0% found this document useful (0 votes)
12 views28 pages

Géométrie Semi-Riemannienne: Loubeau

The document is a set of lecture notes on semi-Riemannian geometry, focusing on differentiable manifolds, tangent spaces, tensors, and geodesics. It includes definitions, theorems, and examples related to these concepts, structured into chapters and sections. The notes were compiled by Téofil Adamski for a Master 2 mathematics course at the Université de Rennes.

Uploaded by

Angelo Oppio
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
12 views28 pages

Géométrie Semi-Riemannienne: Loubeau

The document is a set of lecture notes on semi-Riemannian geometry, focusing on differentiable manifolds, tangent spaces, tensors, and geodesics. It includes definitions, theorems, and examples related to these concepts, structured into chapters and sections. The notes were compiled by Téofil Adamski for a Master 2 mathematics course at the Université de Rennes.

Uploaded by

Angelo Oppio
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 28

Géométrie semi-riemannienne

Éric Loubeau
Master 2 de mathématiques fondamentales · Université de Rennes
Notes prises par Téofil Adamski (version du 5 avril 2023)
i

Sommaire

1 Differentiable manifolds
1.1 Differentiable manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Tangent spaces and tangent bundle . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2 Semi-riemannian manifolds
2.1 First definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.2 Connection and Levi-Civita connection . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.3 Curvature and Ricci tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.4 Killing vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
3 Geodesics
3.1 First definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
4 Examples
5 Calculus of variations
1

Chapitre 1
Differentiable manifolds

1.1 Differentiable manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


1.2 Tangent spaces and tangent bundle . . . . . . . . . . . . . . . . . . . . . . . 1
1.3 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.1. Differentiable manifolds


Let S be a topological connected Hausdorff and paracompact space. A chart is a homeomor-
phism ξ from and open subset of S in a open subset η(U ) ⊂ Rn . It can be written

ξ(P ) = (x1 (p), . . . , xn (p)), ∀p ∈ U

where the maps xi are called the coordinates functions of ξ and we will denote ξ = (x1 , . . . , xn ).
Two charts ξ and η of dimension n intersect in a smooth manner if the maps ξ ◦ η −1 and η ◦ ξ −1
are of classe C ∞ .
An atlas is a collection of charts of dimension n such that

– for all point p ∈ S, there exist an open subset U such that p ∈ U ;


– two charts intersect in a smooth manner.

An atlas is complete if it contains all the charts of S which intersect in a smooth manner. Any atlas
admits a completion.

Définition 1.1. A differentiable manifold is a topological space equipped with a complete atlas.

Exemples. – The euclidean space is a differentiable manifold.


– The sphere Sn ⊂ Rn+1 is a differentiable manifold of dimension n.
– A cartesian product of differentiable manifolds is also a differentiable manifold.

Définition 1.2. Let M be a differentiable manifold. A function f : M −→ R is of class C ∞ if, for


any chart (U, η), the maps
f ◦ η −1 : η −1 (U ) −→ R
is of class C ∞ .
The sum, product and inverse are of class C ∞ .

Définition 1.3. Let M and N be two differentiable manifolds. A map ϕ : M −→ N is of class C ∞


if, for any charts (U, ξ) of M and (V, η) of N , the map
η ◦ ϕ ◦ ξ −1 : ξ(U ) −→ η(V ).
is of class C ∞ .
2 1.2. Tangent spaces and tangent bundle

1.2. Tangent spaces and tangent bundle


Définition 1.4. Let p ∈ M a point. Let F (M ) be the space of functions of class C ∞ on M . A
tangent vector at the point p is a R-linear map v : F (M ) −→ R satisfying the Leibniz rules
v(f g) = f (p)v(g) + g(p)v(f ).
The space of all tangent vectors at the point p is the tangent space at the point p, denoted Tp M .

Let (U, ξ) a chart, p ∈ U a point and f ∈ F (M ) a function. We denote η = (x1 , . . . , xm ) and


∂f ∂(f ◦ η −1 )
(p) := (η(p)).
∂xi ∂ui
where the notation ui are the coordinates on Rm . The map

∂i |p := : F (M ) −→ R
∂xi p

are a tangent vector at the point p. The vectors ∂i |p form a basis of Tp M .


Définition 1.5. Let ϕ : M −→ N be a map of class C ∞ . For all point p ∈ M , we define the R-linear
map
dϕp : Tp M −→ Tϕ(p) N
by the equality
dϕp (v) = vϕ ∈ Tϕ(p) N
where
vϕ (g) := v(g ◦ ϕ).

With coordinate (x1 , . . . , xm ) on M and (y 1 , . . . , y n ), we have


n
X ∂(y i ◦ ϕ) ∂
dϕ(p)(∂j |p ) = (p) .
i=1
∂xj ∂y i ϕ(p)

Remarque. If the maps ϕ : M −→ N and ψ : N −→ P are smooth, then the composition ψ ◦ ϕ is


also a smooth map.
Définition 1.6. A vector field is a a map V which send each point p ∈ M on a tangent vector
V p ∈ Tp M .

If f ∈ F (M ), we denote V (f )(p) := Vp (f ).

Définition 1.7. If V (f ) is of class C ∞ for all f ∈ F (M ), then V is of class C ∞ .


The sum of two vector fields is a vector field. The multiplication of a vector field by a function
is a vector field. The bracket of two vector fields V and W is defined by
[V, W ]p (f ) := Vp (W (f )) − Wp (V (f )).
It is skew-symmectric R-bilinear. It satisfies the Jacobi identity
[X, [Y, Z]] + [Y, [Z, X]] + [Z, [X, Y ]] = 0.
We have
[f X, gY ] = f g[X, Y ] + f (X(g))Y − g(Y (f ))X.

Exemple. We have [∂i , ∂j ] = 0.


Définition 1.8. A differentiable manifold P is a sub-manifold of M if
– P ⊂M;
– the injection map j : P ,−→ M is a map of class C ∞ ;
– it differential djp : Tp P −→ Tj(p) M is injective for all p ∈ P .
Chapitre 1. Differentiable manifolds 3

Théorème 1.9 (Whitney). Let M be a C ∞ -differentiable manifold of dimension n. Then there


exists an immersion M −→ R2n .
F
The tangent bundle of M is TM := p∈M Tp M . With TM = {(p, v) | p ∈ M, v ∈ Tp M }, we
have a natural map π : TM −→ M which satisfies π −1 (p) = Tp M . We can show that the tangent
bundle is a manifold of dimension 2n. Indeed, let be (U, ξ) a chart on M with ξ = (x1 , . . . , xn ).
Let v ∈ Tp M . We can write
X ∂
v= vi i
∂x p

with v i ∈ R. We consider
π −1 (U ) ⊂ TM −→ R2n ,
η̃ :
(p, u) 7−→ (x1 (π(p, u)), . . . , xn (π(p, u)), v 1 , . . . , v n ).
where v i = v(xi ) =: ẋi (u). This defines a atlas on TM . If (u1 , . . . , u2n ) are the coordinates on R2n ,
the transition functions are given by
ui ξ˜ ◦ η̃ −1 = xi ◦ π ◦ η̃ −1 (a, b) = xi η −1 (a),
X ∂xi
ui ξ˜ ◦ η̃ −1 = ẋi ◦ η̃ −1 (a, b) = bk k (η −1 (a)).
∂y
So the map ξ˜ ◦ η̃ −1 are of class C ∞ .

Remarque. A vector field X : M −→ TM is a map of class C ∞ such that π ◦ X = IdM .

