Geodiff Besancon
Geodiff Besancon
Geodiff Besancon
geometry
Besançon - Autumn 2022
Jack Borthwick
December 9, 2022
Avant-propos
Ces notes ont été rédigées dans le cadre d’un cours doctoral de 10h donné à l’École
Doctorale Carnot-Pasteur. Elles s’adressent avant tout aux doctorants désireux d’avoir
une introduction rapide aux variétés différentielles mais également aux étudiants de
master qui n’ont jamais fait de géométrie et qui n’ont que des notions de topologie
de niveau licence.
J’aimerais remercier L. Jeanjean pour sa lecture détaillée de ces notes ainsi que ses
commentaires qui m’ont permis de les améliorer.
1
• In the first version of these notes I had assumed “paracompactness” in order
to treat in a unified way infinite dimensional and finite-dimensional manifolds.
However, upon further reflexion it appealed to me that this was not the ideal
way to develop the basic theory as whilst it is a desired feature of the topol-
ogy of manifolds, paracompactness does not behave well with respect to the
usual set-theoretic operations. For instance, subsets of paracompact spaces, or
products of paracompact spaces may not be paracompact. Instead, one should
work with sufficient conditions for paracompactness that are better behaved.
For finite-dimensional manifolds, it is custom to assume that the manifold is
second-countable; and we will assume this. This means that we suppose that
the topology has a countable basis. Since finite dimensional manifolds are
automatically locally compact and [locally compact + second-countable] ⇒
[paracompact]. Subspaces of second-countable spaces are second-countable;
metric spaces are second-countable if and only if they are separable. In the
infinite dimensional case, a sufficient condition would be to suppose that the
topology is regular and second-countable.
Rn+1 \ {0} π
RP n
f
f˜
X
2
The first step is to show that the quotient topology is Hausdorff and second-
countable; neither are guaranteed for general quotients of topological spaces hav-
ing these properties. However, our quotient comes from a continuous action of a
topological group; the following are general arguments we record here for future
reference:
Proposition I.1
Proof. Let U be open in X . In order to show that π(U ) is open in X/G one must
show that π −1 (π(U )) is open in X . One has:
[
π −1 (π(U )) = {x ∈ X, ∃g ∈ G, g · x ∈ U } = gU.
g∈G
Proof. Let (Vn )n∈N be a countable basis for the topology of X , the claim is that
(π(Vn ))n∈N is a basis for the topology of X/G.
Let [x] = π(x) ∈ X/G and U an open neighbourhood of [x]. By definition, π −1 (U )
is an open neighbourhood of x ∈ X , therefore one can find n0 ∈ N such that
x ∈ Vn0 ⊂ π −1 (U ), it follows that: [x] = π(x) ∈ π(Vn0 ) ⊂ U .
As an immediate consequence:
RPn is second-countable.
We will now give a general argument that guarantees separation; it is slightly more
involved and requires an additional condition on the group action.
3
Definition I.2
A continuous action of a topological group G on a locally compact topological
space X is said to be proper, if for any compact subset K ⊂ X the set:
{g ∈ G, gK ∩ K 6= ∅}
is compact.
Proposition I.2
Proof. We shall divide the proof into two steps; first make the following observation,
since the canonical projection π : X 7→ X/G is an open map, X/G is Hausdorff it
is sufficient to show that R = {(x, g · x), x ∈ X, g ∈ G} is closed in X × X .
To see this, assume that R is closed and suppose that π(x) 6= π(y) for some
x, y ∈ X ; then (x, y) ∈ (X × X) \ R. This set is open, therefore one can find
neighbourhoods Vx and Vy of x and y respectively such that Vx × Vy ⊂ (X × X) \ R,
but then: π(Vx ) and π(Vy ) are open neighbourhoods of π(x) and π(y) such that
π(Vx ) ∩ π(Vy ) = ∅. Indeed: if π(z) ∈ π(Vx ) ∩ π(Vy ) 6= 0, then ∃g1 , g2 ∈ G such that
(g1 z, g2 z) ∈ Vx × Vy but (g1 z, g2 z) = (g1 z, g2 g1−1 (g1 z)) ∈ R and R ∩ (Vx × Vy ) = ∅.
We shall now prove that R is closed; let (x, y) ∈ (X × X) \ R. Using that X
is locally compact, choose relatively compact neighbourhoods Vx and Vy of x and y
respectively; then Vx ×Vy is a relatively compact neighbourhood of (x, y). Introduce
the map:
G × X −→ X × X,
φ̃ :
(g, x) 7−→ (x, g · x).
Note that φ̃(G × X) = R. Moreover, the image under φ̃ of φ̃−1 (Vx × Vy ) is the set
of points in Vx × Vy that belong to R and that we want to avoid. If we can show that
φ̃−1 (Vx × Vy ) is “small”, i.e. compact, then the image will be compact too and easy
to avoid.
Set K1 = Vx × Vy and K2 = π1 (K1 ) ∪ π2 (K1 ) where π1 , π2 are respectively pro-
jections onto the first and second components of X × X . K2 is a compact subset of
X and so, since the action is proper, G0 = {g ∈ G, gK2 ∩ K2 6= ∅} is compact in G.
4
Now:
So, φ̃−1 (K1 ) is a closed subset of the compact G0 × K2 , hence compact. φ̃ being
continuous it follows that φ̃(φ̃−1 (K1 )) is a compact subset of Vx × Vy . Now Vx × Vy
and (X × X) \ φ̃(φ̃−1 (K1 )) are open neighbourhoods of (x, y) so (Vx × Vy ) ∩ (X ×
X) \ φ̃(φ̃−1 (K1 )) is an open neighbourhood of (x, y) contained in (X × X) \ R. This
proves that R is closed and consequently that X/G is Hausdorff.
The following sequential characterisation of properness is often useful.
Proposition I.3
RPn is Hausdorff.
5
(R∗ )N , (xn )n∈N ∈ (Rn+1 \ {0})N are sequences such that λn xn → y ∈ Rn+1 \ {0}
and xn → x ∈ Rn+1 \ {0}. Convergent sequences are bounded so there is M > 0
such that for all n ∈ N, |λn xn | ≤ M . Futhermore, for large enough n, ||xn || > ||x||
2
,
therefore, for large enough n, |λn | ≤ ||x|| . (λn )n∈N is therefore a bounded sequence
2M
Ui −→ Rn
φi :
(x1 , . . . , xn+1 ) 7−→ ( xi , . . . , xi , xi , . . . , xn+1
x1 xi−1 xi+1
xi
).
The result only depends on the class of π(x1 , . . . , xn+1 ), so that the map factors to a
continuous map φ̃i : π(Ui ) → Rn ; this is easily seen to be a homeomorphism as the
inverse map can be given explicitly:
Rn −→ π(Ui )
φ̃−1
i :
(y1 , . . . , yn ) 7−→ π((y1 , . . . , yi−1 , 1, yi , . . . , yn )).
A = {(π(Ui ), φ̃i ), i ∈ J1, n+1K} is then an atlas for RPn . Hence, RPn is a topological
manifold of dimension n.
I.B. C k manifolds
Definition I.3
In general, a full atlas will not be known explicitly and given a new chart (φ, U )
there is no way of knowing if it is in the atlas or not. We should therefore work with
atlases that contain all “reasonable” charts in the following sense:
2
More direct proofs certainly exists, but do not show the general principle.
6
Definition I.4
Definition I.5
Proposition I.4
Definition I.6
A C k manifold is a manifold equipped with a C k -atlas. A maximal C k -atlas is
also said to be a C k -differentiable structure on M . A C ∞ -manifold is said to
be smooth.
It should be understood from the discussion in this section that to define a dif-
ferentiable structure on a manifold M , we only need to specify an atlas, and then
consider the unique maximal atlas it is contained in. For instance:
Example I.3. A = {(E, idE )}, E a Banach space, is a smooth atlas on E and the
(canonical) smooth structure on E is that of the maximal atlas  containing A.
Example I.4. On RPn , consider the atlas A = {(Ui , φ̃i ), i ∈ J1, n+1K}. The transition
map: φ̃j ◦ φ̃−1
i is given on φ̃i (Ui ∩ Uj ) by:
−1 y1 yi−1 1 yi+1 yj−1 yj+1 yn
φ̃j ◦ φ̃i (y1 , . . . , yn ) = ,..., , , ,..., , ,..., .
yj yj yj yj yj yj yj
This is smooth, so A is a smooth atlas for RPn and the maximal atlas  makes RPn
a smooth manifold.
7
I.C. Submanifolds
Already several pages of theory and only one non-trivial example... to remedy this
let us define submanifolds and prove a classical theorem that produces many exam-
ples. The notion of submanifold is trickier than that of a subspace in a topology and
there are, in fact, several non-equivalent definitions (that allow for different phe-
nomena); we refer to [Sha97, Chapter 1] for an interesting discussion on this topic.
Our definition corresponds to “regular” submanifolds.
Definition I.7
A smooth submanifold N has a natural atlas given by the submanifold charts. Let
q ∈ N and (U, ψ) where ψ : U → V1 × V2 is a homeomorphism, then restricting to
U ∩N and projecting onto the first factor we get a homeomorphism: ψ̃ : U ∩N → V1 .
The transition functions are easily seen to be smooth so these charts yield a smooth
structure on N that is compatible with the subspace topology.
In this definition, V1 is an open subset of a Banach space E1 and one may wonder
if E1 could change from point to point. (This was excluded in Definition I.1, but
could in principle be allowed). In fact, since the transition maps between two charts
are smooth3 diffeomorphisms, we can see that all the possible E1 are topolinear
isomorphic (i.e. there is a continuous vector space isomorphism between them) at
points in the same connected component. So we get a manifold structure according
to our definition on each connected component of the submanifold.
Remark I.1. In the finite dimensional case, this is true on topological manifolds due
to the Brouwer invariance of domain theorem, which implies the topological invariance
of dimension; this second statement is harder to prove than one might think naively
! There are nice proofs of this in homology theory and one can find a clear discussion
in [Nab10]; although it is discussed in most standard books in algebraic topology.
Example I.5. Open subsets of manifolds are submanifolds.
We prove now a (sub-optimal) version of:
3
C 1 is enough.
8
Theorem I.1: Implicit function theorem
Proof. Let x0 ∈ f −1 ({0}), and set F = ker f ′ (x0 ). E/F ∼ = Rm so has a basis
(e1 , . . . , em ). For each i ∈ J1, mK, choose ẽi in E such that π(ẽi ) = ei where
π : E → E/F is the canonical projection. Set G = span(e1 , . . . , em ) and note
that we have the topological direct sum decomposition E = F ⊕ G; in particular
the projections are continuous.
