Differential Geometry - Christian Bär
Differential Geometry - Christian Bär
Differential Geometry - Christian Bär
GEOMETRY IN POTSDAM
Differential Geometry
Summer Term 2013
Preface iii
1 Manifolds 1
1.1 Topological manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Differentiable manifolds . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.3 Tangent vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.4 Directional derivatives and derivations . . . . . . . . . . . . . . . . . . . 23
1.5 Vector fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2 Semi-Riemannian Geometry 35
2.1 Bilinear forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.2 Semi-Riemannian metrics . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.3 Differentiation of vector fields . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.4 Vector fields along maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.5 Parallel transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.6 Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
3 Curvature 79
3.1 The Riemannian curvature tensor . . . . . . . . . . . . . . . . . . . . . . . 79
3.2 Sectional curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
3.3 Ricci- and scalar curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
3.4 Jacobi fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
4 Submanifolds 107
4.1 Submanifold of differentiable manifolds . . . . . . . . . . . . . . . . . . . 107
4.2 Semi-Riemannian submanifolds . . . . . . . . . . . . . . . . . . . . . . . . 113
4.3 Totally geodesic submanifolds . . . . . . . . . . . . . . . . . . . . . . . . . 120
4.4 Hypersurfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123
4.5 Trigonometry in spaces of constant curvature . . . . . . . . . . . . . . . . 133
Literature 163
i
Contents
Index 165
ii
Preface
These are the lecture notes of an introductory course on differential geometry that I
gave in 2013. It introduces the mathematical concepts necessary to describe and ana-
lyze curved spaces of arbitrary dimension. Important concepts are manifolds, vector
fields, semi-Riemannian metrics, curvature, geodesics, Jacobi fields and much more.
The focus is on Riemannian geometry but, as we move along, we also treat more gen-
eral semi-Riemannian geometry such as Lorentzian geometry which is central for ap-
plications in General Relativity. We also make a connection to classical geometry when
we apply differential geometry to derive the laws of trigonometry on spaces of constant
curvature. One fundamental result of Riemannian geometry that we show towards the
end of the course is the Bonnet-Myers theorem. It roughly states that the larger the
curvature of a space, the smaller the space itself must be.
The lecture course did not require prior attendance of a course on elementary differ-
ential geometry treating curves and surfaces but such a course would certainly help to
develop the right intuition.
It is my pleasure to thank all those who helped to improve the manuscript by sugges-
tions, corrections or by work on the LATEX code. My particular thanks go to Andrea
Röser who wrote the first version in German language and created many pictures in
wonderful quality, to Volker Branding who translated the manuscript into English and
to Ramona Ziese who improved the layout.
Christian Bär
iii
1 Manifolds
1.1 Topological manifolds
3. If U1 , U2 ∈ O, then also U1 ∩ U2 ∈ O.
The pair (M, O) is called a topological space. By abuse of language, one often speaks
about the topological space M rather than (M, O).
If both (M, OM ) and (N, ON ) are topological spaces, a map f : M → N is called con-
tinuous, if
f −1 (V ) ∈ OM for all V ∈ ON .
In other words, preimages of open sets have to be open. A bijective continuous map
f : M → N , whose inverse f −1 is also continuous, is called a homeomorphism. Two
topological spaces M and N are called homeomorphic, if there exists a homeomor-
phism between them.
1. M is Hausdorff, that is, for all p, q ∈ M with p 6= q there exist open sets U, V ⊂
M with p ∈ U , q ∈ V and U ∩ V = ∅.
V
U b q
b p
M
1
1 Manifolds
2. The topology of M has a countable basis, that is, there exists a countable subset
B ⊂ O, such that for every U ∈ O there are Bi ∈ B, i ∈ I with
[
U= Bi .
i∈I
≈
x V ⊂ Rn
U
b p
M
Remark 1.1.3. The first two conditions in the definition are more of a technical nature
and are sometimes neglected. The important fact is that a topological manifold is locally
homeomorphic to Rn . Loosely speaking manifolds look locally like Euclidean space. If
the topology on M is induced by a metric, then the first condition is satisfied automat-
ically. If M is given as a subset of RN with the subset topology, then both conditions 1
and 2 are satisfied automatically.
2
1.1 Topological manifolds
(iii) We construct two homeomorphisms with the help of the stereographic projec-
tion.
We define U1 := S n \ {SP } with
SP := (−1, 0, . . . , 0) ∈ Rn+1 and set
x(y) N P b
V1 := Rn . Furthermore, we define b
x : U1 −→ V1 , b y
2
y = (y 0 , y 1 , . . . , y n ) 7−→ x(y) = · ŷ.
1 + y0
b
| {z } SP
=:ŷ
y : V1 −→ U1 ,
1
x 7−→ y(x) = (4 − kxk2 , 4x),
4 + kxk2
x̃ : U2 −→ V2 ,
2
y 7−→ x̃(y) = x̃(y 0 , y 1 , . . . , y n ) = · ŷ.
| {z } 1 − y0
=:ŷ
We have seen that the sphere S n is an n-dimensional topological manifold.