Remarque. In general, we have TM ̸= M × Rn . This is the case for S3 .

Exemple. We consider the sphere S2 . It is a manifold of dimension 2. We want to calculate it


tangent space at a point p ∈ S2 . Let γ : ]−ε, ε[ −→ S2 be a curve of class C ∞ on S2 with γ(0) = p.
It acts on functions on S2 . For a function f : S2 −→ R, we denote
d
γ̇(0) · f := (f ◦ γ)(t)
dt t=0

We have

= γ̇(0) ∈ Tp S2
dt t=0

and all tangent vector can be obtained this way. As |γ| = 1, we find
d
|γ(t)|2 = 0 = 2⟨p, γ̇(0)⟩.
dt
Thus we conclude
Tp S2 = {X ∈ R3 | ⟨X, p⟩ = 0}.

Exemples. Open subsets of Rn are differentiable manifolds. The half-plane H2 := R × R∗+ has the
tangent space Tp H2 = R2 and so its tangle bundle is TH2 = H2 × R2 .

Exemples. – The image of the map


]−1, 1[ −→ R2 ,
t 7−→ (t, |t|)

is a differentiable manifold but not a submanifold of R2 .


– The map
R −→ R2 ,
t 7−→ (t3 , t2 )

is differentiable but not an immersion.


4 1.3. Tensors

– The map
R −→ R2 ,
t 7−→ (t3 − 4t, t2 − 4)
is differentiable and an immersion, but there is a self-intersection sot it is not an embedding.
– The map t 7−→ (t, sin(1/t)) is an immersion with no self-intersecting point, but it is not an
embedding.
– The cone x2 + y 2 − z 2 = 0 is not a submanifold of R3 for connectivity reasons.


1.3. Tensors
Let V be a module over a ring K.
Définition 1.10. Let r, s ∈ N be integers with rs > 0. A tensor of type (r, s) is a K-multilinear
function
(V ∗ )r × V s −→ K.
We denote Tr,s (V ) the set of tensors of type (r, s).
A tensor field is a tensor on the ring X (M ) which denotes the set of vectors field on a
differentiable manifold M . The set X (M ) is a module on the ring F(M ) of functions on M . So a
tensor field of type (r, s) is a F(M )-linear map
A : X ∗ (M )r × X (M )s −→ F(M ).

Exemple. The map


X ∗ (M ) × X (M ) −→ F(M ),
C:
(θ, X) 7−→ θ(X)
is a tensor.

Counter-example. Let ω ∈ X ∗ (M ) a linear form. The map


X (M ) × X (M ) −→ F(M ),
F:
(X, Y ) 7−→ X(ω(Y ))
is not a tensor field.
′ ′ ′ ′
Remarque. When A ∈ Tr,s (V ) and B ∈ Tr ,s (V ), we can define the tensor A × B ∈ Tr+r ,s+s (V )
with the equality
′ ′
A⊗B(θ1 , . . . , θr+r , X1 , . . . , Xs+s′ ) = A(θ1 , . . . , θr , X1 , . . . , Xs )A(θr+1 , . . . , θr+r , Xs+1 , . . . , Xs+s′ ).
i
Proposition 1.11. Let p ∈ M and A ∈ Tr,s (M ). Let θ and θi be 1-forms which agree on p. Let X i
and Xi be vector field which agree on p. Then
1 r
A(θ , . . . , θ , X 1 , . . . , X s )(p) = A(θ1 , . . . , θr , X1 , . . . , Xs )(p).
Thus we can define the map
Ap : (T∗p M )r × (Tp M )s −→ R.

Démonstration. We show that, if θi0 (p) = 0 or Xi0 (p) = P 0, then A(θ1 , . . . , θr , X1 , . . . , Xs )(p) = 0.
n
1
Let (U, (x , . . . , x )) be a chart. Then we can write Xj0 = X i ∂i . Let f be a bump function on U
i 2
with f (p)P= 1. We have Xj0 (p) = 0 ⇔ X (p) = 0, ∀i and f Xj0 is a vector field and we can write
f 2 Xj0 = f X i (f ∂i ). So
f 2 A(θ1 , . . . , θr , X1 , . . . , Xs ) = A(θ1 , . . . , θr , X1 , . . . , f 2 Xj0 , . . . , Xs )
X
= f xI A(θ1 , . . . , θr , X1 , . . . , f ∂i , . . . , X s )
i
1 r
and A(θ , . . . , θ , X1 , . . . , Xs )(p) = 0. ⋄
Chapitre 1. Differentiable manifolds 5

Let (U, (x1 , . . . , xn )) be a map. Let p ∈ U . On U , we denote


Aij11,...,i s i1 ir
,...,js := A(dx , . . . , dx , ∂ji , . . . , ∂js )

and we have X
A= Aij11,...,i i1 ir
,...,js ∂ji ⊗ · · · ⊗ ∂js ⊗ dx ⊗ · · · ⊗ dx .
s

The contraction of A on the indices i and j is the tensor field Cji A of type (r − 1, s − 1) which is
the composition of C and the tensor
(θ, X) 7−→ A(θ1 , . . . , θ, . . . , θr , X1 , . . . , X, . . . , Xs ).
The component of Cji A are Aij11,...,m,...,i
,...,m,...,js with m ∈ {1, . . . , n}.
r

Définition 1.12. Let ϕ : M −→ N a differentiable map. If A ∈ T0,s (N ), we set


ϕ∗ A(X1 , . . . , Xs ) := A(dϕ(X1 ), . . . , dϕ(Xs )).
The tensor ϕ∗ A ∈ T0,s (M ) is the pull-back of A by ϕ.
Définition 1.13. A derivation of tensor is a R-linear map
D : Tr,s (M ) −→ Tr,s (M )
such that
D(A ⊗ B) = DA ⊗ B + A ⊗ DB
and
D(CA) = C(DA).

For a function f ∈ F(M ) ⊂ T0,0 (M ), we set f ⊗ A = f A and we have D(f A) = f DA + (Df )A.
The derivation D is a derivation of functions so there exists a V ∈ X (M ) such that Df = V (f ).
The chain rule becomes

D(A(θ1 , . . . , θr , X1 , . . . , Xs )) = (DA)(θ1 , . . . , θr , X1 , . . . , Xs )
X r Xs
+ A((θ1 , . . . , Dθi , . . . , θr , X1 , . . . , Xs )) + (θ1 , . . . , θr , X1 , . . . , DXj , . . . , Xs ).
i=1 i=1

Théorème 1.14. Given a vector field V and an R-linear map δ : X (M ) −→ X (M ) such that
δ(f X) = V (f )W + f (δX),
there exists a unique derivation of tensors which equals to δ on X (M ) and V on F(M ).
Définition 1.15. Let V ∈ X (M ). Then we set the derivation LV as
LV (f ) := V (f ) et LV (X) := [V, X]
for all f ∈ F(M ) and X ∈ X (M ). It is called the Lie’s derivation.
Définition 1.16. Let V be a vector space. The index of a bilinear form b is the dimension of the
largest subspace W ⊂ V such that the restriction b|W ×W is negative definite.
A vector v ∈ V is null or isotropic if v ̸= 0 and b(v, v) = 0.
Lemme 1.17. Let V and W be two linear spaces of the same dimension. Then they are equipped
with inner products with the same indices if and only if the exists a linear isometry V −→ W .
7

Chapitre 2
Semi-riemannian manifolds

2.1 First definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7


2.2 Connection and Levi-Civita connection . . . . . . . . . . . . . . . . . . . . . 8
2.3 Curvature and Ricci tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.4 Killing vector field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.1. First definitions


Définition 2.1. A metric on a differentiable manifold M is a tensor field g on M of type (0, 2)
which is symmetric, non-degenerate and with a constant index. A semi-riemannian manifold is a
manifold M equipped with a metric g.
In general, two different metrics on a same manifold M gives two different semi-riemannian
structures on M . If the index is zero, then we say that the semi-riemannian manifold (M, g) is
riemannian. If the index is one, then we will call it lorentzian.
In local coordinates (U, (x1 , . . . , xn )), we can write g = gi,j dxi ⊗ dxj with gi,j = g(∂i, ∂j).
P
i,j
The matrix (gi,j ) is invertible, the inverse will be denoted (g ).