Indeed, F and G are closed subspaces of E which automatically implies that the
projections pF and pG are continuous. This follows from the closed graph theorem:
suppose that (xn )n∈N converges to x and (pF (xn ))n∈N converges to y , since F is closed
it follows that y ∈ F, similarly, since G is closed and (xn − pF (xn )) is a sequence in
G that converges to x − y : x − y ∈ G. But, x = x − y + y and this decomposition
is unique, thus y = pF (x) and the graph is closed. Therefore, pF and pG = IdE − pF
are continuous.
Define now f˜ : U → F × G, by: f˜(x) = (pF (x), f (x)). We claim that f˜′ (x0 ) ∈
GL(E, F × G). In virtue of the open mapping theorem, it is sufficient to show that
f˜′ (x0 ) is bijective.
• Let (f, g) ∈ F×G, since f ′ (x0 ) is surjective one can find h0 such that f ′ (x0 )h0 =
g . Set h = h0 − πF (h0 ) + f , then f˜′ (x0 )h = (f, g). This proves surjectivity.
We can now apply the inverse function theorem to f˜ which provides neighbour-
hoods V and V1 × V2 ⊂ F × G of x0 and (πF (x0 ), 0) respectively, such that f˜|V :
V → V1 ×V2 is a diffeomorphism. This is the desired submanifold chart near x0 .
Remark I.2. The theorem extends to maps with values in Banach spaces, but we
need to strengthen the hypothesis on the derivative of points in x ∈ f −1 ({0}). The
correct hypothesis is that: f ′ (x) is surjective and the short exact sequence
f ′ (x)
0 −→ ker f ′ (x) −→ E −→ F −→ 0
i
9
Example I.6. S n = {x ∈ Rn+1 , ||x||22 = 1}.
Example I.7. SLn (R) = {A ∈ Mn (R), det A = 1}.
Example I.8. On (R) = {A ∈ Mn (R), At A = In }.
Let us conclude this section by quoting without proof the following theorem:
Remark I.3. 2n is optimal, but on specific examples one can often do better.
1
p(x) = (x1 , . . . , xn , 0) , x = (x1 , . . . , xn+1 ) ∈ UN .
1 − xn+1
Setting S = (0, . . . , 0, −1) and US = S n \ {S} one can define a chart φS in a similar
fashion. Then:
A = {(UN , φN ), (US , φS )},
is an atlas on S n . Let us prove that it is smooth by determining: φS ◦φ−1 N on R \{0},
n
one finds:
1
(φS ◦ φN )−1 (y1 , . . . , yn ) = (y1 , . . . , yn ) .
||y||22
This is smooth on Rn \ {0} so  is a smooth structure on S n .
10
II. The tangent bundle
For simplicity, from now on, all manifolds will be smooth.
Definition II.1
Let M and N be smooth manifolds and f : M → N , we will say that f is
smooth if for any p ∈ M one can find charts (U, φ), (V, ψ), p ∈ U , f (U ) ⊂ V
such that: ψ ◦ f ◦ φ−1 is C ∞ in the usual sense.
This definition does not depend on the charts (U, φ),(V, ψ). Indeed, suppose, for
instance, that: (Ũ , φ̃) is another chart such that p ∈ U , then (on an appropriate
intersection of the domains):
ψ ◦ f ◦ φ̃−1 = (ψ ◦ f ◦ φ−1 ) ◦ (φ ◦ φ̃−1 ),
transition functions are smooth by definition of the smooth atlas, so we can see that
ψ ◦ f ◦ φ̃−1 is C ∞ .
Example II.1. Let N ⊂ M be a submanifold of M , then the inclusion map i : N → M
is smooth. To see this let q ∈ N and (V, ψ), ψ : V → V1 × V2 a submanifold chart;
by definition of its atlas there is a corresponding chart (V1 , ψ̃), now:
ψ ◦ i ◦ ψ̃ −1 : V1 → V1 × V2 ,
is given by:
(ψ ◦ i ◦ ψ̃ −1 )(v) = (v, 0),
which is C ∞ .
Example II.2. The composition of smooth maps is smooth, in particular if f : M →
N is smooth and X is a submanifold of M , then the restriction of f to X is smooth.
Example II.3. Let M be a smooth E-manifold and (U, ψ) a chart, then ψ : U → E is
smooth; indeed: idE ◦ ψ ◦ ψ −1 is smooth.
We would now like to extend the notion of differential between functions on
Banach spaces to manifolds. One could suggest, as above, that we take local charts
and differentiate; but this would not be a chart invariant definition. Indeed let f :
M → N be a smooth map, choose, near p ∈ M , charts: (U1 , φ1 ), (U2 , φ2 ) such that
p ∈ U1 ∩ U2 and charts (V1 , ψ1 ), (V2 , ψ2 ) such that f (p) ∈ V1 ∩ V2 . Let us study how
the local representations of the function change:
(ψ2 ◦f ◦φ−1 ′ −1 ′ −1 −1 ′ −1 −1 ′
2 ) (x) = (ψ2 ◦ ψ1 ) (ψ1 ◦ f ◦ φ2 (x)) ◦(ψ1 ◦f ◦φ1 ) (φ1 ◦φ2 (x))◦(φ1 ◦ φ2 ) (x) .
| {z } | {z }
Q−1 P
11
We can see from this formula that the local chart derivative transforms under a
change of chart like a linear map when we change basis. Inspired by this analogy,
we introduce the following equivalence relation. Let p ∈ M be fixed and consider
triples (U, φ, v), (U, φ) is a chart and v ∈ E is a vector in the model space. We define
an equivalence relation on (U, φ, v) by:
Remark II.1. Let us make the statement of Definition II.2 more explicit. Fix p ∈ M
and let [(φ, U, v)]p denote the tangent vector of which (φ, U, v) is a representative.
We define a vector space structure on Tp M by:
where (φ, U ) is an arbitrary local chart such that p ∈ U . The definition makes
sense since if (ψ, W ) is another chart with p ∈ W , then: [(φ, U, v)]p = [ψ, V, (ψ ◦
φ−1 )′ (φ(p)) · v]p for any v ∈ E, linearity of the derivative guarantees that:
It may seem that the topology could depend on a choice of chart, but this is not the
case as by definition:
12
which is a continuous linear isomorphism E → E, hence the topology is independent
of the choice of chart. Note that, whilst one can use the bijection φ∗p to define a
norm on Tp M , this norm depends on the choice of chart, which is why we only get a
Banachisable topological vector space structure.
where (U, φ) is a chart such that p ∈ U and (V, ψ) is a chart such that f (p) ∈
V . In the diagram we denote (abusively) by φ∗p : Tp M → E1 the canonical
continuous linear isomorphism [(U, φ, v)] → v and similarly for ψ∗f (p) .
Remark II.2. When E is a Banach space, it is clear that for any p ∈ E, there is
a canonical identification Tp E = E in the global chart given by the identity; we
will systematically make this identification. This also justifies the notation φ∗p in
Definition II.3, as one can easily verify that this is indeed the tangent map of the
chart φ at p !
The tangent map behaves like the usual differential, in particular:
Proof. Introduce local charts (U, φ), (V, ψ), (W, η) such that: f (U ) ⊂ V , g(V ) ⊂ W ,
and fix p ∈ U . By definition:
13
Now:
(η ◦ g ◦ f ◦ φ−1 )′ (φ(p)) = (η ◦ g ◦ ψ −1 ◦ ψ ◦ f ◦ φ−1 )′ (φ(p)),
= (η ◦ g ◦ ψ −1 )′ (ψ(f (p))) ◦ (ψ ◦ f ◦ φ−1 )′ (φ(p)),
= (η ◦ g ◦ ψ −1 ◦ ψ ◦ f ◦ φ−1 )′ (φ(p)),
= (η ◦ g ◦ ψ −1 )′ (ψ(f (p))) ◦ ψ∗p ◦ (ψ∗p )−1 ◦ (ψ ◦ f ◦ φ−1 )′ (φ(p))
= η∗g(f (p)) g∗f (p) ◦ f∗p ◦ (φ∗p )−1 .
It follows that:
η∗g(f (p)) ◦ (g ◦ f )∗p = η∗g(f (p)) g∗f (p) ◦ f∗p .
The result follows.
Let p ∈ M , two smooth functions f, g on M are said to define the same germ
at p, if there is an open neighbourhood U of p, such that f |U = g|U . This
defines an equivalence class on smooth functions on M that we call the germ
of f at p, written fp . Let Cp∞ (M ) denote the set of all germs at p, pointwise
multiplication and addition gives Cp∞ (M ) the structure of an R-algebra.
14
In the above formula, f (p) and g(p) denote the value at p of any representative f
(resp. g ) of the germ fp (resp. gp ); by definition it does not depend on this choice.
Note that if fp is the germ of a constant map, then Dfp = 0. Indeed, suppose
fp is the germ of a constant map equal to λ, then for any gp , fp gp = λgp so that:
λDgp = D(λgp ) = D(fp gp ) = D(fp )g(p) + λDgp . Thus D(fp )g(p) = 0 for any
function g , this implies D(fp ) = 0.
The important result is:
Proposition II.2
Proof. Choose a chart (U, φ) near p and note that Dvp f = (f ◦ φ−1 )′ (φ(p))(φ∗p vp )
This formula shows that Dvp f only depends on the germ of f at p so factors to a
map on Cp∞ (M ). Using the product rule,
so that this map is a derivation. For the injectivity of the correspondence, suppose
that for some vp ∈ Tp M , Dvp f = 0 for any f . Let l ∈ E′ and consider f = l ◦ φ.
Then for any l ∈ E′ , l(φ∗p vp ) = 0 so φ∗p vp = 0 ⇒ vp = 0.
Now assume that M is of dimension n. We shall show that all derivations are of
the form Dvp for some vp ∈ Tp M . Let D be a derivation and (U, φ) be a local chart
near p. Let fp ∈ Cp∞ (M ) and f a representative of the germ, using a bump function,
it can be chosen such that suppf ⊂ U . For q in the support of f one can write:
Z 1
f (q) = f (p) + (f ◦ φ−1 )′ (φ(p) + t(φ(q) − φ(p)) · (φ(q) − φ(p))dt.