V ⊂ R2 and a homeomorphism x : U → q2 b
x(q2 )
V with x(0) = 0. V
3
1 Manifolds
c(0) = x(q1 ), c(1) = x(q2 ) and c(t) 6= x(0) for all t ∈ [0, 1].
c̃(0) = q1 , c̃(1) = q2 ,
that is, we have c̃1 (0) > 0 while c̃1 (1) < 0. Applying the mean value theorem we
find, that there exists a t ∈ (0, 1) with c̃1 (t) = 0. Then c̃(t) = (0, 0, 0) and conse-
quently c(t) = x(c̃(t)) = x(0), which contradicts the choice of c. Hence, M is not a
2-dimensional topological manifold.
≈
U x V ⊂ Rn
Thus we get:
Proposition 1.1.6
A topological space M is a 0-dimensional topological manifold, if and only if M is countable
and carries the discrete topology.
4
1.1 Topological manifolds
Definition 1.1.7. We call a topological manifold M connected, if for every two points
p, q ∈ M there exists a continuous map c : [0, 1] → M with c(0) = p and c(1) = q.
Given two points, there has to be a continuous curve in M which connects both. Usu-
ally, in Topology one calls this path-connected, which is in the case of manifolds equiva-
lent to being connected. We do not want to go deeper into this subject at this point.
Proposition 1.1.9
Every connected 1-dimensional topological manifold is homeomorphic to R or to S 1 .
A proof of this fact can be found in the appendix of [M65]. Thus, the only compact,
connected topological manifold of dimension 1 is S 1 .
Theorem 1.1.10
Let M and A be sets. For all α ∈ A assume that Uα ⊂ M and Vα ⊂ Rn are subsets and that
xα : Uα → Vα are bijective maps. Suppose the following holds:
[
(i) Uα = M ,
α∈A
Then M carries a unique topology for which all Uα are open sets and all xα are homeomor-
phisms.
5
1 Manifolds
Uβ
xβ
xα
Vα Uα M Vβ
xβ ◦ xα −1
Now we have to check that this O is a topology and that it has the desired properties:
6
1.1 Topological manifolds
Thus xβ −1 (Y ) ∈ O.
We will use Theorem 1.1.10 to equip RPn with the structure of an n-dimensional topo-
logical manifold. We set
Since α is affine-linear there exist a B ∈ Mat(n × (n + 1), R) and a c ∈ Rn+1 such that
α(x) = Bx + c
α L
0 b c b
bc
α(Rn )
Rn
Rn+1
b
0
7
1 Manifolds
For L ∈ Uα the space L ∩ α(Rn ) consists of exactly one point, because otherwise we
would have L ⊂ α(Rn ) and hence 0 ∈ α(Rn ), a contradiction. Moreover, we have
xα −1 (v) = R · α(v).
Then
n
\
L⊂ ej ⊥ = {0}.
j=0 ej ⊥
n
S S
This is a contradiction, consequently Uαj = RPn and hence Uα = RPn .
j=0 α∈A
8
1.1 Topological manifolds
α(Rn )
α α(v) bc
xβ ◦ xα −1 (v)
b
Vα
b
0
v b
Vβ
bc
β
β(Rn )
9
1 Manifolds
Example 1.1.11 continued. For RPn the finite set A1 := {α0 , . . . , αn } does the job. Con-
sequently, the topology of RPn has a countable basis.
Proof of Proposition 1.1.12. The topology resulting from A has all the properties of the
topology resulting from A1 . Since the topology is unique, A and A1 give the same
topology on M .
Without loss of generality we may therefore assume that A1 = A is countable. Now the
topology of each Vα ⊂ Rn has aScountable basis Bα . Then xα −1 (Bα ) is a countable basis
of the topogoly of Uα . Finally, α∈A xα −1 (Bα ) is a countable basis of M .
xα
b p q b
Uα
M
V1 b b
V2
R n ⊃ Vα
We summarize:
10
1.2 Differentiable manifolds
Corollary 1.1.14
Let M and A be sets and let A1 ⊂ A be a countable subset. For all α ∈ A assume that
Uα ⊂ M and Vα ⊂ Rn are subsets and that xα : Uα → Vα are bijective maps. Suppose the
following holds:
[
(i) Uα = M ;
α∈A1
Then M carries a unique topology which turns M into an n-dimensional topological manifold
such that the xα : Uα → Vα are charts.
x V
p b
x(p)
U b
M
f
R f ◦ x−1
11
1 Manifolds
f ◦ y −1 = (f ◦ x−1 ) ◦ (x ◦ y −1 ) .
| {z } | {z }
diff’able only
at x(p) continuous
This concept of differentiability depends on the choice of chart x and this is really bad
because on a general topological manifold there are no preferred coordinate systems.
The sad truth is that there is no reasonable concept of differentiable functions on a
topological manifold.
But there is one thing we can do, we can refine the notion of a manifold. If x ◦ y −1 were
a diffeomorphism and not only a homeomorphism, then the differentiability of f ◦ x−1
would imply the differentiability of f ◦ y −1 . We enforce this by making the following
definition.
is a C ∞ -diffeomorphism.