Exemple. Let ν ⩽ n be a natural integer. On the space Rn , we have the semi-riemannian


structure Rnν with the metric
ν
X n
X
⟨u, v⟩ = − ui wi + v i wi .
i=1 i=ν+1

Définition 2.2. Let p ∈ M . Let (M, g) be a semi-riemannian manifold. A tangent vector v ∈ Tp M


is
– space-like if v = 0 or g(v, v) = 0 ;
– null if v ̸= 0 and g(v, v) = 0 ;
– time-like if g(u, u) < 0.
Null vectors form the null cone.
If P ⊂ M is a submanifold and M is equipped with a riemannian metric g, the P is a riemannian
manifold. For example, the sphere S2 admits a riemannian metric. But this is not always true for
semi-riemannian metrics.
Lemme 2.3. Let (M, gM ) and (N, gN ) two semi-riemannian manifolds. Let π : M × N −→ M
and σ : M × N −→ N the two projections. Then the map g := π ∗ gN + σ ∗ gN is a semi-riemannian
metric on M × N .
Définition 2.4. An isometry between two semi-riemannian manifolds (M, g) and (N, h) is a
diffeomorphism ϕ : M −→ N which preserves the metrics, that is ϕ∗ g = h or
∀p ∈ M, ∀u, w ∈ Tp M, hϕ(p) (dϕp (u), dϕp (v)) = gp (u, v).
8 2.2. Connection and Levi-Civita connection

Exemple. Let (u1 , . . . , P


un ) be the natural coordinates on Rnν .PLet V and W two vector fields
on R . We denote W =
n
W i ∂i . We define DV W := dW (V ) = V (W i )∂i . This is the covariant
derivative of W with respect to V .

2.2. Connection and Levi-Civita connection


Définition 2.5. A connection on a manifold M is a map
X (M ) × X (M ) −→ X (M ),
D:
(X, Y ) 7−→ DX Y
such that
– DV W is F(M )-linear in V ;
– DV W is R-linear in W ;
– DV (f W ) = V (f )W + f DV W .

Proposition 2.6. Let (M, g) be a semi-riemannian manifold. Let V ∈ X (M ) a vector field. Let V ∗
be the 1-form defined by
V ∗ (X) := g(V, X).
Then the map V 7−→ V ∗ is a F(M )-linear isomorphism.

Exemples. – We take the sphere S2 . For a point p ∈ S2 and two tangent vectors X, Y ∈ Tp S2 ,
we can define
gp (X, Y ) := ⟨X, Y ⟩
where the notation ⟨·, ·⟩ is the standard inner product on R3 . This gives a riemannian metric
on S2 . If we replace the inner product ⟨·, ·⟩ by another semi-riemannian metric on R3 , then
this metric is no longer semi-riemannian in general.
– If g is a riemannian metric, then a another riemannian metric is given by
g̃p (X, Y ) := ef (p) gp (x, y)
for a smooth function f ∈ C ∞ (M, R).
– There exists three vector fields Ei on S3 which form a orthonormal family where the semi-
riemannian is the same as the first example. Then we can define a new semi-riemannian
metric by
◦ ⟨Ei , Ej ⟩ = 0 for all i ̸= j ;
◦ |E1 |2 = −1 and |E2 |2 = |E3 |2 = 1.
– On the half-plane H2 , we can define the metric
dx2 + dy 2
g := .
y2

Question. When can we equip M with a semi-riemannian metric ? It is not always the case for a
semi-riemannian metric with strictly positive index. But it is always the cases for a riemannian
metric. There is two ways to do that :
– by using the Withney’s theorem : the exists an immersion ι : M −→ RN for a large enough
integer N and we take the pullback of the euclidean matric on RN , that is
gp (X, Y ) := ⟨dιp (X), dιp (Y )⟩;
– if (Ui , xi ) are an atlas of M , we define
X
gp := αi x∗i ⟨·, ·⟩Rn .
i

Théorème 2.7. Let (M, g) be a semi-riemannian manifold. Then there exists a unique connection D
such that, for all vector fields V and W , we have
Chapitre 2. Semi-riemannian manifolds 9

– [V, W ] = DV W − DW V ;
– Xg(V, W ) = g(DX V, W ) + g(V, DX W )
Moreover, the connection D is characterized by the Koszul formula
2g(DV W, X) = V (g(X, W )) + W (g(X, V )) − Xg(V, W )
− g(V, [W, X]) + g(W, [X, V ]) + g(X, [V, W ]).
It is called the Levi-Civita connection.

Démonstration. Let D be a connection satisfying these two points. In the right-hand side of the
Koszul formula, using the two points, we obtain 2g(DV W, X). This proves the uniqueness because
of the one-to-one correspondance between vector fields and 1-forms.
Let proves the existence. Let F (V, W, X) the right-hand side of the Koszul formula. Then if we
take two vector fields V and W , then the map F (V, W, ·) : X (M ) −→ R is F(M )-linear. So it is
a 1-form. Thus there exists a unique vector field DV W such that
g(DV W, X) = F (V, W, X), ∀X ∈ X (M ).
This show the Koszul formula and that the map D is a connection. With this formula, we can prove
the two points. ⋄

Notation. We will write ∇ for the Levi-Civita connection. With this notation and g(·, ·) = ⟨·, ·⟩,
the two points of the theorem are
X⟨Y, Z⟩ = ⟨∇X Z, Y ⟩ + ⟨∇X Y, Z⟩,
[X, Y ] = ∇X Y − ∇Y X.

Définition 2.8. The Christoffel symbols for the chart (U, xi ) are the functions on U given by
n
X
D∂i (∂j ) = Γki,j ∂k .
k=1

Recall that [∂i , ∂j ] = 0 = D∂i (∂j ) − D∂j (∂i ) by the Schwarz theorem, so Γki,j = Γkj,i . Moreover,
P j
if W = W ∂j on U , then
X X
D∂i (W ) = (∂i (W j )∂j + W j Γki,j ∂k )
j k
X X
= (∂i (W j ) + W j Γki,j )∂k
k j

By Koszul formula, we have


 
1 X k,ℓ ∂gℓ,j ∂gℓ,i ∂gi,j
Γki,j = g + − .
2 ∂xi ∂xj ∂xℓ

Exemple. On Rnν , we have Γki,j = 0.

2.3. Curvature and Ricci tensor


Definition-proposition 2.9. Let (M, g) a semi-riemannian manifold and ∇ its Levi-Civita connection.
The the map
X (M ) × X (M ) × X (M ) −→ X (M )
R:
(X, Y, Z) 7−→ R(X, Y )Z := ∇[X,Y ] Z − ∇X ∇Y Z + ∇Y ∇X Z
is a tensor field of type (1, 3), called the riemannian curvature.