0
X
n Z 1
f (q) = f (p) + e∗i (φ(q) − φ(p)) (f ◦ φ−1 )′ (φ(p) + t(φ(q) − φ(p)) · ei dt.
i=1 0
15
Applying D in this formula we obtain, due to the Leibniz rule and the fact that it
annihilates germs of constants:
X
n X
n
Dfp = D((e∗i ◦ φ)p )(f ◦ φ−1 )′ (φ(p))(ei ) = D((e∗i ◦ φ)p )f∗p · φ∗p (ei )
i=1 i=1
= f∗p vp ,
Pn
where: vp = φ∗p ( i=1 D((e∗i ◦ φ)p )ei ) .
This justifies:
We can now revisit the definition of the tangent map, thinking of tangent vec-
tors as derivations:
∞
where fϕ(p) ∈ Cϕ(p) (N ) and (f ◦ φ)p denotes the germ at p of f ◦ φ, where f is
any representative of the germ.
16
Definition II.8
Let M be a finite dimensional manifold, p ∈ M , (U, x) a local chart we define:
∂
= x−1
∗x(p) ei .
∂xi p
a
Any tangent vector at p can be written : X = X i ∂
∂xi
. Xi ∈ R
p
a
See the Einstein’s summation convention
To conclude this section we shall check that Definition II.7 and Definition II.3,
coincide on finite dimensional manifolds. Let φ : M → N be a smooth map between
two finite dimensional manifolds dim M = n, dim N = m, choose p ∈ M , (x, U ) a
chart at p and (y, V ) a chart on N such that f (U ) ⊂ V .
First, let us make explicit II.8. In Rn the vector ei is interpreted as the partial
derivative operator ∂ei , so, using Definition II.7, x−1 ∗x(p) ei should be interpreted as
the derivation defined by:
∂
(f ) = ∂ei (f ◦ x−1 )(x(p)).
∂xi p
f is the germ of some smooth function at p and, the right hand side is the ith partial
derivative (in the usual sense) of the function f ◦x−1 , defined
on the open set x(U ) ⊂
R . Now, let us calculate φ∗p Xp , where Xp = X ∂xi again using Definition II.7,
n i ∂
p
X
n
∂
(φ∗p Xp )(f ) = Xp (f ◦ φ) = Xi (f ◦ φ)
i=1
∂xi p
X
n
= X i ∂ei (f ◦ φ ◦ x−1 )(x(p))
i=1
X
n
∂
= X i
(f ◦ φ)
i=1
∂xi p
Xn
= X i ∂ei (f ◦ y −1 ◦ (y ◦ φ ◦ x−1 ))(x(p))
i=1
X
n X
m
using the chain rule = X i ∂ei (y ◦ φ ◦ x−1 )j (x(p))∂ej (f ◦ y −1 )(y(φ(p)))
i=1 j=1
!
X
m X
n
∂
−1
= X ∂ei (y ◦ f ◦ x )j (x(p))
i
(f ).
j=1 i=1
∂yj ϕ(p)
17
Where we have introduced the notation (y ◦f ◦x−1 )j to denote the j -th component
in the canonical basis of y ◦ f ◦ x−1 . Now we compare with Definition II.3, recall
that one must have:
−1
(y ◦ φ ◦ x−1 )′ (x(p)) ◦ x∗p = y∗ϕ(p) ◦ φ∗p ⇔ φ∗p = y∗y(ϕ(p)) ◦ (y ◦ φ ◦ x−1 )′ (x(p)) ◦ x∗p
Moreover: !
X
n
∂ X
X= Xi = x−1
∗x(p) X i ei ,
i=1
∂xi p i=1
hence:
X
m X
n
−1 ′
(y ◦ φ ◦ x ) (x(p)) ◦ x∗p (X) = ej X i (y ◦ φ ◦ x−1 )′j (x(p)) · ei ,
j=1 i=1
X
m X
n
= ej X i ∂ei (y ◦ φ ◦ x−1 )j (x(p)).
j=1 i=1
−1
Since y∗y(ϕ(p)) ej = ∂
∂yj
, it follows that the two definitions are identical.
ϕ(p)
One can also use Definition II.7 to give a much more elegant (and coordinate free)
proof of the chain rule in finite dimensions.
Alternative proof of the chain rule in finite dimensions. Let φ : M → N , ψ : N → S , let
f be the germ of a smooth function at (ψ ◦ φ)(p) for fixed p ∈ M . Let Xp ∈ Tp M ,
thought of as a derivation at p, then:
18
the information contained in the tangent spaces at different points. It is the first
example that we will encounter of a fibre bundle. Furthermore, here the fibres will
also have a topological vector space structure, and hence it is an instance of a vector
bundle. We give the general definition:
The family B = {(Ui , τi )} is said to be a vector bundle atlas. In the exact same
way as for atlases on manifolds, a vector bundle atlas determines a unique
maximal vector bundle atlas that we call a vector bundle structure on π (or
E ). We say that E is the total space and M the base manifold.
a
Fancy way of saying continuous linear.
They contain in fact contain all the essential information of the bundle: given a
covering {Ui } and a family of transition maps {τji,p } satisfying the cocycle condition,
it is possible to reconstruct E and π .
Let us now describe how to construct T M , suppose that M`is an E-manifold.
Consider first the coproduct of sets (i.e. disjoint union) T M = p∈M Tp M . Specify
19
π : T M → M on each Tp M to be the constant map vp ∈ Tp M 7→ p.4 For every
chart (U, φ) consider a map: φ̂ : π −1 (U ) → φ(U ) × E defined by5 :
vp ∈ Tp M ,→ T M 7→ (φ(p), φ∗p vp ).
Now, equip T M with the coarsest topology6 that makes π and all the φ̂ continuous;
with this topology the φ̂ are homeomorphisms. The transition map between any
two charts is:
and hence is smooth. The charts {(π −1 (U ), φ̂), (U, φ) chart on M } therefore deter-
mine a smooth structure on T M . Define now: φ̃ : π −1 (U ) → U × E by φ(vp ) =
(p, φ∗p vp ), vp ∈ Tp M. and observe that these are smooth bundle charts; the transi-
tion maps are easily seen to be given by: p 7→ (ψ ◦ φ−1 )′ (p) and are smooth, which
shows that π is a vector bundle.
Remark II.5. Although T M is a local product, in general it is not globally diffeo-
morphic to a product of manifolds; when this is possible the manifold is said to be
parallelisable.
We quote a few examples that we can describe explicitly:
Example II.5. If U is an open subset of a Banach space E, T U = U × E.
Example II.6. Let S n ⊂ Rn+1 be the Euclidean sphere, then:
Proposition II.3
4
We apply here the universal property of coproducts.
5
As
` before it is sufficient to define the function on each Tp M to define a function on T M =
p∈M Tp M
6
Let (Yi )i∈I be a family of topological spaces and fi : X → Yi a family of maps. The initial
topology of (fi )i∈I is the finest topology on X such that all the maps are continuous. A basis for
this topology are sets of the form ∩j∈J fj−1 (Uj ) where J ⊂ I is finite and Uj ⊂ Yj is open.
20
II.D. Operations on vector bundles
Certain functors of subcategories of Banach spaces extend to functors of vector
bundles, to keep this discussion brief we refer the reader to [Lan95, Chapter III,
§4], we shall simply quote a few important examples.
Example II.7. Consider first the contravariant functor that sends a Banach space E
to its dual7 E∨ ; recall that the map between morphisms is given by: u 7→ t u. There
is a corresponding functor of vector bundles of base M , π : E → M , that yields a
new vector bundle E ∨ whose fibres are the dual spaces of those of E . Applied to
the tangent bundle, this gives us the cotangent bundle.
Example II.8. If E1 and E2 are two vector bundles with the same base M one can
form their direct sum8 , E1 ⊕ E2 which is a vector bundle with base M and whose
fibre at a point x ∈ M is E1x ⊕ E2x .
Example II.9. In finite dimensions, one can also form the tensor product of vector
bundles.
In a local chart (U, φ) this is a map from U to U × E of the form: p 7→ (p, XU (p)).
The map XU is called the local representative of the vector field, X is smooth if and
only for every p ∈ M , one can find a chart (U, φ), such that that XU is smooth. In
finite dimensions, this translates to the fact that: X = X i ∂x∂ i is smooth if and only if
the functions (X i ) are smooth. Note that under a change of chart (U, φ) → (V, ψ),
one has: XV = (ψ◦φ−1 )′ ◦φ·XU . Conversely, if for an atlas on M one is given a family
of smooth maps, indexed by the charts, that verify this compatibility condition then
one can construct a unique vector field locally described by the family.
If X is a vector field to avoid too many parentheses it is customary to write
X(p) = Xp . We denote by Γ(T M ) the set of smooth vector fields on M ; it is a
C ∞ (M )-module; there is an injective map into the set of derivations of C ∞ (M ). A
remarkable fact is that it is also a Lie algebra.
7 |l(x)|
Banach space of continuous linear forms on E, equipped with the norm klk = supx̸=0 ||x||
8
also known as the Whitney sum
21
Theorem II.1
Let X, Y be two vector fields on M , ∂X , ∂Y the corresponding derivations,
then there is a unique vector field [X, Y ] such that: ∂X ∂Y − ∂Y ∂X = ∂[X,Y ] .
In a local chart (U, φ), the local representative is:
[X, Y ]U = YU ∗ (φ−1 −1
∗ · XU ) − XU ∗ (φ∗ · YU ),
= (YU ◦ φ−1 )′ ◦ φ · XU − (XU ◦ φ−1 )′ ◦ φ · YU .
Where in the final step we use the symmetry of the second derivative. We have
therefore shown that, on U , ∂X ∂Y − ∂Y ∂X is a derivation associated with a vector
field over U locally represented by:
[X, Y ]U = YU ∗ (φ−1 −1
∗ · XU ) − XU ∗ (φ∗ · YU ).
It remains to ensure ourselves of the fact that this can be glued together to from
a vector field on M . For this, is sufficient to check that it transforms correctly
under a change of chart. Recall that if (V, ψ) is another chart then on U ∩ V , XV =
(ψ ◦ φ−1 )′ ◦ φ · XU , to simplify notation let us write: h = (ψ ◦ φ−1 )′ ◦ φ now:
Now:
ψ∗−1 h · XU = ψ∗−1 (ψ∗ ◦ φ−1 −1
∗ ) · XU = φ∗ X U ,
and:
(h · XU )∗ (φ−1 −1 ′ −1
∗ YU ) = (ψ ◦ φ ) (XU , YU ) + h · (XU )∗ (φ∗ YU ).