Ũ
y
x
V U M Ṽ
x(U ∩ Ũ ) y(U ∩ Ũ )
y ◦ x−1
12
1.2 Differentiable manifolds
Example 1.2.3
(1) Let M = U ⊂ Rn be open. Then A := {id : U → U } is a C ∞ -atlas.
(2) Let M = S n and A := {(x : U1 → V1 ), (x̃ : U2 → V2 )}, where U1 := S n \ {SP },
U2 := S n \ {N P } and V1 := V2 := Rn , compare Example 1.1.4.3. Furthermore, let
2 0
x(y) = ŷ, where y = y , ŷ ∈ Rn+1 ,
1 + y0
1 2
y(x) = 4 − kxk , 4x and
4 + kxk2
2
x̃(y) = ŷ.
1 − y0
Then we have for v ∈ x(U1 ∩ U2 ) = x(S n \ {SP, N P }) = Rn \ {0}:
−1 4 − kvk2 4v
x̃ ◦ x (v) = x̃ ,
4 + kvk2 4 + kvk2
2 4v
= 2
4 − kvk 4 + kvk2
1−
4 + kvk2
8v
=
4 + kvk − 4 + kvk2
2
4v
= .
kvk2
Hence x̃ ◦ x−1 is C ∞ on Rn \ {0} = x(S n \ {SP, N P }) = x(U1 ∩ U2 ). Similarly one
sees that x ◦ x̃−1 is smooth. This shows that x and x̃ are C ∞ -compatible. Hence A
is a C ∞ -atlas.
(3) Let M = RPn , A := {xα : Uα → Rn | xα is an affine-linear embedding Rn → Rn+1
of maximal rank and 0 6∈ α(Rn )}, compare Example 1.1.11. All changes of charts
xβ ◦ xα −1 are rational functions and hence C ∞ . Therefore A is a C ∞ -atlas.
(4) Analogously, for M = CPn as in Example 1.1.15, the resulting atlas is also a C ∞ -
atlas.
13
1 Manifolds
Definition 1.2.5. An C ∞ -atlas Amax is called maximal (or also differentiable struc-
ture), if every chart that is C ∞ -compatible with all charts in Amax , is already con-
tained in Amax .
(x : U → V ) ∈ Amax (M ) with p ∈ U
is C k on W .
14
1.2 Differentiable manifolds
p b
f f (p) b
U Ũ N
f −1 (Ũ ) ∩ U
M x y
y◦f ◦ x−1
V Ṽ
x(p)
b
−1
W x(f (Ũ ) ∩ U )
Example 1.2.8
(1) Let M = S n with the differentiable structure given by
A = {(x : U1 → V1 ), (x̃ : U2 → V2 )}
15
1 Manifolds
16
1.3 Tangent vectors
The vague answer is: It is the linear approximation of the map at that point. But what
do we mean by the linear approximation in a point of a differentiable manifold? For this
to make sense we first need a concept of “linear approximation” of a manifold at a given
point.
Remark 1.3.2. This definition does not depend on the choice of the chart x : U → V .
Namely, if y : Ũ → Ṽ is another chart with p ∈ Ũ then we get by the chain rule
d d d
(y ◦ c)|t=0 = y ◦ x−1 ◦ (x ◦ c) |t=0 = D y ◦ x−1 |x(p) (x ◦ c)|t=0 . (1.5)
dt dt dt
d d
(y ◦ c1 )|t=0 = (y ◦ c2 )|t=0 .
dt dt
17
1 Manifolds
Lemma 1.3.5
Let M be an n-dimensional differentiable manifold, let p ∈ M and let x : U → V be a chart
of M with p ∈ U . Then the map
d
dx|p : Tp M → Rn , ċ(0) 7→ (x ◦ c)|t=0 ,
dt
is well defined and bijective.
Proof. Well-definedness and injectivity are clear from to the definition of the equiv-
alence relation that defines ċ(0). To show surjectivity let v ∈ Rn and set
−1
c(t) := x (x(p) + tv). Choose ε > 0 so small that x(p) + tv ∈ V whenever |t| < ε.
Then we have
d d
dx|p (ċ(0)) = x ◦ x−1 x(p) + tv |t=0 = x(p) + tv |t=0 = v.
dt dt
x
b
x(p)
b p v
U
M V
18
1.3 Tangent vectors
Definition 1.3.6. We equip Tp M with the unique vector space structure for which
dx|p becomes a linear isomorphism. In other words, for a, b ∈ R and
c1 : (−ε1 , ε1 ) → M , c2 : (−ε2 , ε2 ) → M we set:
a · ċ1 (0) + b · ċ2 (0) := (dx|p )−1 a · dx|p ċ1 (0) + b · dx|p ċ2 (0) .
Lemma 1.3.7
The vector space structure on Tp M does not depend on the choice of chart x : U → V .
Proof. Let y : Ũ → Ṽ be another chart with p ∈ Ũ . We have to show that the map
dy|p : Tp M → Rn is also linear with respect to the vector space structure induced by x.