There is a version of type (0, 4) given by


R(X, Y, Z, W ) = g(R(X, Y )Z, W ).
10 2.3. Curvature and Ricci tensor

Démonstration. We need to check that R(f X, Y )Z = f R(X, Y )Z and R(X, Y )(f Z) = f R(X, Y )Z.
We have [f X, Y ] = f XY − Y (f )X − f Y X and
∇[f X,Y ] Z = f ∇X ∇Y Z
which prove the formula. ⋄

Proposition 2.10. We have the following properties :


1. R(X, Y )Z = −R(X, Y )Z ;
2. g(R(X, Y )Z, W ) = −g(R(X, Y )W, Z) ;
3. R(X, Y )Z + R(Y, Z)X + R(Z, X)Y = 0 ;
4. g(R(X, Y )Z, W ) = g(R(Z, W )X, Y ).

Démonstration. 2. With g(·, ·) = ⟨·, ·⟩, one has


g(R(X, Y )Z, Z) = ⟨∇[X,Y ] Z − ∇X ∇Y Z + ∇Y ∇X Z, Z⟩
= ⟨∇[X,Y ] Z, Z⟩ − ⟨∇X ∇Y Z, Z⟩ + ⟨∇Y ∇X Z, Z⟩
 
⟨Z, Z⟩
= [X, Y ] − X⟨∇Y Z, Z⟩ + ⟨∇Y Z, ∇X Z⟩ + Y ⟨∇X Z, Z⟩ − ⟨∇X Z, ∇Y Z⟩
Z
      
⟨Z, Z⟩ ⟨Z, Z⟩ ⟨Z, Z⟩
= [X, Y ] − XY +Y X = 0.
Z Z Z
So g(R(X, Y )(Z + W ), Z + W ) = 0 and we conclure by bilinearity. ⋄

Remarque. The map R is a tensor. For X, Y, Z ∈ X (M ) and p ∈ M , the quantity (R(X, Y )Z)p
only depend on the values X(p), Y (p) and Z(p). So we can define Rp (u, v)w for u, v, w ∈ Tp M .
Proposition 2.11. Let X, Y and Z be three vector fields. Then
(∇Z R)(X, Y ) + (∇X R)(Y, Z) + (∇Y R)(Z, X) = 0.
Remarque. We have
(∇X R)(Y, Z)W = ∇X (R(Y, Z)W ) − R(∇X Y, Z)W − R(Y, ∇X Z)W − R(Y, Z)∇X W.
Moreover, we have
(∇X g)(Y, Z) = X(g(Y, Z)) − g(∇X Y, Z) − g(X, ∇X Z) = 0.

Démonstration. We prove the identity on a basis. We choose X = ∂i , Y = ∂j and Z = ∂k . So


(∇Z R)(X, Y )W = [∇Z , R(X, Y )]W − R(∇Z X, Y )W − R(X, ∇Z Y )W
and then
(∗) = (∇Z R)(X, Y ) + (∇X R)(Y, Z) + (∇Y R)(Z, X)
= [∇Z , R(X, Y )]W + [∇X , R(Y, Z)]W + [∇Y , R(Z, X)]
− R(∇Z X, Y )W − R(X, ∇Z Y )W
− R(∇X Y, Z)W − R(Y, ∇X Z)W
− R(∇Y Z, X)W − R(Z, ∇Y X)W
= [∇Z , R(X, Y )]W + [∇X , R(Y, Z)]W + [∇Y , R(Z, X)]
+ R([X, Z], Y )W + R([Z, Y ], X)W + R([Y, X], Z)W.
But
[∇Z , R(X, Y )] = [∇Z , [∇X , ∇Y ]] − [∇Z , ∇[X,Y ] ]
where the last term is null on U and so
(∗) = [∇Z , [∇X , ∇Y ]] = [∇X , [∇Y , ∇Z ]] + [∇Y , [∇Z , ∇Y ]] = 0. ⋄

Let p ∈ M be a point and Π be a plane in Tp M . For two tangent vectors v, w ∈ Tp M not null
and not colinear, we define
Q(v, w) := ⟨v, v⟩⟨w, w⟩ − ⟨v, w⟩2 .
Chapitre 2. Semi-riemannian manifolds 11

This gives a semi-riemannian metric. The plane Π is nondegenerate if


Q(v, w) ̸= 0, ∀v, w ∈ Π \ {0}, v ̸∝ w.
The quantity |Q(v, w)| is the volume of the parallelogram defined by the vectors v and w. If Π is
not degenerate, then we define the sectional curvature of Π bu
g(R(v, w)v, w)
K(Π) := .
Q(v, w)
This definition does not depend on the choice of the vectors v and w.
Proposition 2.12. If K(Π) = 0 for all plane Π ⊂ Tp M , then R = 0 at the point p.

Démonstration. 1. If v and w define a nondegenerate plane Π, then it suffices to apply the


implication K(Π) = 0 ⇒ ⟨R(v, w)v, w⟩ = 0.
2. If they define a degenerate plane, then v and w can be approximated by vectors which define
a nondegenerate plane. If v is null, let x be a tangent vector such that ⟨u, x⟩ ̸= 0. If not, let
x be the opposite of the causal type of v. Then Q(u, x) < 0. Let δ ̸= 0 a small real number
such that the vectors v and w + δx define a nondegenerate plane. We assume δ = 1. So thank
to the first case, we get
⟨R(u, w)u, x⟩ + ⟨R(u, x)u, w⟩ = 0
which implies ⟨R(u, w)u, x⟩ = 0 for all x and so R(u, w)u = 0. Thus R(v + x, w)(v + x) = 0
and R(v, w)x + R(x, w)v = 0, so R(u, w)x = R(w, x)u. If we do u ←→ w, we get R(w, u)x =
R(u, x)u. So we have R(u, w)x = R(w, x)u = R(x, v)w. But the first Bianchi identity gives
R(u, w)x + R(w, x)u + R(x, u)w = 0
and so R(u, w)x = 0. Thus R = 0.

Corollaire 2.13. If the sectional curvature of M is, at a point p, constant to c, then
∀x, y, z ∈ Tp M, R(x, y)z = c(⟨y, z⟩x − ⟨x, z⟩y).

Définition 2.14. – The Ricci tensor is a (0, 2)-tensor obtainned by contraction of


n
X
Ric(X, Y ) := εm g(R(X, Em ), Em )
m=1

where (Ei ) is a orthonormal frame (whe g(Ei , Ej ) = 0 if i ̸= j and g(Ei , Ei ) = εi = ±1). This
defines a symmetric form.
– The scalar curvature is the function
n
X
scal := εm Ric(Em , Em ).
m=1

Proposition 2.15. We have dscal = 2 div(Ric).

Démonstration. We set the notations.


– If f is a function, then df is a 1-form defined by df (X) := X(f ).
Pn
– If T is a (0, 2)-tensor, then div T is a 1-form defined by div T (X) := m=1 εm ∇Em T (Em , X).
– We work with a vector field X such that (∇Y X)p = 0 for Y ∈ X (M ).
– We work with an orthonormal basis (Em ) such that (∇Y Em )p = 0 for Y ∈ X (M ).
At p, we have
dscal = X(scal)
X
= εm εj X(g(R(Em , Ej )Ej , Em ))
m,j
X
= εm εj [g(∇X (R(Em , Ej )Ej ), Em ) + g(R(Em , Ej )Ej , ∇X Em )]
m,j
12 2.3. Curvature and Ricci tensor

X
= εm εj g((∇X R)(Em , Ej )Ej + R(∇X Em , Ej )Em + R(Em , ∇X Ei )Em + R(Em , Ej )∇X Em , Em )
m,j
X
= εm εj g((∇X R)(Em , Ej )Ej , Em )
m,j
X
= εm εj [−g((∇Ej R)(X, Em )Ej , Em ) − g((∇Em R)(Ej , X)Ej , Em )]
m,j
X
=− εm εj [(∇Em R)(Ej , X, Ej , Em ) + ∇Ej R(X, Em , Ej , Em )]
m,j
X
= εm εj [(∇Ej R)(Ej , Em , Em , X) + ∇EM R(Em , Ej , Ej , X)]
m,j
X
=2 εm εj (∇Ej R)(Ej , Em , Em , X)
m,j
X
=2 εm εj ∇Ej (R(Ej , Em , Em , X)) because ∇g = 0
m,j
X
=2 εm εj ∇Ej (Ric(Ej , X))
m,j
X
=2 εm εj ∇Ej (Ric(X, Ej ))
m,j
X
=2 εm εj (∇Ej Ric)(Ej , X)
m,j

= 2 div(Ric)X. ⋄

Definition-proposition 2.16. A semi-riemannian manifold (M, g) is an Einstein manifold if there


exists a function f on M such that
Ricp = f (p)gp , ∀p ∈ M. (∗)
If the dimension is greater than 3, then the function f is constant.

Démonstration. The idea is to take the divergence of the equation (∗). We have

X
div(f g)(X) = εi (∇Ei (f g))(Ei , X)
i
X
= εi [Ei (f g(Ei , X)) − f g(∇Ei Ei , X) − f g(Ei , ∇Ei X)]
i
X
= εi [Ei (f )g(X, Ei ) + f (Ei g(X, Ei )) − f g(∇Ei Ei , X) − f g(Ei , ∇Ei X)]
i
X
= εi Ei (f )g(X, Ei )
i
X 
= εi g(X, Ei )Ei (f )
i
= X(f ) = df (X)

and so div(f g) = df . Taking the divergence of the equation (∗), we obtain div(Ric) = df and, by
the last proposition, we have df = dscal/2. Taking its trace, we get scal = (dim M )f . Then we have
scal = 2f + K for a constant K. As dim M ⩾ 2, the function scal is a constant and so does the
function f . ⋄
Chapitre 2. Semi-riemannian manifolds 13

2.4. Killing vector field


Définition 2.17. A Killing vector field is a vector field such that the Lie derivative of the matrix g
with respect to X is zero, that is
L X g = 0.
The associated flow of a vector field is the map
M × I −→ M,
Ψ:
(p, t) 7−→ Ψt (p)
such that
dΨ(t, p)
Ψ(p, 0) = p et = X(p).
dt t=0

Proposition 2.18. A vector field X is a Killing vector field if and only if its flow is by isometries.

Démonstration. First, we show that


1 ∗
L X g = lim [Ψ g − g]. (1)
t−→0 t t
By the chain rule, one has
(L X g)(A, B) = X(g(A, B)) − g(L X A, B) − g(A, L X B)
= X(g(A, B)) − g([X, A], B) − g(A, [X, B])
= g(∇X A, B) + g(A, ∇X B) − g(∇X A − ∇A X, B) − g(A, ∇X B − ∇B X)
= g(∇A X, B) + g(∇B X, A).
So X is a Killing vector field if and only if
∀A, B, g(∇A X, B) = −g(A, ∇B X).
But we have
(Ψ∗t g − g)(A, B) = g(dΨt (A), dΨt (B)) − g(A, B)
= g(dΨt (A), dΨt (B)) − g(AΨt , BΨt ) + g(AΨt , BΨt ) − g(A, B).
On the one hand, with F = G ◦ α, α(t) = Ψt and G = g(A, B), we get
lim [g(AΨt , BΨt ) − g(A, B)] = F ′ (0)
t−→0
= Xg(A, B).
On the other hand, with à = AΨt and A ←→ AΨt , we have
g(dΨt (A), dΨt (B)) − g(AΨt , BΨt ) = g(Ã, B̃) − g(A, B)
= g(Ã − A, B̃) + g(A, B̃ − B)
with
1
lim g(Ã − A, B̃) = lim[g(dΨt (A) − AΨt , dΨt (B))]
t−→0 tt−→0
1
= lim g(dΨt (A − dΨ−t (AΨ(t) )), dΨt (B))
t−→0 t
1
= − lim g(dΨt (dΨ−t (AΨ(t) ) − A), dΨt (B))
t−→0 t
1
= −g( lim [dΨt (dΨ−t (AΨ(t) ) − A)], lim dΨt (BΨt )).
t−→0 t t−→0

But limt−→0 dΨt (BΨt ) = B and


1
lim [dΨt (dΨ−t (AΨ(t) ) − A)] = −[A, X]. (∗)
t−→0t
If we conclude the equality (∗), then we will get the formula (1). Let proove the equality (∗).
14 2.4. Killing vector field

Let Ψ be the flow of a vector field V . We must prove


1
[V, W ] = lim [dΨ−t (WΨt − W )] (2)
t−→0 t

Let Fp (t) = dΨ−t (WΨt (p) . The right-hand side of the equation (2) is exactly Fp′ (0). Assume Vp ̸= 0.
Let (xi ) a system of local coordinates such that ∂/∂xi = V . Then locally
x1 (Ψt (q)) = x1 (q) + t et xj (Ψt (q)) = xj (q), j⩾2
and 

 X ∂Ψj ∂ ∂
dΨt = = (Ψt ).
∂xi j
∂xi ∂xj ∂x i


W i ∂x
P
Let W = i i
. Thus
X ∂
Fp (t) = W i (Ψt (p)) .
i
∂xi

and so
X d(W i ◦ Ψt (p)) ∂
Fp′ (0) =
i
dt t=0 ∂xi
X ∂
= Vp (W i )
i
∂xi
X ∂W i
= ∂i .
i
∂x1
So
[V, W ] = (∂1 , W )
X
= [∂1 , W i ∂i ]
X
= ∂1 (W 1 )∂i + W i [∂1 , ∂i ]
X ∂X i
= ∂i = Fp′ (0). ⋄
∂x1
Lemme 2.19. Let X be a Killing vector field on a connected manifold M such that, for all
point p ∈ M , we have
X(p) = 0 et (∇X)p = 0.
Then X = 0 on M .

Démonstration. The set


A := {q ∈ M | X(q) = 0 et (∇X)q = 0}
is closed and nonempty. To conclude, we show that this set is open. Take p ∈ A et Ψt the associated
flow.
Since X(p) = 0, we have Ψt (p) = 0 for all t. Indeed, the flow satisfies Ψt ◦ Ψs = Ψt+s . Thus we
get
dΨt+s dΨt
(p) = (p) = X(p) = 0
dt t=0 dt t=0
which concludes Ψt (p) = Ψ0 (p) = p.
Let proves that dΨt : Tp M −→ Tp M is the identity. We have [X, Y ]p = (∇X Y )p − (∇Y X)p . But
the points (∇X Y )p depends only on X(p) = 0, so (∇X Y )p = 0. So we have (X(Y ) − Y (X))p = 0
and
dΨt (YΨt ) − Yp
= 0.
t
Let Fp (t) := dΨt (YΨt ). Then Fp′ (0) = 0. With Y = dΨs (u), the function t 7−→ dΨ−t (YΨt ) has a
null derivative, so dΨt = 0. Finally, the flow acts by isométries, so the flow acts by the identity. ⋄
Chapitre 2. Semi-riemannian manifolds 15

Remarque. Klling vector fields form a Lie algebra of finite dimension because [L X , L Y ] = L [X,Y ] .
Moreover, the dimension is less then n(n + 1)/2 where n = dim M .