[X, Y ]V = h · [X, Y ]U .
Therefore, it is clear that this patches together to define a smooth vector field on
M , that we denote by [X, Y ]U .
22
Remark II.6. If M is of finite dimension n and (U, x) is a local chart, the local formula
becomes on U :
Xn X n
∂Y i j ∂X i ∂
[X, Y ] = X − ·Y j
,
i=1 j=1
∂xj ∂xj ∂xi
∂
[X, Y ] = .
∂y
Proposition II.4
• It is bilinear, antisymmetric.
Summation of the three expressions shows the result, using once more symmetry of
the second derivative.
It also worth noting, for some calculations, that we have the following Leibniz
rule for [X, Y ]:
[X, f Y ] = (∂X f )Y + f [X, Y ].
Proof. Exercise.
23
II.F. Flow and bracket
Let γ : I → M be a curve on M defined on an open interval I . Note that T I = I ×R
and we have a canonical section j : t 7→ (t, 1). The canonical lift of γ is the map
γ̇ = γ∗ ◦ j , it is a curve on T M that satisfies π ◦ γ̇ = γ , where π : T M → M is the
projection of T M onto M .
Remark II.7. For each t ∈ I , γ̇(t) has the usual interpretation as the tangent (or
velocity) vector to the curve at the point γ(t). It is sometimes written:
d
γ̇(t0 ) = γ(t) .
dt t=t0
α∗ : T (−ε, ε) ∼
= (−ε, ε) × R → T φ(U ) ∼
= φ(U ) × E,
with: α∗ (t, λ) = (α(t), α′ (t)·λ), here: α′ (t) ∈ L(R, E) is the usual derivative, indeed,
it is naturally identified with α∗t : Tt (−ε, ε) ∼ = R → Tα(t) E ∼ = E. Now recall that
we have the canonical section j : (−ε, ε) → T (−ε, ε) = (−ε, ε) × R given by:
j(t) = (t, 1). Hence: α̇(t) = (α∗ ◦ j)(t) = (α(t), α′ (t) · 1). By the chain rule:
γ̇(0) = φ−1 −1
∗α(0) α̇(0) = φ∗ϕ(p) (φ∗p · vp ).
To understand the last inequality it is important to note that although in all rigour:
φ−1 ∼
∗ϕ(p) : Tϕ(p) φ(U ) = {φ(p)} × E → Tp M we systematically identify {φ(p)} × E with
E and write:
φ−1 −1
∗ϕ(p) (φ(p), v) ≡ φ∗ϕ(p) v.
24
It can sometimes be useful to note that if f : M → N and vp ∈ Tp M then for any
curve γ : I → M such that γ(0) = p, γ̇(0) = vp we have:
˙
f∗p vp = (f ◦ γ)(0), (II.1)
With a bit of elbow grease one can show the following fundamental theorem of
flows:
25
Theorem II.2: Fundamental theorem of flows
Let X be a smooth vector field on a smooth manifold M , then there is a unique
maximal flow domain D such that:
1. (t, p) ∈ D 7→ φX
t (p) is a local flow,
We refer to [Lee03] for the proof. This gives another formula for the Lie bracket
that is sometimes useful:
d X
[X, Y ](p) = φ−t ∗ϕX (p) · Y (φX (p)) .
dt t
t
t=0
This shows that [X, Y ] measures the infinitesimal change in Y under the effect of
the flow of X .
Proof. Exercise...
26
It is a smooth map; this follows from the fact that if E × F is equipped with its norm
||(x, y)||E×F = ||x|| + ||y||, then f : U ⊂ G → E × F is differentiable if and only if
π1 ◦ f and π2 ◦ f are differentiable. Equipping M × N with the maximal atlas Cˆ, this
property generalises to manifolds.
Proposition II.6
Let (p, q) ∈ M × N ; then, as one might expect, we have the toplinear isomor-
phism:
T(p,q) (M × N ) ∼
= Tp M × Tq N.
This follows directly from Definition II.2.
Let (E, π, M ), (E ′ , π ′ , M ) two vector bundles with same base M , a vector bun-
dle morphism is a smooth map: f : E → E ′ such that
• π′ ◦ f = π,
• for each p ∈ M , the induced map between fibres: fp : π −1 (p) → π ′−1 (p)
is continuous linear
Remark II.8. • The final condition is superfluous for finite dimensional mani-
folds.
27
III. Geometry on manifolds, connections
From now on, for simplicity, all manifolds will be assumed finite dimensional.
28
Proposition III.1
φ∗ [X, Y ] = [φ∗ X, φ∗ Y ].
Proof. This is not the most elegant proof, but it is valid in infinite dimensions. Let
us work in a local charts, suppose that (U, x) is a chart at p, (V, y) a chart at φ(p)
such that φ(U ) ⊂ V . Let XU and X̃V be the local representations of the fields near
in the chart then they are φ-related means that for any p ∈ U :
Now:
[X̃, Ỹ ]V (φ(p)) = (ỸV ◦ y −1 )′ (y(φ(p)) · X̃U (φ(p)) − (X̃V ◦ y −1 )′ (y(φ(p)) · ỸU (φ(p))
Furthermore:
((y ◦ φ ◦ x−1 )′ (x(p)) · YU (p))′ · XU (p) =(y ◦ φ ◦ x−1 )′′ (x(p)) · (XU (p), YU (p))
+ (y ◦ φ ◦ x−1 )(x(p)) · YU′ (p) · XU (p).
Treating the second term in exactly the same fashion and using the symmetry of the
second derivative, we find that:
29
Since left-invariant vector fields are uniquely determined by their value at the
identity, we can identify g and Te G as vector spaces, and transfer the Lie algebra
structure to Te G, i.e. by setting:
[Xe , Ye ] = [X, Y ]e ,
where X (resp. Y ) is the unique left-invariant vector field such that X(e) = Xe (resp.
Y (e) = Ye ).
Proposition III.2
• The integral curves of a left-invariant vector field are complete, i.e. de-
fined on all of R.
exp(X) = γ(1).
30
Example III.2. Let G = GLn (R), the tangent space at In is identified with Mn (R).
The exponential map of X ∈ gln (R) = Mn (R) is given by the usual power series:
P
k∈N k! X . This enables us to determine the Lie bracket on Mn (R). Using Propo-
1 k
31
It is standard that we have a bijection G/H ∼ = An ; in fact G/H equipped with the
quotient topology is a topological manifold and has a unique smooth structure such
that the canonical projection π : G → G/H is a smooth submersion; it is then the
case that G/H is diffeomorphic to An . We omit this development and instead make
some further observations:
x In x
1. π admits a smooth global section ∈ A 7→
n
. The interpretation
1 0 1
of σ is as follows: at each point x ∈ Rn we choose a frame anchored at x and
composed of the n vectors (e1 , . . . , en ) of the canonical basis of Rn .
3. H acts on the right on G and we have φ(gh) = φ(g)h, where on the rhs10 this
is the canonical right action on the product An × H i.e. (x, h1 )h2 = (x, h1 h2 )
This is our first example of a principle H -bundle that we will define below.
With respect to this structure, G is interpreted as the “frame bundle” of An :
at each point we glue the set of all possible linear frames anchored at x; the
right-action can be thought of as a (global) change of frame.
To describe the properties of the Maurer-Cartan form we need to recall some alge-
bra.
9
σ(x) = ϕ−1 (x, e)
10
right-hand side
32
III.D. Differential forms and the exterior algebra
1 X
(α∧β)(v1 , . . . , vk ) = ε(σ)α(vσ(1) , . . . , vσ(k) )β(vσ(k+1) , . . . , vσ(k+l) )
k!l! σ∈S
k+l
Example III.4. Let u and v be one forms, then: (u ∧ v)(x, y) = u(x)v(y) − v(x)u(y).
Let V have finite dimension n ∈ N∗ , and B = (e1 , . . . , en ) a basis of V . Write the
dual basis (e∗1 , . . . , e∗n ), then by induction one can show that:
33
Proposition III.3
L Vk ∨
The multiplication ∧ is bilinear and associative, making R ⊕ ∞ k=1 V an
associative algebra with unit called the exterior algebra. If α and β are respec-
tively a k -form and l-form then:
(α ∧ β) = (−1)kl β ∧ α.
V
If dim V = n then m V ∨ = 0, if m > n and if (e1 , . . . , en ) is a basis for V ,
then:
{ei∗1 ∧ · · · ∧ e∗ik , 1 ≤ i1 < i2 < · · · < ik ≤ n}
V
is a basis of k V ∨
1. If f is a smooth function df = f ′ .
3. d2 = 0.
34
Example III.6. Let U be an open set of Rn , (e1 , . . . , en ) the canonical basis. Define,
xi : U → R which maps p to its ith coordinate of p in the canonical basis. Then for
Xn
∂f
any p ∈ U , dxip = e∗i . If f is a smooth function then we can write df = dxi .
i=1
∂x i
∂ ∂ ∂ ∂ ∂ ∂
= cos θ + sin θ , = −r sin θ + r cos θ
∂r ∂x ∂y ∂θ ∂x ∂y
The dual basis is
xdx + ydy xdy − ydx
dr = p , dθ =
x2 + y 2 x2 + y 2
V
To extend this to manifolds, one can construct k Tp∨ M at each point p, in the
V ` V
same way as for a vector space. Then, one can equip k T ∨ M = p∈M k Tp∨ M
V
with the structure of a vector bundle where the fibres are given by k Tp∨ M . This is
an example of the more general procedure touched upon in Section II.D.
V
The gory details are as follows. Define π : k T ∨ M → M by αp ∈ Tp∨ M 7→ p.
V
Choose a chart, (x, U ) and define a map x̂ : π −1 (U ) → U × k (Rn )∨ by:
^k
x̂(αp ) = (p, αp (x−1 −1 −1
∗x(p) , x∗x(p) , · · · , x∗x(p) )), αp ∈ Tp∨ M.
These are our natural candidates for our vector bundle charts. Composing with the
chart we get a bijective map:
^k
x̃ : π −1 (U ) → x(U ) × (Rn )∨
that is a natural candidate for a coordinate chart. As with the tangent bundle we
V
equip k T ∨ M with the coarsest topology making π and every x̃ (where x runs
through the charts on M ) continuous. The x̃ are easily seen to be homeomorphisms
with this choice of topology. If (y, V ) is another chart then:
35
Since M is a smooth manifold
V these transition charts are also smooth and so define
a smooth structure on k T ∨ M . This is precisely the smooth structure that makes
the bundle charts x̂ smooth diffeomorphisms. It is straightforward to check that
Vk ∨
T M is a vector bundle.