This holds true since by (1.5)
dy|p = D y ◦ x−1 |x(p) ◦ dx|p
| {z } |{z}
linear linear
Lemma 1.3.8
Let M and N be differentiable manifolds, let p ∈ M , and let f : M → N be differentiable
near p. Then the map
ċ(0) f df |p (ċ(0))
b
c p f (p)b
f ◦c
M
N
19
1 Manifolds
Consequently, we have
−1
df |p = dy|f (p) ◦ D(y ◦ f ◦ x−1 )|x(p) ◦ dx|p .
Remark 1.3.10. If U ⊂ M is an open subset, then the differential of the inclusion map
ι : U ֒→ M is the canonical isomorphism dι : Tp U → Tp M , given by
We will identify tangent spaces via this isomorphism and simply write Tp U = Tp M .
M → Tp M, v p
v 7→ ċp,v (0), b cp,v
0
where cp,v (t) := p + tv.
Remark 1.3.12. For a chart x : U → V the differential dx|p has two meanings which are
20
1.3 Tangent vectors
∈
∈
dx|p
Tp U Tx(p) V = Tx(p) Rn
∼
dx|p ∼
= =
al
onic
d can rphism
dt (x ◦ c)|t=0 ∈ Rn isom
o
This proof of the chain rule was very simple. One may wonder why the proof of the
chain rule that one remembers from one’s course on calculus of several variables re-
quired a lot more work. The reason for the simplicity here is that one has already built
the chain rule into the definition of the differential of a map.
is a C k -diffeomorphism.
21
1 Manifolds
Proof. W.l o.g. let f be a C k -diffeomorphism. For a curve c : (−ε, ε) → M with c(0) = p
we have:
d(idM )|p ċ(0) = (idM ◦ c)˙(0) = ċ(0)
and hence
d(idM )p = idTp M .
f |U : U → Ũ
is a C k -diffeomorphism.
22
1.4 Directional derivatives and derivations
U p b f f (p) Ũ b
U2 N
U1 f −1 (U2 )
x y
M
y ◦ f ◦ x−1 |V
V1 Ṽ V2
x(p)
b
b
−1
y(f (p))
V x(f (U2 ) ∩ U1 )
On x(U1 ∩ f −1 (U2 )) the map y ◦ f ◦ x−1 is defined. Since df |p is invertible, we also have
that D(y ◦ f ◦ x−1 )|x(p) is invertible.
The ”classical” inverse function theorem says that there exists an open neighborhood
V ⊂ x(U1 ∩ f −1 (U2 )) of x(p) and an open neighborhood Ṽ ⊂ V2 of y(f (p)), such that
y ◦ f ◦ x−1 |V : V → Ṽ
Definition 1.4.1.
Let M be a differentiable manifold, let p ∈ M and and
let ċ(0) ∈ Tp M . For a function f : M → R, differenti- ċ(0) b
23
1 Manifolds
C k (U ) := {f : U → R | f is C k }.
α · f ∈ C k (U ), (α · f )(q) := α · f (q)
k
f + g ∈ C (U ∩ Ũ ), (f + g)(q) := f (q) + g(q)
f · g ∈ C k (U ∩ Ũ ), (f · g)(q) := f (q) · g(q)
and [
C∞
p := C ∞ (U ).
U open
p∈U
∂f = ∂(f |Ũ ).
Example 1.4.4
∂
(1) Let M = Rn and p ∈ M . Then ∂ = ∂xi p
is a derivation.
(2) Let M be an arbitrary differentiable manifold, let p ∈ M and ċ(0) ∈ Tp M . Then ∂ċ(0)
is a derivation. We check (iii):
d
∂ċ(0) (f · g) = (f · g) ◦ c |t=0
dt
24
1.4 Directional derivatives and derivations
d
= (f ◦ c) · (g ◦ c) |t=0
dt
d d
= (f ◦ c)|t=0 · g c(0) + f (c(0)) · (g ◦ c)|t=0
dt dt
= ∂ċ(0) f · g(p) + f (p) · ∂ċ(0) g.
Lemma 1.4.6
The map ∂· : Tp M → Der(Cp∞ ), ċ(0) 7→ ∂ċ(0) , is linear.
∂(f ◦ x−1 ) ∂f
∂(dx|p )−1 (ej ) (f ) = grad f ◦ x−1 |x(p) , ej = =:
∂xj x(p) ∂xj p
25
1 Manifolds
≈
U x V ⊂ Rn
(dx|p )−1 (e 2)
e2
p b
(dx|p )−1 (e1 ) b
e1
M x(p)
Proposition 1.4.8
Let M be a differentiable manifold and let p ∈ M . Then the map
form a basis of Der(Cp∞ ). Namely, then we know that the linear map ∂· maps the basis
(dx|p )−1 (e1 ), . . . , (dx|p )−1 (en ) of Tp M onto the basis ∂x∂ 1 p , . . . , ∂x∂ n p of Der(Cp∞ ) and is
hence an isomorphism.
n
X ∂
a) Linear Independence: Let αi = 0. We have to show: α1 = . . . = αn = 0.
∂xi p
i=1
Choose f = xj . Then
n
X ∂xj
0= αi = αj for j = 1, . . . , n.
∂xi p
i=1 | {z }
=δij
26
1.4 Directional derivatives and derivations
n
X ∂
δ= αj · .