Proposition 2.20. Let X be a Killing vector field and f := |X|2 /2. Then
∆f = − Ric(X, X) + |∇X|2 .

Démonstration. Only calculus. ⋄

Remarque. We compute grad f = −∇X X.


Définition 2.21. The divergence of a vector field X is
div V := tr(∇V ).
Théorème 2.22 (Bochores). Let M be a compact riemannian manifold with Ric(X, X) ⩽ 0 for
all X. Then a Killing vector field is parallel, that is ∇X = 0 on M . If Ric < 0, then there are no
nonzero Killing vector field.

Démonstration. But the Stokes theorem, we have


Z
div V = 0.
M

Let f = |X|2 /2. We get Z Z


∇f = 0 = − Ric(Y, X) + |∇X|2
M M

and so ∇X = 0. Moreover, if Ric < 0, then Ric(Y, X) = 0 for all Y and so X = 0. ⋄

Théorème 2.23 (Berger). Let M a compact riemannian manifold of even dimension with positive
sectional curvature. The any Killing vector field has a zero.

Démonstration. Let X be a vector field. Let f := |X|2 /2. Then grad f = −∇X X. If X has no
zero, then f has a positive minimum at a point p ∈ M . Then Hess f (p) ⩾ 0. Let V be a vector field.
Ten
Hess f (V, V ) := ⟨∇V (∇f ), V ⟩ = ⟨−∇V ∇X X, V ⟩
= ⟨R(V, X)X, V ⟩ + ⟨∇V X, ∇V X⟩.
But B : X 7−→ ∇X X is skew symmetric, so (∇X X)(p) = (grad f )p = 0, so B admits λ = 0 as
an eigenvalue with X(p) as a eigenvector. As dim M is even, there exists another eigenvector V
corresponding to λ = 0. ⋄
17

Chapitre 3
Geodesics

3.1 First definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

3.1. First definitions


Définition 3.1. Let M be a differential manifold. Let I := ]−ε, ε[ ⊂ R be a interval centered at the
origin and γ : I −→ M a curve. A vector field along the curve γ is a map X : I −→ TM such that
∀t ∈ I, X(t) ∈ Tγ(t) M.

Exemple. The map t 7−→ (γ(t), γ ′ (t)) is a vector field along the curve γ.
Proposition 3.2. Let M a semi-riemannian manifold et γ : I −→ M a curve. Then there exists a
unique R-linear operator
D
: {vector fields along γ} −→ {vector fields along γ}
dt
such that
D df D
– dt (f X) = dt X + f dt X;
D
– If X(t) = Y (γ(t)), then dt X = (∇γ Y ) ◦ γ.

Démonstration. Let t0 ∈ I. Let (U, x) a chart on M and J ⊂ I an interval such that γ(J) ⊂ U .
Let Xi := ∂/∂xi . If Y is a vector field along γ, we have
X
Tγ(t) M ∋ Y (t) = αj (t)(Xj )γ(t).
j

With the first two conditions, we get


D X D X
Y = αj (Xi ◦ γ) + αk′ Xk (γ)
dt j
dt
k

and, by the third condition, we obtain


X
γ̇t = γ̇i Xi (◦γ)
and
D X
(Xi ◦ γ) = (∇j Xj ) ◦ γ = γ̇i (∇Xi Xj ) ◦ γ.
dt i

Put everything together


D X X 
Y = αk′ Γki,j,◦γ γi′ αj Xk ◦ γ.
dt i,j
k

Therefore the operator exists and is unique. ⋄

Remarque. The quantity (∇γ ′ X)(t) depends only on γ̇(t). We denote D


dt Y by ∇γ̇ Y .
18 3.1. First definitions

Définition 3.3. Let M be a semi-riemannian manifold and γ : I −→ M a curve of class C ∞ . A


vector field X along the curve γ is parallel if ∇γ̇ X = 0.
Définition 3.4. A curve γ is a geodesic if ∇γ̇ γ̇ = 0.

Théorème 3.5. Let M be a semi-riemannian manifold and γ : ]a, b[ −→ M be a curve. Let t0 ∈ I


be a real number and X0 ∈ Tγ(t0 ) M a tangent vector. Then there exists a unique vector field Y
along the curve γ such that Y (t0 ) = X0 .

Démonstration. Let (U, x) be a chart such that γ(t0 ) ∈ U . Let Xi := ∂/∂xi . Let J ⊂ I be a
interval such that γ(J) ⊂ U . We denote
X X
γ̇(t) = γ̇ i (t)Xi (γ(t)) et Y (t) = αj (t)Xj (γ(t)).
j

Then
DY Xh X i
(t) = α̇k (t) + αj (t)γ̇i (t)Γki,j (γ(t)) Xk (γ(t))
dt i,j
k

and so
DY X
(t) = 0 ⇐⇒ ∀k, α̇k (t) + αj (t)γ̇i (t)Γki,j (γ(t)) = 0. (∗)
dt i,j

Let admits the Picard-Lindelöf-Cauchy theorem :


Let f : I × U −→ Rn a continuous function which is Lipschitz in x. Then there exists
an unique solution x : I −→ Rn of the system
x′ (t) = g(t, x(t)) et x(t0 ) = x0 .
So there exists a solution to the equation (∗) for any initial data. One can extend Y (t) to I because
the coefficients in the equation (∗) are bounded for t ∈ I. ⋄

Lemme 3.6. Let X and Y be two parallel vector field along a curve γ. The the map
t 7−→ gγ(t) (X(t), Y (t))
a constant. For X = Y = γ̇, if γ is a geodesic, then g(γ̇, γ̇) = |γ|2 is constant.
Remarque. So causal type of geodesics is preserve on frame (Xi ).
Théorème 3.7. Let M be a semi-riemannian manifold. Let p ∈ M and v ∈ Tp M . Then there exists
an open interval I and a unique geodesic γ : I −→ M such that
γ(0) = p et γ̇(0) = v.

Démonstration. Let (U, x) be a chart such that γ(t0 ) ∈ U . Let Xi := ∂/∂xi . Let J ⊂ I be a
interval such that γ(J) ⊂ U . We write
X
γ̇ = γ̇i (Xi ◦ γ).
i

We have Xh X i
∇γ̇ γ̇ γ̈k (t) + γ̇j (t)γ̇j (t)Γki,j ◦ γ Xk (γ(t)).
k i,j

So γ is a geodesic if and only if


X
γ̈k (t) + γ̇j (t)γ̇j (t)Γki,j ◦ γ, ∀k
i,j

if on only if its components satisfy the systems of second order nonlinear ordinary differential
equation. Existence is given, for any initial data p and v, by the Picard-Lindelöf-Cauchy theorem. ⋄
19

Chapitre 4
Examples

Exemple. The euclidean space Rn is a semi-riemannian manifold. The geodesics are straight lines.
Indeed, we have Γki,j = 0 and a path γ must verify the equation
γ̈ k + Γki,j γ̇ i γ̇j = 0

Exemple. The sphere Sn is a riemannian manifold. Indeed, it is a differentiable manifolds by the


charts
Sn \ {N } −→ Rn ,
πN :
 
x1 xn
(x1 , . . . , xn ) 7−→ ,...,
1 − xn+1 1 − xn+1
and
Sn \ {S} −→ Rn ,
πS :
 
x1 xn
(x1 , . . . , xn ) 7−→ ,...,
1 + xn+1 1 + xn+1
where the points N and S are the north and south poles. These two charts are bijective and we can
verify that there compositions are C ∞ maps.
We find the tangent spaces. Let p ∈ Sn . Take a curve γ : ]−ε, ε[ −→ Sn with γ(0) = p. Then we
have |γ(0)|2 = 1 and thus γ̇(0) ∈ Tp Sn . We can prove Tp Sn = {X ∈ Rn+1 | ⟨p, X⟩ = 0}.
We must equip the sphere with a metric. For X, Y ∈ Tp Sn , we set
gSn ,p (X, Y ) := ⟨X, Y ⟩Rn+1 .
Then the tensor g is a metric on the sphere Sn . We get a riemannian manifold.
We must understand the Levi-Civita connection. We define the connection ∇ on Sn by
∇X Y := (∂X Y )tangent
and we will check that it is indeed the Levi-Civita connection. Here, the « tangent » is the projection
on the tangent space according the decomposition Rn+1 = Rp⊕Tp Sn and we denote ∂X Y = dY (X).
First, we prove that
∇X Y = ∂X Y + ⟨X, Y ⟩p.
The normal part of ∂X Y is ⟨∂X Y, p⟩p. But ⟨Y, p⟩ = 0, so X⟨Y, p⟩ = 0 and ⟨∂X Y, p⟩ + ⟨Y, ∂Y p⟩ = 0
and ∂X p = dp(X) = X. So the normal part of ∂X Y is −⟨X, Y ⟩p. Next, we observe that
⟨Z, ∇X Y ⟩ = ⟨Z, ∂X Y ⟩.
By the Koszul formula, we have
2⟨Z, ∂X Y ⟩ = X⟨Z, Y ⟩ − Z⟨X, Y ⟩
and
⟨Z, ∂X Y ⟩ = ⟨Z, ∇X Y ⟩ + Y ⟨X, Y ⟩ − ⟨X, [Y, Z]⟩ + ⟨Y, [Z, X]⟩ + ⟨Z, [X, Y ]⟩.
So we get the Koszul formula. By the uniqueness, this is the Levi-Civita connection.
20

Let us find the curvature. We have


−R(X, Y )Z = ∇X ∇Y Z − ∇Y ∇X Z − ∇[X,Y ] Z
= ∇X (∂Y Z + ⟨Z, Y ⟩p) − ∇Y (∂X Z + ⟨Z, X⟩p) − (∇[X,Y ] Z + ⟨Z, [X, Y ]⟩p)
= ∂X ∂Y z − ∂Y ∂X Z − ∂[X,Y ] z) + (⟨X, ∂Y Z⟩p − ⟨Y, ∂X Z⟩p − ⟨[X, Y ], Z⟩p) + (∂X ⟨Y, Z⟩p − ⟨X, ⟨Y, Z⟩p⟩p −
= ⟨Y, ∂Y Z⟩p − ⟨Y, ∂X Z⟩p − ⟨[X, Y ], Z⟩p + ∂X (⟨Y, Z⟩p) − ∂Y (⟨X, Z⟩p).
But ∂X (⟨Y, Z⟩p) = ⟨∇X Y, Z⟩p + ⟨Y, ∇X Z⟩p + ⟨Y, Z⟩p and so
−R(X, Y )Z = ⟨∇X Y, Z⟩p − ⟨∇Y X, Z⟩p − ⟨[X, Y ], Z⟩p + ⟨Y, Z⟩p − ⟨X, Z⟩p
= ⟨Y, Z⟩X − ⟨X, Z⟩Y.
The sectional curvature is K = +1. The Ricci tensor is
Ric(X, Y ) = (n − 1)⟨X, Y ⟩
and the scalar curvature is
scal = n(n − 1).
Let us find the geodesics. Let γ a geodesics. Then ∇γ̇ γ̇ = 0. But
∇γ̇ γ̇ = (∂γ γ̇)
= γ̈ tangent
= γ̈ − γ̈ normal
= γ̈ − ⟨γ̈, γ⟩γ.
After calculus, we find that the geodesics are great circles.

Exemple. The hyperbolic space is Hm := R∗+ × Rm−1 . Its tangent spaces are Tp Hm ≃ Rm . We
equip this manifold with the metric
⟨X, Y ⟩
g(X, Y ) = .
x21
It is a riemannian manifold with sectional curvature equal to −1.
We can choose others models of the hyperbolic space such as
Hm = {(x0 , . . . , xm ) ∈ Rm+1 | x0 > 0, −x20 + x21 + · · · + x2m = −1}.
Equipped with the induce metric, it is a riemannian manifold. An other model is the Poincaré
model
Dm := {x ∈ Rm | |x| < 1}
with the metric
4
g(X, Y ) = ⟨X, Y ⟩.
(1 − |x|2 )2
The sectional curvature is also equal to −1.

Exemple. The curvature of Rm


1 is zero, its geodesics are straight lines.

Exemple. We set the pseudo-sphere Sn−1


ν ⊂ Rnν . The tangent space is
Tp Sn−1
ν = {X ∈ Rn | ⟨p, X⟩Rnν = 0}.
The pseudo-spere equipped with the metric ⟨·, ·⟩Rnν is a riemannian manifold of signature (ν, n−1−ν).
It is diffeomorphic to Rν × Sn−1−ν and it sectional curvature is +1. The geodesics are branches
of hyperboloids, straight line or periodic curves on ellipsoids : we can prove this by considering
different cases (the vectors to join are time, space or light like). More over, a curve γ is a geodesic
if and only iff the curves γ̈ and γ are parallel.
21

Chapitre 5
Calculus of variations

Let M be a semi-Riemannian manifold. Let γ : I −→ M be a curve. We recall that the set Aγ


is the set of maps Y : I −→ TM along the curve γ, that is such that
∀t ∈ I, Y (t) ∈ Tγ(t) M.
A such map Y can be write X
Y (t) = αj (t)(Xj ◦ γ)(t)
j

on a chart (U, xi ) with Xj := ∂/∂xj . The derivation for Aγ is


D X X 
Y (t) = α̇k (t) + Γkij (γ(t))γ̇i (t)αj (t) Xk (γ(t)).
dt i,j
k

Facts.
D
1. For all X0 ∈ Tγ(0) M , there exists Y ∈ Aγ such that Y (0) = X0 and dt Y = 0.
D
2. For all X0 ∈ Ta M , there exists t0 > 0 and γ : [0, t0 [ −→ M such that γ(a) = X0 and dt γ = 0.
Such a γ is called a geodesic.
3. We also write dtD
Y = Y ′ = Ẏ = ∇γ̇ Y .
4. If γ is a geodesic, then
d
⟨γ̇, γ̇⟩ = ⟨γ̈, γ̇⟩ = 0.
dt
Définition 5.1. Let M be a semi-Riemannian manifold. A variation of a function α : [a, b] −→ M
of class C ∞ is a map x : [a, b] × ]−δ, δ[ −→ M of class C ∞ with δ > 0 such that x(u, 0) = α(u).
The variation vector field is the vector field V such that
∂x
V (u) := (u, 0).
∂v
The length of α is
Z b
L(α) := |α′ (s)| ds
a
p
where |·| = |⟨·, ·⟩|. The length of V is
Z b
∂x
L(v) = Lx (v) = (s, v) ds.
a ∂u
We consider curves such that |γ ′ (t)| > 0, called regular curves of space-like. We denote ε the
sign of ⟨α′ , α′ ⟩.
Lemme 5.2. If x is a variation of α with |α′ | > 0, then
Z b ′
α (u)
L′x (0) = ε ⟨ ′ , V ′ (u)⟩ du.
a |α (u)|
22

∂x
Démonstration. With xu = ∂u , we have
Z b
L(u) := |xu (u, v)| du.
a
We have α′ = xu (u, 0)′ . So for δ small enough, we have |xu (u, v)| > 0 for u ∈ ]−δ, δ[. So
Z b
′ d
L (0) = |xu | dt.
a du u=0
But we get
d 1 ε⟨xu , xuv ⟩
|xu | = (ε⟨xu , xu ⟩)−1/2 2ε⟨xu , xuv ⟩ = .
du 2 ⟨xu , xu ⟩
Take u = 0, we get xu (u, 0) = α′ (0) and xv (u, 0) = V (u) and xuv (u, 0) = V ′ (u). ⋄

Proposition 5.3 (first variation). Let α : [a, b] −→ M be a continuous and smooth curve piece-wise
of constant speed c > 0 and of sign ε. Let x be a variation of α. Then
n
ε b ′′
Z
εX ε
L′ (0) = − ⟨α , V ⟩ du − ⟨∆α′ (Ui ), V (Ui )⟩ + ⟨α′ , V ⟩|ba
c a c i=1 c
with U1 < · · · < Uk are points where α is not C ∞ and
∆α′ (Ui ) = α′ (Ui+ ) − α′ (Ui− ) ∈ Tα(Ui ) M.

Démonstration. We have
α′ 1
⟨ ′
, V ⟩ = ⟨α′ , V ′ ⟩.
|α | c
On ]Ui , Ui+1 [, we have
d ′
⟨α′ , V ′ ⟩ = ⟨α , V ⟩ − ⟨α′′ , V ⟩.
du
So Z Ui+1 Z Ui+1
U
⟨α′ , V ′ ⟩ du = ⟨α′ , V ⟩Ui+1
i
− ⟨α′′ , V ⟩ du.
Ui Ui
We sum up to obtain the desired formula. ⋄

Corollaire 5.4. A piece-wise smooth curve α with constant speed c > 0 is a geodesic if and only if
the first variation of L is zero for any variation with fixed ends.
Remarque. Fixed ends imply that V is zero at a and b and
ε ′
⟨α , V ⟩|ba = 0.
2
Démonstration. Suppose that α is a geodesic, that is α′′ = 0. Then α is smooth, so ∆α′ (Ui ) = 0.
In particular, we get V (a) = V (b) = 0 and so L′ (0) = 0.
Suppose that L′ (0) = 0. First we show that α is a geodesic on ]Ui , Ui+1 [, that is α′′ (t) = 0
for t ∈ ]Ui , Ui+1 [. Let y be in Tα(t) M and f a smooth function defined on [a, b] with supp f ⊂
[t − δ, t + δ] ⊂ ]Ui , Ui+1 [ and f ∈ [0, 1] and f = 1 on ]t − δ/2, t + δ/2[. Let Y be the vector field
D
obtained by parallel transport of y along α, that is dt Y = 0 and Y (t) = 0. Let V := f Y . Then
V (a) = V (b) = 0. Let exp be the exponential map, that is the map
expp : D ⊂ Tp M −→ M
with p ∈ M where
expp (v) = B(1)
where B is the geodesic starting at p with initial speed v and where
D = {v ∈ Tp M | B(1) exists}.
Let x(u, v) = expα(u) (vV (u)). Then x(u, v) is a variation of α with fixed ends. So L′ (0) = 0 and
then
Z b
0= ⟨α′′ , v⟩ du
a
Chapitre 5. Calculus of variations 23

Z t+δ
= ⟨α′′ , f Y ⟩.
t−δ

This implies that


∀y ∈ Tγ(t) M, ⟨α′′ (t), y⟩ = 0
and so α′′ (t) = 0 on each ]Ui , Ui+1 [.
If y ∈ Tα(Ui ) M , let f have its support in ]Ui−1 , Ui+1 [ with f = 1 around Ui . So
ε
0 = L′ (0) = − ⟨∆α′ (Ui ), y⟩, ∀y
c
and so
∆α′ (Ui ) = 0. ⋄

We will compute L′′ (0) if L′ (0) = 0. Any vector field Y along α decomposes as Y = Y T + Y ⊥
where Y T = ε⟨Y, α′ ⟩α′ =: f α′ and Y ⊥ is orthogonal to α′ . If α is a geodesic, then
Y ′ = f ′ α′ + (Y ⊥ )′ .
Moreover, we have (Y ′ )⊥ = (Y ⊥ )′ .
Théorème 5.5 (second variation). Let γ be a geodesic of constant speed c > 0 and of sign ε. If x
is a variation of γ, then
ε b ′⊥ ′⊥
Z
ε
L′′ (0) = ⟨V , V ⟩ − ⟨R(V, γ ′ )V, γ ′ ⟩ du + ⟨γ ′ , A⟩|ba
c a c
where V (u) = xv (u, 0) and A(u) = xvv (u, 0).

Let Ω(p, q) be the space of smooth piece-wise curves from [a, b] to M starting at p and ending
at q. The tangent space to Ω(p, q) at α is the set Tα Ω(p, q) of vector fields V along α with
V (a) = V (b) = 0. The index of σ ∈ Ω(p, q) is the bilinear symmetric form
Iσ : Tσ Ω −→ Tσ Ω
such that Iσ (V, V ) = Lx (σ) where x is a variation with fixed ends and variation vector V , that is
ε b ′⊥ ′⊥
Z
Iσ (V, W ) = ⟨V , V ⟩ − ⟨R(V, σ ′ )W, σ ′ ⟩ du.
c a
We have Iσ (V, W ) = Iσ (V ⊥ , W ⊥ ).
Lemme 5.6. Let σ be a non-null geodesic with sign ε. Let M be a semi-Riemannian manifold with
index ν. Then
1. if Iσ is semi-definite positive, then ν = 0 or n ;
2. if Iσ is semi-definite negative, then ν = 1 or n − 1.
Définition 5.7. Let γ be a geodesic. A vector field Y along γ is called Jacobi field if
Y ′′ = R(Y, γ ′ )γ ′ .
If x is a variation of γ such that
∀v, u 7−→ x(u, v) is a geodesic,
then the variation vector u 7−→ ∂x∂v (u, 0) is a Jacobi field.
For all v, w ∈ Tp M , there exists an unique Jacobi field Y along γ such that Y (0) = v and
Y ′ (0) = w.

Définition 5.8. Two points σ(a) and σ(b) with a ̸= b on a geodesic σ are conjugate if there exists
a nontrivial Jacobi field Y such that J(a) = J(b) = 0.

Then σ(a) and σ(b) are conjugate if and only if there exists a variation x of σ such that the map
∂x
u 7−→ x(u, v) is a geodesic, for all v, started from σ(a) such that ∂u (b, 0) = 0. This is equivalent
to the fact that the exponential map expp : Tp M −→ M is singular at bσ ′ (0), that is there is a
tangent vector x to p at bσ ′ (0) such that d(expp )bσ′ (0) (x) = 0.
24

Lemme 5.9. Let σ be a geodesic such that σ ⊥ (s) ∈ Tσ(s) M is space-like. If ⟨R(v, σ ′ )v, σ ′ ⟩ ⩽ 0 for
all v ⊥ σ ′ , then there is no conjugate points along σ.

You might also like