Smooth sections of this bundle are called differential k -forms on M , they form a
C ∞ (M ) module denoted by Ωk (M ).
Example III.10. Let xi = e∗i ◦ x where (x, U ) is a local chart. Define a differential
form on U , dxi , by:
dxip = xi∗p = e∗i ◦ x∗p .
V
Then dxi is a smooth section of 1 T ∨ U ∼ = T ∨ U , indeed, in the chart (x, U ) this
∗
is locally represented by ei . For each p ∈ U , (dx1p , . . . , dxnp ) is by definition the dual
basis to the coordinate basis ( ∂x∂ 1 , · · · , ∂x∂n ).
p p
We can inductively define differential k -forms on U , dxi1 ∧ · · · ∧ dxik , 1 ≤ i1 <
V
· · · < ik ≤ n that at each point p give a basis of k Tp∨ M . An arbitrary differential
k -form can be represented locally:
X
α= αi1 ...ik dxi1 ∧ · · · ∧ dxik ,
0≤i1 <···<ik ≤n
where αi1 ...ik are functions on U , the differential form is smooth if and only if they
are smooth for every chart (x, U ).
This is in a nutshell how we work with forms on manifolds in a very basic fashion
and in a way that is designed to ressemble the case of an open set of Rn .
Whilst this is satisfying for computations, to develop the theory it is sometimes
nice to avoid using charts by adopting another point of view. Let α ∈ Ωk (M ), if
X1 , . . . , Xk are vector fields on M then one can define a C ∞ (M )-linear k -alternating
form A : Γ(T M ) × · · · × Γ(T M ) → C ∞ (M ) by:
On finite dimensional manifolds, it turns out that the two points of view are equiva-
lent. The key to this correspondence is the following lemma:
36
Lemme III.1
Let T : Γ(T M ) × · · · × Γ(T M ) → C ∞ (M ) be a k -multilinear map on the
C ∞ (M ) module Γ(T M ), then for each p ∈ M , and any X1 , . . . , Xk ∈ Γ(T M ),
the value of T (X1 , . . . , Xk )(p) only depends on the values of the vector fields
X1 , . . . , Xk at the point p.
Consequently, for each p, T induces a k -multilinear map: Tp : Tp M ×
· · · Tp M → R. Furthermore, if T is alternating p 7→ Tp is a smooth section of
Vk ∨
T M.
Proof. It is sufficient to treat the case k = 1 and then use induction. Fix p ∈ M and
choose a chart (x, U ) near p, let b : M → R be a smooth bump function with support
in U . Let X ∈ Γ(T M ) and decompose Xp on the coordinate basis as Xp = v i ∂
∂xi
.
p
Set Y = v i b ∂x∂ i . C ∞ (M )-linearity gives:
∂
α(bX − Y ) = (X i − v i )α(b )
∂xi
hence: α(bX − Y )(p) = 0, but:
∂
α(X)(p) = b(p)α(X)(p) = α(bX)(p) = α(Y )(p) = v i α(b( ))(p).
∂xi
Thus showing that α(X)(p) only depends on Xp .
This correspondence is useful for defining tensor objects on manifolds without
referring to a specific choice of chart or working in local coordinates. (Although to
actually calculate things we do have to choose coordinates at some point). For in-
stance, one can extend the exterior derivative to manifolds and we have the invariant
formula:
Let α be a k -form on M , then its exterior derivative can be defined by the same
conditions as in Proposition III.4. We have the following invariant formula:
X
k
dα(X1 , . . . , Xk+1 ) = (−1)i+1 Xi (α(X1 , . . . , X̃i , . . . , Xk+1 ))
i=1
X
+ (−1)i+j ω([Xi , Xj ], X1 , . . . , X̃i , . . . , X̃j , . . . , Xk+1 )
i<j
(III.1)
where ˜ denotes omission. Recall that vector fields are to be thought of as
derivations.
37
Example III.11. The most important case is when α is a one form then the formula
reads:
dα(X, Y ) = X(α(Y )) − Y (α(X)) − α([X, Y ]).
1 X
[α, β](v1 , . . . , vk+l ) = ε(σ)[α(vσ(1) , . . . , vσ(k) ), β(vσ(k+1) , . . . , vσ(k+l) )].
k!l! σ∈S
k+l
Example III.12. Let α be a g-valued one form, then [α, α](v1 , v2 ) = 2[α(v1 ), α(v2 )]
The exterior derivative can be extended to g-valued forms by choosing an arbi-
trary basis E i of g and declaring that d acts on components (which are usual k -forms)
like the usual exterior derivative; one should check that this does not depend on the
choice of basis. The invariant formula in Proposition III.5 then extends without
any modification.
We can now summarise the main properties of the Maurer-Cartan form. Recall
that for every p ∈ G, Xp ∈ Tp G,
ωp (Xp ) = Lp−1 ∗p Xp ∈ Te G ∼
= g.
Now H is a closed subgroup of G and H acts on G from the right, let us denote by
Rh : G → G the map p 7→ ph. Calculating the pullback of ω under Rh , we have:
where he have defined Adh : p 7→ h−1 ph. Its tangent map at the identity element,
Adh∗e – which we will abusively also denote Adh in the sequel – defines a represen-
tation of H with representing space g, called the adjoint representation. We also say
that g is a H -module.
38
The second property we have is that for any X ∈ h ⊂ g, one can consider a vector
in Tp G for any other p defined by:
d
Xp∗ = p exp(tA) ,
dt t=0
then:
ωp (Xp∗ ) = X. (III.3)
Note that the vector Xp∗ is in ker π∗p where π : G 7→ G/H is the canonical projec-
tion. Tangent vectors in ker π∗p are known as vertical vectors; intuitively they point
in the direction of the fibres of G above G/H . In fact all vertical vectors can be
obtained as fundamental vectors. The vector field p 7→ Xp∗ is called the fundamental
vector field associated to X . Summarising this:
Proposition III.6
The Maurer-Cartan form satisfies one further equation; later we will interpret this
as the local expression of the flatness of affine space.
1
dω + [ω, ω] = 0
2
Proof. Let us work with the invariant formulae. First ω is a one form so for any
vector fields X, Y :
and
1
[ω, ω] = [ω(X), ω(Y )].
2
39
Now, recall from Lemma III.1 that the value at any given point g ∈ G of the function
dω(X, Y ) + 21 [ω, ω] only depends on the values of X and Y at g . Therefore it is
sufficient to study the case where X and Y are left-invariant vector fields, In this case:
1
dω(X, Y ) = −ω([X, Y ]) = −[Xe , Ye ] = −[ω(X), ω(Y )] = − [ω, ω].
2
Remark III.1. Recall that a chart of a 2-dimensional manifold like affine space is
defined as a map from an open set of a manifold onto an open subset of R2 . Equiv-
alently this can be described by two maps: r : U → R and θ : U → R.
Let us now consider the pullback of the Maurer-Cartan form of G onto U . First,
let us determine what the Maurer-Cartan form is in matrix terms. By definition:
ωg = Lg−1 ∗g i.e. it is the derivative at g of the restriction to G of the linear map:
q 7→ g −1 q . It can be written:
ωg = g −1 dg
where dg is to be understood as the derivative at p of the injection g ,→ gln+1 (R).
The pullback along σ is given by:
σ ∗ ω p = σ −1 (p)σ∗p p∈U
40
Since:
cos θ(p) sin θ(p) −r(p)
σ −1 (p) = − sin θ(p) cos θ(p) 0 ,
0 0 1
− sin θ(p)dθ − cos θ(p)dθ cos θ(p)dr − r(p) sin θ(p)dθ
σ∗p = cos θ(p)dθ − sin θ(p)dθ sin θ(p)dr + r(p) cos θdθ
0 0 0
We have:
0 −dθ dr
σ ∗ ωp = dθ 0 rdθ .
0 0 0
Analysing this result, notice that the last column is the dual basis of (er , eθ ); this can
be interpreted as the infinitesimal displacement vector, expressed in the basis er , eθ
as we move from a point p to p + δp. On the other hand, the matrix:
0 −dθ
,
dθ 0
is a matrix describing the infinitesimal change in the basis vectors. Indeed, differ-
entiate in the classical sense the equations defining er and eθ we find:
Hence, to first order, it is the change of basis matrix from (er , eθ ) to (er+δr , eθ+δθ ).
In other words the Maurer-Cartan form measures how the moving basis (section)
changes as we move from point to point.
41
Definition III.8: Principal bundles
1. ∀r ∈ P, h ∈ H, π(r · h) = π(r),
ϕ
π −1 (U ) U ×H
π
πU
U
Then π is a principal H -bundle with total space P , base M and structure group
H.
Remark III.2. • The right-action is free, and the orbits are the fibres:
π −1 ({π(p)}) = {p · h, h ∈ H}
• The fibres are diffeomorphic to H but do not have a natural group structure.
Every smooth n-dimensional manifold M has a natural smooth GLn (R)-principal
bundle, L(T M ) known as the frame bundle. The fibre at p ∈ M of this bundle will
be the set of linear frames of Tp M , described as follows
L(T M )p = GL(Rn , Tp M ).
`
There is a natural right-action of GL(Rn ) ∼= GLn (R) on L(T M ) = p∈M L(T M )p
given by u · g = u ◦ g , the natural projection defined by: π : up ∈ L(R, Tp M ) ,→
L(T M ) 7→ p, certainly satisfies π(ug) = π(u); i.e. preserves the fibres.
Let (x, U ) be some chart on M , then for each p ∈ U we have a natural map:
L(Rn , Tp M ) → {p} × GL(Rn ) ∼ = GLn (R) given by: up 7→ (p, x∗p ◦ up ), hence we
−1
get a bijective map φ : π (U ) → U × GLn (R) such that the following diagram
commutes:
ϕ
π −1 (U ) U × GL(Rn )
π
πU
U
42
We just need to define a topology and a smooth structure to make all of this smooth;
we proceed as for the tangent bundle. Compose φ with the chart of the manifold
and define:
φ̂ : π −1 (U ) → x(U ) × GL(Rn ),
these will be our local charts. Endow L(T M ) with the coarsest topology that makes
π and all φ̂ for any chart (x, U ) continuous; they are then homeomorphisms. If
(x, U ), (y, V ) U ∩ V 6= ∅ are two charts and φ̂, ψ̂ are the corresponding charts on
L(T M ), then: ψ̂ ◦ φ̂−1 : U ∩ V × GL(Rn ) → U ∩ V × GL(Rn ) is given by:
This is smooth (because M is smooth) and so the maximal atlas determined by the
charts φ̂ is a smooth structure on L(T M ) with respect to which the maps φ are
smooth diffeomorphisms and π is smooth !