∂xj p
j=1
b1) We have
(iii)
δ(1) = δ(1 · 1) = δ(1) · 1 + 1 · δ(1) = 2δ(1)
(ii)
δ(α) = δ(α · 1) = α · δ(1) = 0.
b2) Let f ∈ Cp∞ , more precisely f ∈ C ∞ (Ũ ) with p ∈ Ũ open. Choose a neighbor-
˜ of p with Ũ
hood Ũ ˜ ⊂ Ũ ∩ U and x(Ũ
˜ ) = B(x(p), r).
x
Ũ
b
x(p)
˜ p
b
U Ũ B(x(p),r)
M V
x(U ∩ Ũ )
Lemma 1.4.9 (see below) with h = f ◦ x−1 says that there exist g1 , . . . , gn ∈
C ∞ (B(x(p), r)) such that
n
X
f ◦ x−1 (x) = f ◦ x−1 (x(p)) + xi − xi (p) · gi (x) and
i=1
∂(f ◦ x−1 )
i
(x(p)) = gi x(p) .
∂x
It follows that
27
1 Manifolds
(i)
δ(f ) = δ(f | ˜ )
Ũ
n
X
= δ f (p) + xi − xi (p) (gi ◦ x)
i=1
(ii) n
X
(b1)
= δ xi − xi (p) (gi ◦ x)
i=1
n
X
(iii)
= δ xi − xi (p) · gi x(p) + xi − xi (p) |p δ(gi ◦ x)
i=1
| {z }
=0
(ii) n
X
(b1)
= δ xi gi x(p)
i=1
n
X ∂f
= αi · .
∂xi p
i=1
Lemma 1.4.9
Let h ∈ C ∞ (B(q, r)). Then there exist g1 , . . . , gn ∈ C ∞ (B(q, r)) with
n
X
(i) h(x) = h(q) + xi − q i gi (x) and
i=1
∂h
(ii) (q) = gi (q).
∂xi
Proof. For x ∈ B(q, r) set wx : [0, 1] → R, wx (t) := h(tx + (1 − t)q). It follows that
h(x) − h(q) = wx (1) − wx (0)
Z1
= ẇx (t) dt
0
Z1 X
n
∂h
= · (xi − q i ) dt
∂xi tx+(1−t)q
0 i=1
n Z1
X
i i ∂h
= (x − q ) dt
∂xi tx+(1−t)q
i=1 0
| {z }
=: gi (x)
28
1.4 Directional derivatives and derivations
With this definition of the gi , (i) holds. Moreover, (ii) follows from (i) by differentiation
at q.
∼
= ∼
=
dx|p ∂ ◦ (dx|p )−1
Rn
|{z}
∋ ej
depend on
the choice of x
From now on we identify Tp M with Der(Cp∞ ) via the isomorphism ∂· . For example, we
write for ξ ∈ Tp M
n
X ∂
ξ= ξi
∂xi p
i=1
n
P n
P
∂
instead of ∂ξ = ξ i ∂x i p and ξ = ξ i (dx|p )−1 (ei ) where (ξ 1 , . . . , ξ n )⊤ = dx|p (ξ).
i=1 i=1
Now we want to compute the coefficients ξ i in terms of the η j and vice versa. Using the
Chain Rule (Theorem 1.3.13) we compute
1 1
ξ1 η η
.. −1 .. −1 ..
. = dx|p (ξ) = (dx|p ) (dy|p ) . = D(x ◦ y )|y(p) . .
ξn ηn ηn
! !
η1 ξ1
Interchanging the roles of x and y, we also get .. = D(y ◦ x−1 )|x(p) .. . Thus
.n .n
η ξ
n
X ∂(y j ◦ x−1 )
ηj = · ξi (1.6)
∂xi x(p)
i=1
29
1 Manifolds
In the physics literature this transformation rule is put at the heart of the definition of a
tangent vector, then usually called a contravariant vector. For a physicist, a contravariant
vector is a vector (ξ 1 , . . . , ξ n ) associated to a chart which transforms as in (1.6) when the
chart is changed. We have now understood that this vector is the coefficient vector of
an (abstractly defined) tangent vector with respect to the basis ∂x∂ 1 p , . . . , ∂x∂ n p of Tp M
induced by the chart x.
∂ 1 n ⊤ = e . By (1.6), we get
Let us look at the special case ξ = ∂x i p , that is, (ξ , . . . , ξ ) i
n
∂ X ∂
= ηj
∂xi p ∂y j p
j=1
n Xn
X ∂(y j ◦ x−1 ) ∂
= ξk ·
∂xk x(p) ∂y j p
j=1 k=1
n
X ∂(y j ◦ x−1 ) ∂
= · ,
∂xi x(p) ∂y j p
j=1
hence
n
∂ X ∂(y j ◦ x−1 ) ∂
= · (1.7)
∂xi p ∂xi x(p) ∂y j p
j=1
In the physics literature it is customary to use the Einstein summation convention mean-
ing that when an index appears twice in an expression, once as an upper index and
once as a lower index, then summation over this index is understood. So (1.7) would
be written as
∂ ∂(y j ◦ x−1 ) ∂
i
= i
· j
∂x p ∂x x(p) ∂y p
or even shorter as
∂ ∂y j ∂
= · .
∂xi ∂xi ∂y j
This makes formula (1.7) easy to memorize; we simply cancel ∂y j . In these lecture notes
we will not use the Einstein summation convention unless explicitly stated otherwise.
But when you do computations for yourself, the Einstein summation convention can
be quite convenient and is recommended as long as you are aware of it.
30
1.5 Vector fields
Ux := π −1 (U ) ⊂ T M,
Vx := V × Rn ⊂ R2n and
Xx (ξ) := x π(ξ) , dx|π(ξ) (ξ) .
TM Ux
Xx Vx
U V
M x
By construction we have:
[
Ux = T M.
(x:U →V )
∈AM,max
1
Source: http://www.weatheronline.co.uk
31
1 Manifolds
M
π −1 (U1 ) ∩ π −1 (U2 ) = ∅.
AT M = {Xx : Ux → Vx | (x : U → V ) ∈ AM,max }.
x ◦ π ◦ Xx−1 : V × Rn → V, (v, w) 7→ v.
32
1.5 Vector fields
Definition 1.5.3
A map ξ : M → T M is called a vector field on M , ξ(p)
if for all p ∈ M we have p b
M
π ξ(p) = p.
Since a vector field is a map from the differentiable manifold M to the differentiable
manifold T M we know what it means that the vector field is C k . We investigate how
this can be characterized in terms of the coefficient functions. For the chart x of M we
consider the corresponding chart Xx on T M . The commutative diagram
ξ
M TM
∪ ∪
−1 ξ|U
x (v) ∈ U Ux ∋ ξ(x−1 (v))
x Xx
=x−1 (v)
z }| {
v∈ V V × Rn ∋ x π ξ(x−1 (v)) , dx| ξ(x−1 (v))
ξ(x−1 (v))
1 n
π
∩ ∩ = v, ξ (v), . . . , ξ (v)
Rn R2n
shows that ξ corresponds in these coordinates to the map v 7→ v, ξ 1 (v), . . . , ξ n (v) .
Thus ξ is C k on U if and only if the coefficient functions ξ 1 , . . . , ξ n are C k on V .
Example 1.5.5.
We consider M = R2 with polar coordinates. For ϕ0 ∈ R we set U :=
cos ϕ0
R2 \ R≥0 · , V := (0, ∞) × (ϕ0 , ϕ0 + 2π) and y : U → V such that
sin ϕ0
∂
On U the vector field ξ := r is defined. Using (1.7) we express this vector field in
∂r
33
1 Manifolds
∂
ξ = r
∂r
1
∂x ∂ ∂x2 ∂
= r +
∂r ∂x1 ∂r ∂x2
∂(r cos ϕ) ∂ ∂(r sin ϕ) ∂
= r +
∂r ∂x1 ∂r ∂x2
∂ ∂
= r cos ϕ 1 + sin ϕ 2
∂x ∂x
∂ ∂
= x1 1 + x2 2 .
∂x ∂x
In Cartesian coordinates:
ξ 1 (x1 , x2 ) = x1 ,
ξ 2 (x1 , x2 ) = x2 .
b
In polar coordinates:
η 1 (r, ϕ) = r,
η 2 (r, ϕ) = 0.
∂
Similarly, we can express the vector field
∂ϕ
in Cartesian coordinates:
∂ ∂x1 ∂ ∂x2 ∂
= + b
∂ϕ ∂ϕ ∂x1 ∂ϕ ∂x2
∂ ∂
= −r sin ϕ 1 + r cos ϕ 2
∂x ∂x
2 ∂ 1 ∂
= −x +x .
∂x1 ∂x2
34
2 Semi-Riemannian Geometry
On topological manifolds one can consider continuous maps. In order to be able to
define differentiable maps we had to add structure to a topological manifold which
gave rise to differentiable manifolds. We were then able to define linear approxima-
tions to manifolds (tangent spaces) and and to maps (the differential). The concept of a
differentiable manifold is what one needs to do analysis.
In order to do geometry we need to enrich our manifolds once more. We want to mea-
sure lengths of and angles between tangent vectors. This requires scalar products on
the tangent spaces and leads to the concept of a Riemannian manifold.
gij := g(bi , bj ) ∈ R
n
X n
X n
X
i j
g(v, w) = g α bi , β bj = αi β j gij .
i=1 j=1 i,j=1
35
2 Semi-Riemannian Geometry
Notation 2.1.2. Let b∗1 , . . . , b∗n the dual basis of the dual space V ∗ = {linear maps V →
R} of b1 , . . . , bn , that is b∗i (bj ) = δij . Often, we write
n
X
g= gij b∗i ⊗ b∗j .
i,j=1
in other words,
−1 0
..
.
(gij )i,j=1,...,n = −1 . (2.1)
1
..
.
0 1
Such a basis is called a generalized orthonormal basis. We the number of −1’s oc-
curring in (2.1) the index of g and denote it by Index(g). We observe that for a non-
degenerate symmetric bilinear form the following are equivalent:
(3) Index(g) = 0.
If B = (b1 , . . . , bn ) and B̃ = (b̃1 , . . . , b̃n ) are two bases of V , we define the transformation
matrix T = (tji )i,j=1,...,n by
n
X
b̃i = tji bj .
j=1
36
2.2 Semi-Riemannian metrics
that is, if Φ∗ gV = gW .
37
2 Semi-Riemannian Geometry
(x)
M , we define gij = gij : V → R by
!
∂ ∂
gij (v) := g|x−1 (v) , .
∂xi x−1 (v) ∂x j
x−1 (v)
Remark 2.2.2. Note the similarity of the definition of smoothness of g with the charac-
terization of smoothness of vector fields in Remark 1.5.4. We express the vector field
or semi-Riemannian metric with respect to the basis ∂x∂ 1 , . . . , ∂x∂ n of the tangent space
induced by a chart and then require smoothness of the coefficient functions.
n
(y)
X ∂(xk ◦ y −1 ) ∂(xl ◦ y −1 ) (x)
gij (v) = · · gkl (x ◦ y −1 )(v) (2.3)
∂y i v ∂y j v
k,l=1
38
2.2 Semi-Riemannian metrics
Consequence. The condition that g is smooth does not have to be checked for all charts,
if suffices to check it for a subatlas of Amax (M ) which covers M .
Remark 2.2.4. Recall that dx|p : Tp M → Rn is a linear isomorphism for any chart
x : U → V with p ∈ U . In particular, dx1 |p , . . . , dxn |p ∈ (Tp M )∗ .
Definition 2.2.5. The dual space (Tp M )∗ =: Tp∗ M is called cotangent space of M
at p.
Lemma 2.2.6
∂ ∂
The dx1 |p , . . . , dxn |p form the dual basis of ,..., .
∂x1 p ∂xn p
∂
Proof. Since dx|p = ei we have dxj |p ∂
∂xi p
= δij for i = 1, . . . , n.
∂xi p
n
X
g|p = gij x(p) · dxi |p ⊗ dxj |p
i,j=1
In the physics literature you will find the following short version of this equation:
n
P
If one changes the basis of a vector space by the transformation b̃i = tji bj , then we
j=1
get
n
X
b∗i = tij b̃∗j .
j=1
39
2 Semi-Riemannian Geometry
= δki ,
hence
n
X
tij b̃∗j = b∗i .
j=1
n
X ∂(xi ◦ y −1 )
dxi |p = · dy j |p
∂y j y(p)
j=1
∂xi j
dxi = dy
∂y j
If you have forgotten the transformation formula (2.3), you can quickly deduce it in
“physics style” as follows:
(y) (x)
gkl · dy k · dy l = gij · dxi · dxj
i j
(x) ∂x k ∂x l
= gij · · dy · · dy
∂y k ∂y l
∂xi ∂xj (x)
= · ·g · dy k · dy l .
∂y k ∂y l ij
40
2.2 Semi-Riemannian metrics
g|p := Φ∗p β. We check the smoothness of g in Cartesian coordinate, i.e., in the chart
x = id : U = M → V = M .
∂ ∂ ∗
∂ ∂
g|p , = Φp β ,
∂xi p ∂xj p ∂xi p ∂xj p
∂ ∂
= β Φp , Φp
∂xi p ∂xj p
∂ ∂
= β dx|p , dx|p
∂xi p ∂xj p
= β(ei , ej ).
Consequently, the gij are constant, hence C ∞ . In this manner, we can equip M with a
semi-Riemannian metric with arbitrary index.
x = F −1 : U = F (V ) → V
41
2 Semi-Riemannian Geometry
∂F ∂F
= (v), j (v) .
∂xi ∂x
∂F ∂F
Hence gij = ,
∂xi ∂xj
, in particular, the gij are smooth.
Definition 2.2.9. A semi-Riemannian metric g, for which g|p is always positive de-
finite, is called Riemannian metric. A pair (M, g), consisting of a differentiable
manifold M and a (semi-)Riemannian metric g on M is called (semi-)Riemannian
manifold.
A semi-Riemannian metric g is called Lorentz metric, if g|p has always index 1. The
pair (M, g) is then called Lorentz manifold.
1
F : R2 → S 2 ⊂ R3 , F (x) = (4 − kxk2 , 4x).
4 + kxk2
One computes
∂F 1
1
= (−16x1 , 4(4 − (x1 )2 + (x2 )2 ), −8x1 x2 ),
∂x (4 + kxk2 )2
∂F 1
2
= (−16x2 , −8x1 x2 , 4(4 + (x1 )2 − (x2 )2 )).
∂x (4 + kxk2 )2
Moreover,
∂F ∂F 16
g11 = , =
∂x1 ∂x1 (4 + kxk2 )2
and similarly for the other gij . The metric in these coordinates turns out to be
16 1 0
(gij ) = .
(4 + kxk2 )2 0 1
42
2.2 Semi-Riemannian metrics
Example 2.2.11. Let M ⊂ Rn+1 be open. The Minkowski scalar product hh·, ·ii on Rn+1 has
index 1, where
hhx, yii = −x0 y 0 + x1 y 1 + · · · + xn y n
The Lorentz manifold (Rn+1 , gMink ) is called Minkowski space. The four-dimensional
Minkowski space is the mathematical model for spacetime in special relativity.
Example 2.2.12. We express the Euclidean metric geucl = dx1 ⊗ dx1 + dx2 ⊗ dx2 of R2 in
polar coordinates. Here x1 and x2 are the Cartesian coordinates. With x1 = r cos ϕ and
x2 = r sin ϕ we then find:
∂x1 ∂x1
dx1 = dr + dϕ = cos ϕ dr − r sin ϕ dϕ
∂r ∂ϕ
∂x2 ∂x2
dx2 = dr + dϕ = sin ϕ dr + r cos ϕ dϕ.
∂r ∂ϕ
Thus
geucl = (cos ϕ dr − r sin ϕ dϕ) ⊗ (cos ϕ dr − r sin ϕ dϕ)
+(sin ϕ dr + r cos ϕ dϕ) ⊗ (sin ϕ dr + r cos ϕ dϕ)
= cos2 ϕ dr ⊗ dr − r cos ϕ sin ϕ dr ⊗ dϕ − r sin ϕ cos ϕ dϕ ⊗ dr + r 2 sin2 ϕ dϕ ⊗ dϕ
+ sin2 ϕ dr ⊗ dr + sin(ϕ)r cos ϕ dr ⊗ dϕ + r cos ϕ sin ϕ dϕ ⊗ dr + r 2 cos2 ϕ dϕ ⊗ dϕ
= dr ⊗ dr + r 2 dϕ ⊗ dϕ
and hence
Polar 1 0
gij =
0 r2
This matrix tells us: ∂
∂ ∂ϕ
• has length 1,
∂r
b
∂ b
• has length r, ∂
∂ϕ
∂r
∂ ∂
• and are orthogonal to each other
∂r ∂ϕ
In Cartesian coordinates we have: ∂
∂x2
1 0
Cartes
gij = b
0 1 ∂
∂x1
43
2 Semi-Riemannian Geometry
Remark 2.2.15. The metric ϕ∗ g is the unique semi-Riemannian metric on M , for which
ϕ is a local isometry.
Isom(M, g) := {ϕ : M → M isometry}
Remark 2.2.17. The set Isom(M, g) is a group with respect to composition of maps. The
neutral element is idM .
44
2.2 Semi-Riemannian metrics
We will see later that the inverse conclusion also holds; the isometries of (Rn , geucl ) are
precisely the Euclidean motions.
Example 2.2.19. To find isometries of Minkowski space (M, g) = (Rn+1 , gMink ) we de-
fine
O(n, 1) := A ∈ Mat((n + 1) × (n + 1), R) | hhAy, Azii = hhy, zii ∀ y, z ∈ Rn+1
= A ∈ Mat((n + 1) × (n + 1), R) | A⊤ I1,n , A = I1,n
where
−1 0 ··· 0
.. ..
0 1 . .
I1,n = ..
.
.. ..
. . . 0
0 ··· 0 1
Now affine transformations ϕ : →Rn+1 Rn+1 ,
ϕ(x) = Ax + b with A ∈ O(n, 1) and
b ∈ R n+1 , are called Poincaré transformations. The same discussion as for Euclidean
space shows
{Poincaré transformations} ⊂ Isom(Rn+1 , gMink ).
45
2 Semi-Riemannian Geometry
Again, we will see later that equality holds; the isometries of Minkowski space are
precisely the Poincaré transformations.
Example 2.2.20. To find isometries of the sphere (M, g) = (S n , gstd ) let A ∈ O(n + 1).
We set ϕ := A|S n : S n → S n . Let Φp : Tp S n → Rn+1 be as in Example 2.2.8. Then the
diagram
dϕ|p
Tp S n Tϕ(p) S n
Φp Φϕ(p)
A
Rn+1 Rn+1
commutes because:
d d d
c|t=0 A· c|t=0 = (A · c)|t=0
dt dt dt
Therefore
gstd (dϕ|p (ξ), dϕ|p (η)) = hΦϕ(p) (dϕ|p (ξ)), Φϕ(p) (dϕ|p (η))i
= hAΦp (ξ), AΦp (η)i
= hΦp (ξ), Φp (η)i
= gstd (ξ, η).
46
2.3 Differentiation of vector fields
The first idea that comes to one’s mind is to differentiate the coefficient functions η j in
the direction ξ. This would yield the expression
n
X ∂η j ∂
ξi · ·
∂xi x(p) ∂xj p
i,j=1
∂
ξ=η= .
∂ϕ
Then the derivative of η in direction ξ equals 0 because the coefficient functions η j are
constant. On the other hand, in Cartesian coordinates (x1 , x2 ) we get
∂ ∂
ξ = η = −x2 1
+ x1 2 .
∂x ∂x
For the derivative of η in direction ξ we would then find
∂ ∂ ∂ ∂ ∂ ∂
−x2 1 + x1 2 (−x2 ) 1 + −x2 1 + x1 2 (x1 ) 2
∂x ∂x ∂x ∂x ