with Rn
Thus T M is also locally Euclidean by construction.
We know what happens when we change coordinate chart from (U, ϕ) to (V ψ) on M
and Tp M , putting these stuff together we obtain
∂ x̃1 (x) i
Ψ ◦ Φ−1 (x1 , .., xn , a1 , .., an ) = (x̃1 (x), .., a , ...)
∂xi
that is clearly smooth. We thus can construct in this way a smooth atlas {T Ui , Φi }.
prove 2: By construction
prove 3: It is easy to show that Φ̃ = (ϕ−1 × idRn ) ◦ Φ : π −1 (U ) ⊂ T M → U × Rn is a
trivialization map
Lemma 5.9. In every coordinate chart, the component functions of X ∈ X(M ) are
smooth functions. Moreover a vector field, for any f ∈ C ∞ (M ), naturally define the
function X · f : M → R defined by (X · f )(p) = Xp · f ; this one is smooth.
Proof. Given the chart (U, ϕ) with coordinate function x = (x1 , .., xn ) for M we construct
the coordinate rep of X : M → T M
thus the component functions âi on U must be smooth functions, thus in the given
[
coordinate chart the coordinate representation X ·f
[
X · f = âi (x)(∂i fˆ)(x)
that is obviously smooth being âi and ∂i f smooth functions. For this reason we think at
a vector field in a given coordinate chart as something of the form ai ∂i with ai smooth
functions on M
22
Take now the dual space of Tp M ; it is called the cotangent space and denoted by Tp∗ M .
In a given coordinate open setU , with coordinate functions (x1 , .., xn ) we have the co-
∂ i
ordinate basis for Tp M given by { ∂x i p }, and we denote the dual one by { dx p }. Thus
∂xj
∂i |p → ∂˜i |p = ∂ |
∂ x̃i ϕ(p) j p
∂ x̃i
ai → ãi = ∂xj ψ(p)
aj
∂ x̃i
dxi |p → dx̃i |p = ∂xj ψ(p)
dxj |p
∂xj
ωi → ω̃i = ω
∂ x̃i ϕ(p) j
2
• Given a smooth vector field X one has that ω(X)(p) = ωp (Xp ) is smooth.
Given a smooth function we can naturally produce a covector field df called the
differential of f .The symbol d will be also discussed in the following it willplay a crucial
role; at thsi stage let s just view df as a unique symbol ; df is a covector field defined
such that
dfp (Xp ) = Xp · f
In a given coordinate chart we have
dfp (∂i |p ) = (∂i f )(p)
so locally we may write df = ∂i f dxi . Note that by construction df is a smooth map from
M to T ∗ M . The “dual” manouvre with respect to the pushforward is named pullback:
give the smooth map F : M → N we construct
F ∗ : TF∗ (p) N → Tp∗ M
defined by
(F ∗ ωF (p) )(Xp ) = ωF (p) (F∗ Xp )
Observation If F : M → N is as usual a smooth map and ω ∈ Ω1 (N ) then F ∗ ω ∈
Ω1 (M ). It is easy to check in a coordinate system. Note that with respect to the
vector field case we don’t have an ambiguity (that can be caused for example by an non
injective map) and everything is well defined since since we are pulling back objects. We
will sometimes use the following notations to specify the same thing
(F ∗ (ωF (p) ))p = (F ∗ ω)p = F ∗ (ωF (p) )
7. Tensors
We start defining tensors on a real finite dimensional vector space V and we then
generalize the construction to a general manifold.
Definition 7.1. A covariant k tensor is a multilinear map:
| × {z
τ :V .. × V} → R
k times
Observe that T k V and Tp V and Tpk V can be endowed with the structure of a real
vector space. We can in principle be more general and for example consider multilinear
map from V × W → R with both V and W real finite dimensional vector spaces.
Being i < j only one term survive , namely the first one, proving that λ1,2 = λ(b2 , b1 ).
In this way we prove that all λij vanishes proving again linear independence. Do our
basis span Λk V ? Consider a general ρ ∈ Λ2 V and define ρ0 := ρ(bi , bj )β i ∧ β j the on
every pair (bk , bm ) one finds ρ = ρ0
Proposition 7.7. Given V and W finite dim vector spaces and S any vector space.
Given a BILINEAR map f : V ×W → S there is a unique LINEAR map f˜ : V ⊗W → S
such that the following diagram commutes:
f
V ×W S
f˜
π
V ⊗W
where π(v, w) = v ⊗ w
Proof. Sketchy: we first extend f uniquely to a linear map f¯ : R < V × W >→ S defined
byf¯(v, w) = f (v, w) whenever (v, w) ∈ V × W ⊂ R < V × W >. We then note that
the subset I defined previously is contained in the kernel of f¯ therefore f¯ descends to a
linear map f˜ = R < V × W > /I → S satisfying f˜ ◦ π = f by construction. Since every
element of V ⊗ W can be written as linear combination of object of the form v ⊗ w
and on such elements f˜ is uniquely determined by f˜(v ⊗ w) = f¯(v, w) = f (v, w), then
uniqueness follows.
Proof. In order to simplify the notation we consider the case V = W . First define the
map f : V ∗ × V ∗ → Bil(V, V ) as follows:
g(B) = B(bi , bj )β i ⊗ β j
(g ◦ f˜)(ρ) = f˜(ρ)(bi , bj )β i ⊗ β j
Let us better understand the meaning of those objects with some example.
• T00 V = R
• T 1V = V ∗
• T1 V = V
∗
• on V = Rn det(v1 , .., vn ) ∈ Λn V ⊂ V .. ⊗ V }∗
| ⊗ {z
n times
• a scalar product on V is an element of Σ2 V ⊂ V ∗ ⊗ V ∗ of element of the form
gij β i ⊗ β j with gij = gji
What about End(V )? We note that there is a canonical isomorphism φ : End(V ) → T11 V
given by (φ(L))(v, α) = α(L(v)); this map is injective and since dimEnd(V ) = n2 =
dim(V ∗ ⊗ V ∗ ) we get the desired result. In components, given T ji β i ⊗ bj we construct
the endomorphism LT (v) := T ji β i (v) ⊗ bj . In analogy with this construction we have
k
Proposition 7.10. the vector space Tp+1 is canonically isomorphic to the space of mul-
tiliear maps
∗
| × {z
V .. × V} × V
| × {z.. × V }∗ → V
k times p times
where the components T j1 ..jr i1 ..ik are smooth functions. As usual we will denote by
Tp ∈ Trk (Tp M ) the value of a tensor field at a point namely
Tp (X1p , ., Xkp , ω1p , .., ωrp ) = T (X1 , .., Xk , ω1 , .., ωr )(p)
Consider now an element S of T11 (M ). In a given coordinate chart it looks like
S = S ji dxi ⊗ ∂j . When it acts on a vector v = v i ∂i and ω = ωi dxi , using ∂i (dxj ) = δik =
dxj (∂i ) we have
S(v, ω) = S ji v k ωm dxi (∂k )dxm (∂j ) = Sij v i ωj
It is kind of evident now that this map is C ∞ (M ) linear. This can obviously generalized
to any tensor field. So a (1,1) tensor field is a C ∞ (M ) multilinear map from X(M ) ×
Ω1 (M ) → C ∞ (M ).There is something more:
Lemma 7.12. a map
X(M ) × .. × X(M ) × Ω1 (M ) × .. × Ω1 (M ) → C ∞ (M )
| {z } | {z }
k times r times
or
X(M ) × .. × X(M ) × Ω1 (M ) × .. × Ω1 (M ) → X(M )
| {z } | {z }
k times r−1 times
is induced by a rank (k, r) tensor field if and only if it is multilinear over C ∞ (M )
Let ρ ∈ T k (N ) and F : M → N a smooth map. We can pullback covariant k tensor
in analogy with cavariant vector as follows:
(F ∗ ρ)p (X1 , .., Xk ) = ρF (p) (F∗ X1 , .., F∗ Xk )
with X1 , .., Xk vector fiedlds on M .
Poperties:
• F ∗ is linear over R
• F ∗ (f ρ) = (f ◦ F )F ∗ ρ
• F ∗ (ρ ⊗ ω) = F ∗ ρ ⊗ F ∗ ω
• (G ◦ F )∗ = F ∗ ◦ G∗
7.1. Differential forms and integrations.
Definition 7.13.
Λk M = tp Λk (Tp M )
Sections of Λk M are named differential forms of rank k and the set of those objects
is denoted by Ωk (M ). A differential k form can be written, in a given coordinate chart,
as
1
ω = ωi1 ..1k dxi1 ∧ .. ∧ dxik
k!
Lemma 7.14. F : M → N a smooth map, and consider the coordinate functions
(x) = (x1 , .., xn ) for M and (y) = (y 1 , .., y m ) for N (defined locally obviously)
then:
F ∗ (ωi1 ..ik dy i1 ∧ .. ∧ dxyk ) = (ωi1 ..ik ◦ F )d(y i1 ◦ F ) ∧ .. ∧ d(y ik ◦ F )
Proof. it follows easily from the properties discussed previously see Lemma (6.5).
This Lemma has a useful corollary:
30
Corollary 7.15. Let F : M → N smooth map between n manifolds then choose the
coordinates functions (x) on U ⊂ M and (y) on V ⊂ N we have on U ∩ F −1 (V )
F ∗ (f dy 1 ∧ .. ∧ dy n ) = (f ◦ F ) det(∂i F j )dx1 ∧ .. ∧ dxn
Proof. Note that
ω 1 ∧ .. ∧ ω n (X1 , .., Xn ) = det(ω i (Xj ))
as one can easily observe in an example for n = 3 (discussed in the exercise session).In
particular using the previous Lemma we have
F ∗ (f dy 1 ∧ .. ∧ dy n ) = (f ◦ F )d(F 1 ) ∧ .. ∧ d(F n )
with dF i = y i ◦ F . The components of d(F 1 ) ∧ .. ∧ d(F n ) are
d(F 1 ) ∧ .. ∧ d(F n )(∂1 , .., ∂n ) = det(∂i F j )
thus
d(F 1 ) ∧ .. ∧ d(F n ) = det(∂i F j )dx1 ∧ .. ∧ dxn
For any smooth manifold there is a differential operator
d : Ωk (M ) → Ωk+1 (M )
We will write it in coordinates now, in the following we will show a more general defini-
tion.
1
dω := (dωi1 ..ik ) ∧ dxi1 ∧ .. ∧ dxik
k!
that can be written as
1
dω := (∂ ω )dxj ∧ dxi1 ∧ .. ∧ dxik
(k + 1)! [j i1 ..ik ]
where [ji1 , , ik ] means that we are taking the totally alternating part on the indices, that
is if we exchange two of them we get a minus sign morally.
Properties of the exterior derivative:
• d is well defined independently of the coordinates chosen (we will see that later
with the coordinate independent formula)
• d is linear over R
• d2 = 0
• ω ∈ Ωk (M ) and ρ ∈ Ωn (M ) then
d(ω ∧ ρ) = (dω) ∧ ρ + (−1)k ω ∧ dρ
Let’s see an example.Consider on R3 the one form
ω = f dx + gdy + hdz
then
dω = df ∧dx+dg∧dy+dh∧dz = (∂x g−∂y f )dx∧dy+(∂x h−∂z f )dx∧dz+(∂y h−∂z g)dy∧dz
where you see the components of the curl of the vector field f ∂x + g∂y + h∂z (modulo
signs).
Let’s go back to vector spaces: Let V be a finite dimensional vector space then an
orientation for V is an equivalence class of ordered basis defined to be equivalent if they
are related by a positive determinant matrix. A basis is called positively oriented if it
belong to the chosen orientation.
31
Lemma 7.16. Consider V a real vector space of dimension n and Ω ∈ Λn V then the
set of bases (b1 , .., bn ) so that Ω(b1 , .., bn ) > 0 is an orientation for V
Proof. Given (b1 , .., bn ) (with dual basis (β 1 , .., β n )) and (b01 , .., b0n ) two basis related
by
b0i = Aji bj
and so that Ω(b1 , .., bn ) = cβ 1 ∧ .. ∧ β n ((b1 , .., bn )) = c > 0 then
Ω(b01 , .., b0n ) = c det(β i (b0j )) = c det(Aij )
then Ω define a class of basis with the same orientation by the previous formula
We will say that given an orientd vector space Ω ∈ Λn V is positively oriented if
itinduces the same orientation of V . In the case of a manifold M an orientation is a
choice of orientation for each Tp M that we want to be smooth in the sense that in a
neighborhood of point we cna always find a local frame pointwise positively oriented. A
∂
coordinate chart (U, ϕ) is said to be positive oriented if the coordinate basis { ∂x i
|p } for
Tp M is positively oriented for each p ∈ U .
Proposition 7.17. Let M be a smooth manifold of dimension n. A nowhere vanishing
Ω ∈ Ωn (M ) (sometimes called volume form) determines a unique orientation of M for
which Ω is positively oriented at each point.
Proof. In a coordinate chart (U, ϕ) with U connected, we can write Ω = f dx1 ∧ .. ∧ dxn
with f nowhere vanishing and we have
Ω(∂1 , .., ∂n ) = f
thus on U it is alway positive or negative oriented. If it’s negative oriented it is enough
to change the sign of one of the coordinate representation of any point and obtain thus
a continuous orientation.
One can also prove that conversely an orientation for M induce a nonvanishing n form
positively oriented at each point. Consider now a top form defined on a compact domain
D ⊂ Rn
ω = f dx1 ∧ .. ∧ dxn
To avoid convergence issues we will always require that ω is compactly supported in U
and we define Z Z Z
ω= ω= f dx1 ..dxn
U D D
where supp(f ) ⊂ D ⊂ U
Proposition 7.18. Suppose U , V are open subsets of Rn , F : V → U is an orientation-
preserving diffeomorphism, then
Z Z
ω= F ∗ω
U V
Proof. This is just a consequence of 7.15 and the change of coordinates for an integral
(absolute value of the determinant of the Jacobian).
We are ready to consistently define the integral on a manifold M . Given ω compactly
supported in an oriented coordinate chart (U, ϕ)
Z Z
ω := (ϕ−1 )∗ ω
M ϕ(U )
32
By construction this definition does not depend on the choice of oriented charts whose
domain contain suppω. If the n form is supported on M oriented we can use the partition
of unity {ψi } subordinate to a coordinate charts (Ui , ϕi ) covering M considering on each
Ui the one form ψi ω supported there, and then summing overall i
Z XZ
ω := ψi ω
M i M
Note: One can prove that the final answer does NOT depend on the choice of coordi-
nates charts and the partition of unity.
• F l0X = idM
The vector field X is called the infinitesimal generator of the flow and it is strictly
related to the concept of symmetries of a manifold. Let’s see that with examples:
Example 8.7.
X = x∂x + y∂y ∈ X(R2 )
then
F ltX (x0 , y0 ) = (et x0 , et y0 )
and we will say that it generates the dilations
Example 8.8.
X = ∂x ∈ X(R2 )
then
F ltX (x0 , y0 ) = (x0 + t, y0 )
and we will say that ∂x generates the translation along the x axis.
Observation: Fixed t ∈ R we can define
Mt = (p ∈ M : (t, p) ∈ D)
notice it can be empty. Then there is a well defined map
Φt : Mt → M
The map Φt induced by a vector field X is sometimes called the local 1 parameter group
of local diffeomorphisms generated by X, and X is called the infinitesimal generator.
This terminology is motivated by (8.6).
Definition 8.9. Given f ∈ C ∞ (M ) and X and Y vector fields we have
• (LX f )(p) = ∂t |0 (f ◦ F ltX (p))
X) Y
• (LX Y )(p) = ∂t |0 ((F l−t ∗ F ltX (p) ) the Lie derivative of a function and a vector
field respectively.
In order to better understand the Lie bracket of a vector field we take a small detour
and discuss a famous algebraic object. Given two vectors fields X and Y we can induce
a third one, we denote by [X, Y ], by the following algebraic operation called Lie bracket
(or sometimes commutator) defined at a point by:
[X, Y ]p f = Xp · (Y · f ) − Yp · (X · f )
Proposition 8.10. We have:
• [X, Y ]p ∈ Tp M
34
The first thing to observe (and prove) is that dα is a tensor thus a C ∞ (M ) multilinear
map. Let’s compute its components in a coordinate chart
1
dα = aij dxi ∧ dxj
2
then
dα(∂i , ∂j ) = aij = ∂i α(∂j ) − ∂i α(∂j ) = ∂i αj − ∂j αi
In general we have
i+1 X α(X , ..X
P
dα(X1 , .., Xk+1 ) := i (−1) i 1 i−1 , Xi+1 , .., Xk+1 )
8.1. Sumbmanifolds.
Definition 8.18. A subset S of a manifold M (of dimension n) is called a (regular)
submanifold of dimension k if for every p ∈ S we can find a coordinate chart (U, ϕ) in
the maximal Atlas, with p ∈ U and coordinate functions (x1 , .., xn ) such that U ∩ S
(better to say U
\ ∩ S ) is defined by the vanishing of n − k coordinates functions that we
can assume to be the last n − k coordinates functions without loosing generalities.
We call (U, ϕ) adapted relative to S. Note that ϕ|U ∩S = (x1 , .., xk , 0, .., 0) and we can
define using the projection on the first k components pr(k)
ϕS := pr(k) ◦ ϕ : U ∩ S → Rk
so that (U ∩ S, ϕS ) is a coordinate chart for S with the subspace topology.
Proposition 8.19. Let S be a regular submanifold of N and U = (U, ϕ) a collection
of compatible adapted charts of N that covers S. Then (U ∩ S, ϕS ) is an atlas for S.
Therefore, a regular submanifold is itself a manifold.
Submanifolds are typically presented as images or level sets of smooth maps. A level
set of a map F : M → N is a subset
F −1 (q) := {p ∈ M s.t. F (p) = q}
for some q ∈ M . In the case N = Rn we call ξ(F ) = F −1 (0) the zero set of F .
Definition 8.20. Given F : M → N we will say that q ∈ N is a regular value of F
if either q is not in Im(F ) or for every p ∈ F −1 (q) we have F∗ |p : Tp M → TF (p)N is
surjective. The preimage of regular value is called a regular level set.
Before we proceed with our analysis let’s state a couple of natural results that is the
inverse function theorem and its generalization to case of manifolds:
Theorem 8.21. Let W be an open subset of Rn and F : W → Rn a smooth map defined
on an open subset. For any point p in W the map F is locally invertible at p if and only
if the Jacobian determinant ∂i F j is not zero.
that can be generalized for manifolds as
Theorem 8.22. Let F : M → N a smooth map between two manifolds of the same
dimension, and p ∈ M . Suppose we have some charts (U, ϕ) with coordinates functions
(x1 , .., xn ) around p and (V, ψ) with coordinates functions (y 1 , ..., y n ) around F(p) with
F̂ j
F (U ) ⊂ V . Then F is locally invertible at p if and only if det ∂∂x i |ϕ(p) is nonvanishing.
Proof. (*) Consider the charts (U, ϕ) with coordinates functions xa = x1 , .., xm on M
and (V, ψ) with coordinate functions (y i ) = y 1 , .., y n and centered at a point q with
38
F −1 (V ) containing U and F −1 (q). Observe that o that we can view F −1 (q) as the
zero set of ψ ◦ F since ψ(q) = 0. Note now that since the regular level set is assumed
to be non empty it means that the differential map at F −1 (q) is a surjection then we
must have m ≥ n. Call now ψ ◦ F = F 1 , .., F n with F i smooth functions on M : Note
that F i (p) = 0 for every p ∈ F −1 (q). By regularity me must then have that every
p ∈ F −1 (q) are such that the jacobian matrix ∂a F j |p has rank n. Without loosing of
generality we assume last n × n block of the Jacobian is non singular (we will denote it
by ∂â F i |p ). Then we use F 1 , .., F n as the last coordinates. In particular we claim that
we have a neighborhood Up of a fixed p ∈ F −1 (q) such that (Up , ϕF ) is a chart, with
ϕF (r) = (x1 , ..., xm−n , F 1 , .., F m ) for some r ∈ Up . This chart is well defined because of
the inverse function theorem and it is an adapted coordinate chart for F −1 (q).
i I 0
∂a F |p =
0 ∂â F i |p
where â = m + 1, .., n
Let us now characterize regular submanifolds.
Definition 8.25. Let the F : M → N with dimensions m and n. The rank of F at p is
the rank of the pushforward at a point that is rankF∗ |p that is dimIm(F∗ |p ).
• If m ≤ n and for every p ∈ M we have rankF = m, we say that F is an
immersion
• If m ≥ n and for every p ∈ M we have rankF = n, we say that F is an
submersion
• If m = n and for every p ∈ M we have rankF = n = m, we say that F is an
local diffeomorphism
Theorem 8.26. Let F : M → N a smooth map of manifolds and q ∈ N . Suppose
F has constant rank k in a neighborhood of p ∈ M .The we can find a coordinate (U, ϕ)
centered at p and (V, ψ) centered at F (p) so that F̂ (x1 , .., xm ) = (x1 , .., xk , 0, .., 0)
This theorem called the constant rank theorem has nice consequences for example if
f : M → N has constant rank in a neighborhood of a level set, then the level set is a
regular submanifold.
Consider now a one to one immersion F : N → M ; the image F (N ) is called immersed
submanifold. In general its topology and smooth structure has nothing to do with the
one on M and has to be considered as extra data. This observation leads to the following
definition:
Definition 8.27. A map F : M → N is called an embedding if it is a one-to-one
immersion and f (M ) with the subspace topology is homeomorphic to M trough f
Let’s see some example of non embedding to understand which type of situation we
want to avoid:
Example 1 f : R → R2 defined by f (t) = (t2 , t3 ).This map is one to one but not an
immersion since the differential map at zero f∗ |0 is not injective.
Example 2 f : R → R2 defined by f (t) = (t2 − 1, t3 − t).This map is NOT one to one
because f (1) = f (−1) and an immersion since the differential map at zero f∗ |0 is always
injective.
Example 3 Consider the map given in the picture. It is a one to one immersion but
39
the topology induced on R2 doesn t match the original topology because, for example,
there are point close to f (p) corresponding in R to point far away from p. Thus M and
f (M ) in this example are not homeomorphic thus it is not an embedding.
Theorem 8.28. If F : M → N is an embedding then F (M ) is a regular submanifold.
Proof. (*) By the constant rank theorem we know we can choose local coordinates charts
(U, ϕ) and (V, ψ) such that
F̂ (x1 , .., xm ) = (x1 , .., xm , 0, .., 0)
We may have trouble since V ∩ f (N ) may be larger then F (U ). It is here where we use
the subspace topology to find a V 0 such that V 0 ∩ F (N ) = F (U ) (we skip this details)
and thus on V ∩ V 0 we can construct the adapted coordinate chart as before.
= −f (p)X̃σ(p) + f (p)σ∗ Xp
= f (p)∇X σ(p)
(2) The second relation is a direct consequence of (2)
(3) The third relation comes from proposition (9.2) and the definition of ∇;
d X̃
∇X (f σ) = dt t=0 (F l−t ◦ f σ ◦ F ltX (p))
d X̃ ◦ f (γ (t)) σ(γ (t))
= dt t=0 F l−t p p
d X̃
by prop (9.2) = dt t=0 f (γp (t))F l−t (σ(γp (t))
(4) The last relation again comes from the fact that F ltX is linear
Definition 9.5. A linear connection on a vector bundle is a choice of a linear isomor-
phism among fibers, equivalently a linear horizontal lift.
We are ready now to work in coordinates. Consider the coordinates functions (x1 , .., xn )
for M , the coordinate basis for Tp M , and the trivialization map Φ associated to the local
frame (α ) with α = 1, .., k; writing the vector field X = ai ∂i the section as σ = σ α α ,
we have using (9.4)
∇X σ = ∇ai ∂i (σ α α )
by (1) = ai ∇i (σ α α )
by (3) = ai ∂i (σ α )α + ai σ α ∇i α
where ∇∂i := ∇i . Let’s focus now on the element ∇i α . Due to the linearity of the
construction we have that
∇i α = B(∂i )βα β
for some objects B(∂i )βα we now analyze. Fixed the index i this is nothing else that a
k × k matrix associated to an endomorphism of Rk . Fixed the indices α, β and due to
the first two of (9.4) we view B βα as maps X(M ) → C ∞ (M ) linear over C ∞ (M ). In
conclusion we have that locally B βα ∈ Ω1 (M, End(Rk )). We will call it connection one
form and just denote its components by (Bi )βα or simply Bi βα and call them Christoffel
symbols. It is common and useful to write
∇i σ α = ∂i σ α + Bi αβ σ β
Observe now that this construction depends on the choice of coordinates and triv-
ialization. The natural question is what happens when we change trivialization and
coordinate chart. Suppose on U we have the coordinate chart (x1 , .., xn ) and local frame
1 , .., k while on V we have (x̃1 , .., x̃n ) and e ˜1 , .., ˜k . We know that σ̃ β = τ βα σ α We
want to compare on U ∩ V the expressions
ai (∂i σ α + Bi αβ σ β )α
and
ãi (∂˜i σ̃ α + B̃i αβ σ̃ β )˜
α
where B̃ is the connection one form one would obtain in the trivialization induced by the
frames ˜α and the coordinate x̃, that is B̃i αβ σ̃ β ˜α = B̃(∂˜i )αβ σ̃ β ˜α d. Remembering that
j
∂˜i = ∂x ∂ and observing that the one form connection B̃ can be viewed as a one form
∂ x̃i j
taking values in the algebra of k × k matrices (i.e. Mk ) thus recalling that τ : M → Mk
one has
∂xj
∂˜i σ̃ α + B̃i αβ σ̃ β = ∂ j (τ α γ
γ σ ) + (τ )β
γ ( B̃ )
j β
α γ
σ
∂ x̃i
from which we get, combining with α = τ βα ˜β , that
∂xj −1 α δ γ ∂xj −1 α
(B̃i )αβ = (τ ) δ (B ) τ
j γ β + (τ ) δ ∂j τ δβ
∂ x̃i ∂ x̃i
This ugly formula is often written in a compact form, suppressing the matrix indices
and using differential forms notation, as
B̃ = τ −1 Bτ + τ −1 dτ
43
When dealing with T M instead of E we have a natural coordinate basis for the fiber
{∂i } and we denote the connection one form by Γjk = (Γi )jk dxi = Γi j k dxi . In this case
we have τ ij = ∂˜i xj and the connection is called affine connection. We have defined the
covariant derivative associated to an affine connection on vector fields only so far. We
now generalize to every tensor field T ∈ Tlk M by the following.
(1) k = 0, l = 1 we have that ∇X Y is defined as before, in components by
∇i Y j = ∂i Y j + Γ i j k Y k
(2) k = 0 = l we define it as
∇X f = X · f
(3) T = F ⊗ G then
∇X T = ∇X F ⊗ G + F ⊗ ∇X G
(4) Denoting by Yi and ω j with i = 1, .., k and j = 1, .., l two sets of vector fields
and covectors (NOT the components of a (co)vector field) we require
(∇X )F (Y1 , .., Yk , ω 1 , .., ω l ) = X · F (Y1 , .., Yk , ω 1 , .., ω l )
−F (∇X Y1 , Y2 , ..) − ..
−F (Y1 , .., Yk , ∇X ω 1 , ..) − ..
Observe that the extension is unique. We can in fact define the covariant derivative on
covectors by
(∇X ω)(Y ) = X · ω(Y ) − ω(∇X Y )
and once we know it again using the number (4) we can uniquely extend the construction
to every tensor field.In components in the coordinate basis we have
∇i ωj = ∂i ωj − Γik j ωk
and in general
∇i T j1 ..jlm1 ..mk = ∂i T j1 ..jlm1 ..mk
+Γij1j T j..jl m1 ..mk + ..
−Γimm1 T j1 ..jlm..mk − ...
It is often useful to work along curves. Consider γ : I → M a smooth curve and define
a vector field along γ as a smooth map Ỹ : I → T M s.t. Ỹ (t) ∈ Tγ(t) M . A vector field
along γ is called extendible if exists a vector field Y at least in an open subset of M
containing γ, such that Y (γ(t)) = Ỹ (t). This construction naturally extends to all tensor
fields. An affine connection induces on the space of vector field along a curve a unique
linear operator Dt : X(γ) × X(γ) → X(γ) satisfying Dt (f Ỹ ) = f˙Ỹ + f Dt Ỹ Consider then
an extendible vector field along γ and define
Dt Ỹ (t) := ∇t Y := ∇γ̇(t) Y
In components we have
(3) ∇γ̇(t) Y = γ̇ i ∂i Y k (γ(t))∂k |γ(t) + γ̇ i Γi k j Y j ∂k |γ(t)
| {z }
Ẏ k
In the following we will deal with vector field along curves obtained by restriction thus
extendable.
44
Given ω and α one form it is sometimes useful to write their symmetric product as
follows
1
ωα := (ω ⊗ α + α ⊗ ω)
2
With this notation in mind we write the metric in a given coordinate chart as
g = gij dxi dxj
The pair (M, g) is called Riemannian manifold
Warning: A common mistake made by novices is to assume that one can find coordi-
nates near p such that the coordinate vector fields ∂i are orthonormal. The coordinate
basis for Tp is a canonical choice but sometimes not the best one. At each point for
example one can find an orthonormal basis. We define then an othonormal frame as a
set of sections of T M (e1 , .., ep ) such that
1 0 .. 0 0
0 1 ... 0 0
g(ea , eb ) = δab = 0 0 . . . 0 0
0 0 ... 1 0
0 0 ... 0 1
Dually we a have the orthonormal coframe (pointwise basis for Tp∗ M ) (1 , .., n ) where
by construction b (ea ) = δab . Thus we could write in this basis
g = δab a b
We could always write ea as linear combination over C ∞ M of ∂i that is
ea := eia ∂i
as well as
a = eai dxi
for some functions (or matrices if you prefer) eia and ebj that are such that
eai eib = δba , eia eaj = δji
being ea and a (as well as dxi and ∂i ) pointwise dual basis. Observe that by construction
we have
g = δab a b = δab eai ebj dxi dxj
thus
(4) gij = δab eai ebj , and viceversa δab = gij eia ejb .
Those objects are the main ingredients of the Cartan moving frame theory and other
physical models like supergravity and superstring theory, because, in some sense, they
encode the information of the metric in a more geometrical way.
Musical isomorphism:
Definition 9.12.
[ : T M → T ∗M
by [(X :) = X [ is the covector such that X [ (Y ) = g(X, Y ) for every Y ∈ T M
Working in component we have Xi := X [ (∂i ) = gij X j . Being the metric invertible
one can also define g −1 that in components looks like g −1 = g ij ∂i ⊗ ∂j with g ij = g ji
where gij g jk = δik . This induces an inner product on the cotangent space Tp∗ M , namely:
< ωp , αp >p := g −1 (p)(ωp , αp ) = g ij (p)ωi (p)ωj (p)
The inverse map of the flat map is given by the sharp map
] : T ∗M → T M
47
What happen in a different frame?? Consider the orthonormal frame {ea } and define
c by
the smooth functions fab
c
[ea , eb ] = fab ec
Then one gets
1 c b0 cc0 a0 cc0
Γcab = (fab − fac0δ δb0 b − fbc0δ δa0 a )
2
Consider now an embedded surface in R3 etc...
Suppose M and M 0 are smooth manifolds and ∇0 an affine connection on M 0 and F :
M → M 0 is a diffeomorphism then F ∗ ∇0 defined by
(F ∗ ∇0 )X Y = (F −1 )∗ (∇F∗ X F∗ Y )
is an affine connection on M called the pullback connection.
Proposition 9.19 (Naturality of the LC connection). Suppose (M, g) and (M 0 , g 0 ) are
Riemannian with Levi Civita connections ∇ and ∇0 . If F : M → M 0 is an isometry
then F ∗ ∇0 = ∇
Ledt’s now go back to geodesics for the LC connection.
Definition 9.20. Let (M, g) be a Riemannian manifold. An admissible curve γ in M
is said to be a minimizing curve if Lg (γ) ≤ Lg (γ̃) for every admissible curve γ̃ with the
same endpoints.
Now we state the following result whose proof involve calculus of variations tools we
are not going to discuss in this notes
Proposition 9.21. In a Riemannian manifold, every minimizing curve is a geodesic
when it is given a unit-speed parametrization.
51
It is easy to see that the literal converse is not true, because not every geodesic segment
is minimizing. For example, every geodesic segment on S 2 that goes more than halfway
around the sphere is not minimizing, because the other portion of the same great circle
is a shorter curve segment between the same two points. What can be proved by using
Riemann normal coordinate is that geodesics γ are locally minimizing that is for every
t0 ∈ I we can find a neighborhood I0 such that γ restricted to I0 is minimizing
9.4. Geodesics in Riemannian geometry (**). We have already discussed the ex-
istence and uniqueness of a geodesics given the initial speed and point. Now we will
focus on the case of geodesics obtained out of the Levi Civita connection only. Let now
denote by ΥX p the geodesics with initial point p and speed Xp , be carefull this is not the
integral curve for X. From the construction it is easy to see the the following relation
holds:
ΥaX X
p (t) = Υp (at)
whenever either side is defined. Consider now E = {X ∈ X(M ) s.t. ΥX p } is defined in
an interval containing [0,1] for all p; then construct the following map
exp(X) = ΥX
• (1)
or simply expp (X) when we want to specify also the initial point
Proposition 9.22. For each X the geodesics ΥX
p is given by
ΥX
p = expp (tX)
F
M N
Consider now T0 (Tp M ); it can be canonically identified with Tp M . If we consider the
differential of the exponential map at the origin that is d(expp )|0 it can be viewed as a
map from Tp M to itself. Consider now the curve on T pM given by c(t) = tX and use it
to compute the differential of the exponential map:
d d d
d(expp )|0 (X) = |0 expp (c(t)) = |0 expp (tX) = |0 ΥX p (t) = X
dt dt dt
We reassume this result in the following proposition
Proposition 9.24. The differential at the origin of the exponential map is the identity
map on Tp M
This result is crucial since by the inverse function theorem we can say that there
is a local diffeomorphism from Tp M to M around the origin. Better to say , there
are neighborhoods V of 0 in Tp M and U of p in M such that expp : V → U is a
diffeomorphism. A neighborhood U of p that is diffeomorphic trough the exponential
map to some neighborhood V of 0 in Tp M is called a NORMAL NEIGHBORHOOD of
52
U
Observe that given this coordinate we have that the coordinate basis for Tp M is or-
thonormal so in some sense we combine the beauty of the coordinate and orthonormal
frame. Nice things happens in this coordinates:
Proposition 9.25. Given a Riemannian manifold (M, g) and a normal coordinate chart
centered at p we have
(1) In this coordinate chart gij (p) = δij
(2) For every X = ai ∂i the geodesics ΥX p is represented in this coordinates by the
1 n
line (ta , .., ta )
(3) the Christoffel symbols at p in this coordinates vanish
Proof.
From this proposition it is evident that locally Riemannian geodesic substitute the
notion of straight lines. One can easily prove that for S 2 grate circles are geodesics.
R(∂i , ∂j , ∂k ) = Rlkij ∂l
or if you prefer
Rlkij = dxl (R(∂i , ∂j , ∂k ))
53
The Riemann tensor satisfies many identities one can easily recover out of its defi-
nition. We list them in the following by using the components notation and using the
(4,0) tensor obtained by the Riemann one raising an index Rijkl = gim Rmjkl :
•
Rijkl = −Rjikl Rijkl = −Rijlk
•
Rijkl = Rklij
• First Bianchi idenity
Rijkl + Rkijl + Rjkil = 0
• Second Bianchi identity
∇m Rijkl + ∇i Rjmkl + ∇j Rmikl = 0
Other interesting tensors are the Ricci tensor Ric that is a (2, 0) symmetric tensor defined
in components as
R = Rij dxi dxj
with Rij = Rk ikj and the scalar curvature S that is its trace S = g ij Rij . A Riemannian
manifold is called an Einstein manifold if
Rij = λgij
for some λ ∈ R
Definition 9.29. A vector field X is named Killing vector field if LX g = 0
Proposition 9.30. If X = ak ∂k is a Killing vector field then ∇i aj + ∇j ai = 0
Aα = gAβ g −1 − (dg)g −1