Example III.13. Let M be a smooth manifold and L(T M ) its frame bundle, a local
section: σ : U → L(T M ) is a smooth choice of linear frame of Tp M for each p ∈ U .
Indeed by definition for each p ∈ U , σ(p) ∈ Lp (T M ) = GL(Rn , Tp M ) hence setting
for each i ∈ {1, . . . , n}, Ei (p) = σ(p) · ei , we get n-smooth vector fields that span
Tp M for each p ∈ U .
In particular if (x, U ) is a local chart: p ∈ U 7→ x−1 ∗x(p) is a smooth section of
L(T M ) on U and the corresponding frame is the coordinate frame ( ∂x∂ i ).
Remark III.3. Smooth sections on L(T M ) are equivalent to local bundle trivialisa-
tions, indeed, define:
φ : U × H −→ π −1 (U )
(p, h) 7−→ σ(p) · h
This is a smooth bijective map since the H -action is smooth, the derivative: φ∗(p,h) =
Rh∗σ(p) σ∗p + Lσ(p) ∗h this is a bijective map at each p and so by the local inversion
theorem φ is invertible and φ−1 is a bundle chart. Conversely if φ : π −1 (U ) → U ×H ,
then σ(p) = φ−1 (p, e) is a local section on M . It follows in particular that a principal
bundle is trivial (i.e. a product) if and only if there is a global section. Moreover, in
general we should only expect there to be local sections.
The frame bundle L(T M ) on M will play the same role as G = Rn ⋊ GLn (R) in
the case of affine space, we now want an analogue of the Maurer-Cartan form ωG ...
43
III.H. Affine connections on manifolds and curvature
Throughout this section let G = Rn ⋊ GLn (R), H = GLn (R) and g, h their
Lie algebras.
We are one step away from our definition of affine connections. Since the right
action of GLn (R) preserves the fibres we have a natural generalisation of a funda-
mental vector field.
Definition III.10
Let X ∈ h, then define a vector field X ∗ on L(T M ) by:
d
Xp∗ = (p exp(Xt))
dt t=0
2. Rh∗ ω = Adh−1 ω , h ∈ H
44
Definition III.12: Curvature of a connection
Let ω be an affine connection on a smooth manifold, the curvature form Ω is
the g-valued 2-form defined by:
1
Ω = dω + [ω, ω].
2
We shall say that the connection is torsion-free if Ω is h-valued.
The intuitive meaning of this definition is that we are using affine geometry as
a local model for a geometry on M . The Cartan connection tells us how to “slide”
affine space along M and the curvature Ω is an infinitesimal measure of the differ-
ence between M and the model space. The curvature has the following remarkable
property:
Proposition III.8
dΩ + [ω, Ω] = 0.
Proof. This is a short computation that reduces to the Jacobi identity of the Lie
algebra g:
1
dΩ + [ω, Ω] = [dω, ω] + [ω, Ω] = [Ω − [ω, ω], ω] + [ω, Ω]
2
1
= − [[ω, ω], ω].
2
However, for any vector fields X, Y, Z ∈ Γ(T M ):
[[ω, ω], ω](X, Y, Z) = [[ω(X), ω(Y )], ω(Z)] + [[ω(Y ), ω(Z)], ω(X)]
+ [[ω(Z), ω(X)], ω(Y )]
=0
Our definition of affine connection is slightly more general than the classical def-
inition. Let us make an algebraic observation:
Proposition III.9
Let g be the Lie algebra of Rn ⋊GLn (R), h the Lie subalgebra of GLn (R), then
the GLn (R)-module g (with the adjoint representation) admits a H -module
decomposition:
g = Rn ⊕ h.
45
Proof. Recall that:
M b
g= , M ∈ Mn (R), b ∈ R .
n
0 0
which gives us immediately the vector space decomposition,
we just
need to check
A 0
that each component is indeed Ad(H)-invariant. Let h = , let us first con-
0 1
I b
sider the Rn component and an element: x = n , then:
0 1
−1
−1 A 0 In b A 0 In Ab
hxh = = .
0 1 0 1 0 1 0 1
This proves that the Rn component is Ad(H)-invariant, and the adjoint representa-
tion reducesto the
standard representation of GLn (R). A similar calculation shows
B 0
that if: x = then:
0 1
−1 ABA−1 0
hxh = .
0 1
We conclude that the GLn (R) component of the direct sum decompostion is also
Ad(H) invariant and reduces to the standard adjoint representation of H on h =
Mn (R).
This means that any affine connection ω can be decomposed as: ω = θ + γ ,
where θ is a Rn valued one-form with Rh∗ θ = h−1 θ and γ a h-valued one form with
Rh∗ γ = Adh−1 γ .
Remark III.4. For some calculations it is nice to notice how to calculate the bracket
in terms of the semi-direct product of Lie algebras, g = Rn ⊕ h, we have:
46
Now:
(Rh∗ θ)up = (up ◦ h)−1 ◦ (π ◦ Rh )∗up = h−1 ◦ θup .
(Where we have used that π ◦ Rh = π ), it therefore satisfies the correct transform
rule (we say that it is GLn (R)-equivariant). Furthermore: θ(X ∗ ) = 0 for any X ∈ h.
Proposition III.10
1. Rh∗ γ = Adh−1 α,
We will now assume that all the affine connections we consider arise in this
manner from a linear connection.
Proposition III.11
1
Ω = dθ + [γ, θ] + dγ + [γ, γ] .
| {z } | 2
{z }
Rn -valued 2-form
gln (R) valued 2-form
47
Proposition III.12: Bianchi identities
1. dT + [γ, T ] = [R, θ]
2. dR + [γ, R] = 0
where:
1 X
[R, θ](X1 , X2 , X3 ) = ε(σ)R(Xσ(1) , Xσ(2) ) · θ(Xσ(3) ).
2 σ∈S
3
To understand this formula recall that R is gln (R)-valued and therefore acts
naturally on the Rn -valued form θ.
Recall that: σ(p) : Rn → Tp M . Define, Ei (p) = σ(p) · ei where (ei ) is the canonical
basis of Rn , and define (ω i (p)) the dual basis in Tp∨ M , one can therefore, express,
σ ∗ θ as the Rn valued one form:
ω1
σ ∗ θ = ... . (III.5)
n
ω
48
The pullback of the canonical one form is hence simply the map that to a vector
associates its coordinates in the basis determined by the frame σ(p) and is clearly an
isomorphism in the fibres.
The property Rh∗ ω = Adh−1 ω , tells us how to change local gauge. Suppose that
σ̃ : V → L(T M ) is another local gauge defined on V such that U ∩ V 6= ∅. Let us
set ωV = σ̃ ∗ ω and compare ωV and ωU on U ∩ V . First note that there is in fact
a smooth function h : U ∩ V 7→ GLn (R) such that σ̃ = σh = Rh ◦ σ . For every
p ∈ U ∩ V , h(p) is simply the map that changes basis of the frame at p. We claim
that:
(σ̃ ∗ ω) = Adh−1 ωU + h∗ ωH ,
where ωH is the Maurer-Cartan form of H . To prove this, we need to evaluate:
ωσ(p)h(p) ◦ (Rh ◦ σ)∗p ,
The key is to understand how to calculate the derivative of L(T M ) × H → L(T M ),
(r, h) 7→ rh, in terms of those of the maps: Rh and Lr : H → L(T M ), h 7→ rh, using
the isomorphism: T(r,h) (L(T M ) × H) = Tr L(T M ) × Th H it can be described as the
map:
(X, Y ) 7→ Rh∗r X + Lr∗h (Y ).
We therefore have:
(σ̃ω)p = ωσ(p)h(p) · Rh(p) ∗r σ∗p +ωσ(p)h(p) · Lσ(p) ∗h(p) ◦ h∗p .
| {z }
∗
σ ∗ (Rh(p) ω)=σ ∗ (Adh(p)−1 ω)=Adh(p)−1 ωU
Where γ : I → H is any curve such that γ(0) = h(p) and γ̇(0) = h∗p Xp , but we
can take: γ(t) = h(p) exp(tωH h(p) h∗p Xp ) (recall that ωH h(p) = Lh(p)−1 ∗h(p) ). It then
follows that:
Lσ(p) ∗h(p) ◦ h∗p Xp = (ωH h(p) h∗p Xp )∗σ(p)h(p) .
But, ω sends fundamental vector fields to their generator, so that:
ωσ(p)h(p) · Lσ(p) ∗h(p) ◦ h∗p = h∗ ωH .
Proposition III.13
49
Remark III.6. Of course the θ component is known so one only needs the data that
determines γ . If σ̃ = σh for some map h : U ∩ V → GLn (R) then
∀p ∈ U ∩ V, (σ̃ ∗ θ)p = h(p)−1 (σ ∗ θ)p ,
simply because θ(X ∗ ) = 0, so it follows that:
σ̃ ∗ γ = h∗ ωH + Adh−1 σ ∗ γ.
Hence, we only need a family of h-valued one forms, γUi such that the above is
satisfied between two local sections.
such that:
∇X (f Y ) = X(f )Y + f ∇X Y.
50
Ei (p) = σU (p) · ei . Any vector field X can be written on U , X = X i Ei where X i are
smooth functions on U . On U define:
∇Y X = Y (X i )Ei + Ei [(σU∗ γ)ij · Y ]X j . (IV.1)
Recall that (σU∗ γ) · Y ∈ gln (R) = Mn (R) is matrix, equivalently one can view σU∗ γ as
a matrix whose components are one-forms, which justifies the notation: (σU∗ γ)ij · Y .
We must check that the above formula defines a vector field on all of M . For this,
we must investigate how this transforms under a change of local section. Suppose
σV : V → L(T M ) is another local section defining vector fields Ẽi = σV · ei . Let
h : U ∩ V → GLn (R) be the smooth map such that, σV = σU h. Write X = X̃ i Ẽi
then:
X1 X̃ 1
.. ..
. = h . .
Xn X̃ n
So
Y (X 1 ) X̃ 1 Y (X̃ 1 )
.. . .
. = h∗ (Y ) .. + h ..
Y (X n ) X̃ n Y (X̃ n )
It follows that:
X̃ 1
Y (X i )Ei = Y (X̃ i )Ẽi + Ẽ1 · · · E˜n h−1 h∗ (Y ) ...
X̃ n
Now σV∗ γ = Adh−1 σU∗ γ + h∗ ωgln (R) , or: σU∗ γ = Adh σV∗ γ − Adh h∗ ωgln (R) so:
Proposition IV.1
γ ij = ω i (∇ej ).
51
IV.B. Torsion and curvature in terms of the covariant derivative ∇
Recall that ω is said to be torsion free if the torsion form T vanishes. We will now
translate this into a property of the covariant derivative:
Proposition IV.2
∇X Y − ∇Y X = [X, Y ], X, Y ∈ Γ(T M )
d(σ ∗ θ) + [σ ∗ γ, σ ∗ θ] = 0.
Using Equation (III.5), and writing (Γij ) = σ ∗ γ the local connection form we get the
following system of equations:
X
dω i + Γij ∧ ω j = 0.
j
Let us calculate:
Now by definition:
Hence:
Therefore:
It follows that:
52
We will now investigate how the other part of the curvature of the connection –
the curvature form of γ – translates in terms of the covariant derivative. As is ap-
parent from the above, curvature measure commutators of derivatives, let us define
the Riemann tensor of the connection:
R(X, Y )Z = ∇X ∇Y Z − ∇Y ∇X Z − ∇[X,Y ] Z
Proof. Exercice.
We shall now derive the link between R(X, Y )Z and R = dγ + 12 [γ, γ]. Choose a
local section σ : U → L(T M ) and pull back R to a h-valued one form on U that we
shall denote by (Ωij ). Again write σ ∗ γ = (Γij ). Now for two matrix-valued one-forms
α, β , define:
(α ∧ β)ij = αki ∧ βjk .
We can then write:
[σ ∗ γ, σ ∗ γ] = 2(Γ ∧ Γ).
It follows that:
Ωij = dΓij + Γki ∧ Γkj .
Proposition IV.4
Proof. Exercice.
53
Remark IV.1. An alternative way of thinking about a vector field along γ is as a section
of the pullback of the tangent bundle along γ . This is a vector bundle on I defined
by:
γ ∗ T M = {(t, v) ∈ I × M, π(v) = γ(t).}
We now state:
Proposition IV.5
1. ∀t ∈ I, dt
D
(f Y )(t) = f˙(t)Y (t) + f (t) dt
D
Y (t), f : I → R smooth,
Remark IV.2. In essence, this is just the pullback of the connection to the pullback
bundle.
Definition IV.2
Let γ : I → M be a smooth curve, we define its acceleration to be: ∇γ̇ γ̇ . γ is
said to be a geodesic if:
∇γ̇ γ̇ = 0.
A vector field Y along the curve γ is said to be parallel along the curve if:
∇γ̇ Y = 0.
V. Riemannian geometry
Up to now we have developed a curved version of affine geometry built around the
homogeneous model G/H where G is the affine group and H the linear group. Eu-
54
clidean geometry can also be thought of as the study of a homogeneous space G/H ,
where G = Rn ⋊ O(n) and H = O(n). There is again a Maurer-Cartan form and one
can again interpret G as a frame bundle of Euclidean space. However, the relevant
frames for Euclidean geometry are orthonormal frames.
To get a curved version of this we can still base our construction on the frame
bundle L(T M ), but this is a GLn (R)-bundle and not a On (R)-bundle. We therefore
need to reduce L(T M ) to an On (R)-bundle, which is essentially specifying which
frames are orthonormal.
On the model this is achieved by the scalar product; so we can expect that to
generalise this to manifolds we should choose a scalar product on each Tp M , this
leads to the following definition:
g : Γ(T M ) × Γ(T M ) → C ∞ (M ),
We call (M, g) a Riemannian manifold. Given this extra structure there is a canon-
ical choice of connection on L(T M )
Now add the first two equations and substract the last one, and use that the con-
nection is torsion free to find:
X(g(Y, Z)) + Y (g(X, Z)) − Z(g(X, Y )) = 2g(∇X Y, Z) − g([X, Y ], Z)
+ g([X, Z], Y ) + g(X, [Y, Z]).
55
Hence:
2g(∇X Y, Z) = X(g(Y, Z)) + Y (g(X, Z)) − Z(g(X, Y ))
+ g([X, Y ], Z) + g([Z, X], Y ) + g([Z, Y ], X).
i.e.
(∇X g)(Y, Z) = X(g(Y, Z)) − g(∇X Y, Z) − g(Y, ∇X Z).
The compatibility condition between ∇ and g can then simply be stated as the van-
ishing of the covariant derivative of g .
Let V be a vector space S 2 V ∨ the space of symmetric bilinear forms on V . If
α, β ∈ V ∨ , define:
1
(α β)(v1 , v2 ) = (α(v1 )β(v2 ) + α(v2 )β(v1 )) .
2
If (e1 , . . . , en ) is a basis of V , then: (ei `ej )i≤j is a basis of S 2 V ∨ .
On manifolds, define S 2 (T ∨ M ) = p∈M S 2 Tp∨ M and equip it with a topology
and smooth structure as we have done for the other bundles. Any local section
σ : U → L(T M ), corresponding to vector fields (E1 , . . . , En ) that at each point
p ∈ U spans Tp M , gives rise to smooth fields: (ω i ω j )i≤j that at each p ∈ U span
S 2 Tp∨ M where (ω i ) is the dual frame. We will often use this to expression the metric
locally.
Example V.1. The standard Riemannian metric on R3 can be written, in the standard
global frame:
g = dx dx + dy dy + dz dz.
It is custom to write: dx dx = dx2 . Let us now consider the unit sphere S 2 in
R3 , and introduce spherical coordinates according to x = cos φ sin θ, y = sin φ sin θ,
z = cos θ. Then on S 2 we have the relations:
dx = − sin φ sin θdφ + cos φ cos θdθ,
dy = cos φ sin θdφ + sin φ cos θdθ, dz = − sin θdθ
56
Let us determine the Levi-Civita connection in the local orthonormal frame:
E1 = ∂θ∂
, E2 = sin1 θ ∂ϕ
∂
. Let ω 1 = dθ, ω 2 = sin θdφ. By definition it is the unique
one-form valued matrix (Γij ) such that:
dω i + Γij ∧ ω j = 0, (torsion-free).
Let us denote an equivalence class by [p, v]. The idea behind this is that the couple
(p, v) consisting of a frame p and the expression of the vector in the frame p. The
vector is then the equivalence class of all couples under a change of frame. The
projection, π̃ is given by factorisation of the map (p, v) → π(p); this is possible
because: π(ph) = π(p) and so π(p) only depends on the equivalence class of [p, v].
57
If φ : π −1 (U ) → U × H is a local trivialisation, we construct a vector bundle chart
by factorisation of:
π̂ −1 (U ) = π −1 (U ) × V −→ U ×V
(p, v) 7−→ (π(p), πH (φ(p))v)
Since φ(ph) = φ(p) · h. These will be our local vector bundle charts, it is then an
exercise to show that equipping P ×H V with the quotient topology, constructing
charts from these in the usual way defines a smooth structure on P ×H V that makes
these smooth vector bundle charts.
Example VI.1. Let P = L(T M ) be the frame bundle of a smooth manifold M and
ρ the standard representation of GLn (R) on V = Rn . Then P ×H L(T M ) ∼ = TM.
The isomorphism is given by [up , v] 7→ up · v ∈ Tp M ,→ T M , this is a vector space
isomorphism between the fibres and therefore a vector bundle isomorphism. All
tensor bundles arise in this way as associated bundles to the frame bundle.
We will now use associated vector bundles to relate objects on L(T M ) to objects
on M . First let us consider the h-valued one form γ , γup : Tup P → h is no longer
an isomorphism and has a kernel. Define Vup = ker π∗up , where π : P → M is
the projection . We shall refer to these vectors as vertical vectors; intuitively these
are vectors that point in the fibres of L(T M ). Since imγup = Vp by the property
γ(X ∗ ) = X , its kernel is a complementary subspace to Vp that we shall call the
horizontal subspace, Hp . Intuitively, this is a choice of subspace that is isomorphic
to Tp M in Lup (T M ). Any vector X in Tup L(T M ) therefore has a decomposition
X = XH + XV .
Now consider a representation ρ of H on some vector space V , and say that a
V -valued k -form α on L(T M ) equivariant if:
Rh∗ α = ρ(h)−1 · α.
Proposition VI.1
58
Example VI.2. • The curvature dα + 12 [α, α] is equivariant (with respect to the
adjoint representation) and horizontal; the corresponding field on M is the
Riemann tensor.
• The solder form is a horizontal and equivariant one form (with respect to the
standard representation).
59
A. Compactness in second-countable spaces
It is well known that for metric spaces compactness is equivalent to sequential com-
pactness, i.e. the following two notions of compactness coincide:
Definition: Compactness
For second-countable spaces, these notions also coincide. Let B = (Vn )n∈N be a
countable basis for the topology:
• Assume first compactness, then X is sequentially compact. Recall first that:
– A family of closed sets (Fi )i∈I in X is said to satisfy the finite intersection
property if for any finite subset J ⊂ I , ∩j∈J Fj 6= ∅.
– A Hausdorff topological space X is compact if and if for any family of
closed subsets (Fi )i∈I that satisfy the finite intersection property one has
∩i∈I Fi 6= ∅.
Indeed, suppose first that there is a family (Fi )i∈I of closed subsets that
satisfy the finite intersection property but such that ∩i∈I Fi = ∅. Then
(X \ Fi )i∈I is a family of open subsets that has no finite subcover, hence
X is not compact. Conversely, suppose that X is not compact, then
there is an open cover (Ui )i∈I of X that has no finite subcover, but then
(Fi = X \ Ui )i∈I is a family of closed sets with the finite intersection
property such that ∩i∈I Fi = X \ ∪i∈I Ui = ∅.
The decreasing family of closed subsets (Fn )n∈N , defined by Fn = {xm , m ≥ n}
for some sequence (xn )n∈N clearly satisfies the finite intersection property,
hence, X being compact ∩n∈N Fn 6= ∅. Choose x ∈ ∩n∈N Fn then one can con-
struct a subsequence converging to x in the following manner. Observe that
one can construct from B a non-increasing basis of open neighbourhoods of
x, (Vmx )m∈N ; this will play a role similar to open balls in the metric case. We
construct an increasing extraction map φ : N → N by induction. First, for
n = 0, since V0x is a neighbourhood of x there is an infinite number of n ∈ N
such that xn ∈ V0x , let φ(0) = min{n ∈ N, xn ∈ V0x }. Suppose that we have de-
fined φ(k) for k ∈ J0, nK such that φ(0) < φ(1) < · · · < φ(n) and for all k ∈ N,
xϕ (k) ∈ Vkx . To define φ(n + 1), using the fact that {n ∈ N, xn ∈ Vn+1x
} is infi-
nite one can again choose φ(n+1) > φ(n) such that xϕ(n+1) ∈ Vn+1 . (xϕ(n) )n∈N
x
60
converges to N, because if V is any neighbourhood of x, then there is n0 ∈ N
such that Vnx0 ⊂ V , by definition of φ for any n ≥ n0 , xϕ(n) ∈ Vnx ⊂ Vnx0 ⊂ V.
• Assume now sequential compactness and let (Ui )i∈N be an arbitrary cover of
X by open sets.
1. (Ui )i∈N has a countable subcover, indeed for every x ∈ X, ∃ix ∈ I, x ∈
Uix , but one can find nx such that Vnx ⊂ Uix . However: {nx , x ∈ X} = C
is countable and ∪c∈C Vc covers X . Moreover by construction, for each
c ∈ C , one can find at least one i(c) such that Vc ⊂ Ui(c) . (Ui(c) )c∈C is
then a countable subcover.
2. Countable covers have finite subcovers. Let (Un )n∈N be a countable cover
of X by open sets. Suppose that it has no finite subcover then for every
n ∈ N there is xn ∈ X \ ∪ni=1 Ui . By sequential compactness, (xn )n∈N has
a convergent subsequence. Let us call the limit x ∈ X ; by construction:
x ∈ X \ ∪ni=1 Ui for all n ∈ N. Which contradicts the fact that (Un )n∈N
covers X . Therefore (Un )n∈N has a finite subcover.
61
B. Tensor algebra, exterior algebra, differential forms
(to be completed)
The following is a short review on the notion of tensor algebra, it gives a more ab-
stract foundation to our notion of exterior algebra. We begin with the basic propo-
sition that we shall understand as our definition for tensor products:
Let V1 , V2 be two vector spaces over the same field K, then up to isomorphism
there is a unique K-vector space written V1 ⊗ V2 equipped with a bilinear map:
j : V1 × V2 → V1 ⊗ V2 satisfying the following universal property: for any
bilinear map b : V1 × V2 → V3 , there is a unique linear map: l : V1 ⊗ V2 → V3
such that the following diagram commutes:
V1 × V2 b
V3
l
j
V1 ⊗ V2
Proof. See your favourite reference in general algebra; for instance: [Hun74] or [Lan05].
Remark B.1. • The same construction is valid for modules over rings.
• Note that elements of V1 ⊗ V2 are not all of the form v1 ⊗ v2 , they are linear
combinations of such products. Indeed, let l : V1 ⊗ V2 → K be a linear
map that vanishes on any element of the form vi ⊗ vj , then l fits into the
commutative diagram:
V1 × V2 0
R
l
j
V1 ⊗ V2
However, this diagram the null map fits into this same diagram, therefore by
uniqueness: l = 0.
62
• K⊗V ∼
= V (map v ∈ V to 1 ⊗ v )
The tensor product of two spaces solves the problem of exchanging bilinear maps
for linear maps. As an exercise let us prove that ⊗ is associative in the following
sense: (V1 ⊗ V2 ) ⊗ V3 ∼ = V1 ⊗ (V2 ⊗ V3 ). Using the universal property there is a
unique linear map L : (V1 ⊗ V2 ) ⊗ V3 → V1 ⊗ (V2 ⊗ V3 ) such that (v1 ⊗ v2 ) ⊗ v3 7→
v1 ⊗ (v2 ⊗ v3 ) for all (v1 , v2 , v3 ) ∈ V1 × V2 × V3 . It is the required isomorphism,
indeed construct in the same way the map L̃ : V1 ⊗ (V2 ⊗ V3 ) → (V1 ⊗ V2 ) ⊗ V3 such
that v1 ⊗ (v2 ⊗ v3 ) 7→ (v1 ⊗ v2 ) ⊗ v3 then L ◦ L̃ fits into the diagram:
j
V1 ⊗ (V2 ⊗ V3 ) V1 ⊗ (V2 ⊗ V3 )
L◦L̃
j
V1 × (V2 ⊗ V3 )
But this diagram is also satisfied by id, hence, by uniqueness L̃ ◦ L = id. Inversing
the roles of the spaces we also get L ◦ L̃ = id.
We are particularly interested in the case where V1 = V2 , in this case we will write
V ⊗ V = ⊗2 V , then define by induction: ⊗n V = V ⊗ (⊗n−1 V ) and set:
M
+∞
T (V ) = ⊗k V,
k=0
63
| × ·{z
for any u : V · · × V} → W , k -multilinear, alternated, there is a unique linear map
k times
V
ũ : k V → W such that if j is the map: (v1 , . . . , vk ) 7→ v1 ∧ · · · ∧ vk then u = ũ ◦ j .
(This can be shown using the universal property of ⊗k V and taking the quotient).
With this in mind we can show that if (e1 , . . . , ek ) is a basis of V then,
64
References
[Hun74] T. W. Hungerford. Algebra. Graduate Texts in Mathematics. Springer
New York, NY, 1974.
[Lan05] S. Lang. Algebra. Graduate Texts in Mathematics. Springer New York,
2005. ISBN: 9780387953854.
[Lan95] S. Lang. Differential and Riemannian Manifolds. Third Edition. Graduate
Texts in Mathematics. Springer-Verlag New York, 1995.
[Lax02] P. D. Lax. Functional Analysis. Pure and Applied Mathematics. Wiley, 2002.
[Lee03] J. M. Lee. Introduction to Smooth Manifolds. Second Edition. Graduate Texts
in Mathematics. Springer, 2003.
[Nab10] G. Naber. Topology, Geometry and Gauge fields : Foundations. Texts in Applied
Mathematics. Springer, 2010.
[Rud69] M. E. Rudin. “A new proof that metric spaces are paracompact”. In: Proc.
Amer. Math. Soc. Vol. 20. 1969, p. 603.
[Sha97] R. Sharpe. Differential Geometry: Cartan’s Generalization of Klein’s Erlangen
Program. Vol. 166. Graduate Texts in Mathematics. Springer-Verlag New
York, 1997.
65
Glossary
Basis (Topology) A collection B of open subsets of a topological space X such that
any open subset is a union of elements of B . A collection of open subsets B
is a basis if and only if for any x ∈ X and any open subset U such that x ∈ U ,
one can find B ∈ B such that x ∈ B ⊂ U . Example: open balls in metric
spaces.
Bump function A bump function is a smooth map f : M → R with compact sup-
port. For any n ∈ N∗ , one can find bump functions f : Rn → R. Indeed, first
consider: f : R → R+ defined by:
( 1
e− t t > 0
f (t) = .
0 t≤0
Closed graph theorem Let E, F be Banach spaces and u ∈ L(E, F ) then u is con-
tinuous if and only if its graph:
G = {(x, f (x)), x ∈ E}
is closed in E × F .
Comparison of topologies Let T1 , T2 , be two topologies on the same set X , we say
that T1 is finer than T2 , written: T2 ⊂ T1 if U ∈ T1 ⇒ U ∈ T2 . In this case T2 is
also said to be coarser than T1 . The analogy is that of grains of sand, the finer
the topology the more ”grains of sand” i.e. open sets.
66
1. ◦ is associative and for each object O1 of C ,
2. there is a morphism: idO1 : O1 → O1 which acts as a left and right
identity under the composition law.
The class of all sets is a category where morphisms are maps between sets,
topological vector spaces form a category where morphisms are continuous
linear maps, smooth manifolds form a category where the morphisms are
smooth maps, Banachisable spaces form a category where morphisms are con-
tinuous linear maps, etc. A functor between categories C1 and C2 , is a mapping
between categories that:
• maps any object O1 of C1 to an object F (O1 ) of C2
• maps any morphism f : O1 → O2 to a morphism F (f ) : F (O1 ) →
F (O2 ), such that F (idO1 ) = idF (O1 ) and F (f ◦ g) = F (f ) ◦ F (g).
Maps between categories that satisfy all the above conditions except that they
reverse the order of composition, i.e. F (f ◦ g) = F (g) ◦ F (f ) are called
contravariant-functors. They are functors in the above sense on the opposite
category. See [Hun74, Chapter I.7, Chapter X] for a more detailed account.
The map that maps a Banach space to its dual is a contravariant functor.
Locally compact topological space A Hausdorff topological space such that every
point has a neighbourhood with compact closure (i.e. relatively compact).
67
If we defined: r · m = m · r then this property would translate to:
Paracompact topological space A Hausdorff space with the property that every open
covering admits a locally finite refinement. Guarantees the existence of parti-
tions of unity. All metric spaces are in fact paracompact [Rud69].
Regular space A topological Hausdorff space X in which one can separate points
and closed sets with open sets. i.e. if x ∈ X , F ⊂ X closed and x ∈
/ F then
there are disjoint open sets U and V such x ∈ U , F ⊂ V . This is equivalent
to the statement that for any x ∈ X and any open subset U of X , one can find
an open subset V such that x ∈ V ⊂ V ⊂ U .
The subspace topology Let (X, T ) be at topological space, A ⊂ X , then the topol-
ogy of X induces a natural topology on A, defined by:
TA = {U ∩ A, U ∈ T }.
Topological group A group (G, ·) equipped with a topology compatible with its
group structure, i.e. the maps µ : G × G → G, ι : G → G defined by:
µ(g1 , g2 ) = g1 · g2 , ι(g) = g −1 , g1 , g2 , g ∈ G,
are continuous.
68
Topological space A topology on a set X is a collection T of subsets of X that
satisfies the following stability conditions:
1. ∅, X ∈ T ,
2. Arbitrary unions of elements of T are elements of T ,
3. Finite intersections of elements of T are elements of T .
Elements of T are called open sets. A topology is the minimum set of data
required to talk about limits, neighbourhoods, continuous maps, etc.
French-English Dictionary
arbitrary quelconque.
bundle fibré.
chart carte.
countable dénombrable.
data données.
hence d’où.
induction récurrence.
manifold variété.
69
map application.
neighbourhood voisinage.
non-decreasing croissant.
non-increasing décroissant.
onto surjectif.
regular regulier.
sequence suite.
smooth lisse.
submanifold sous-variété.
subsequence sous-suite.
70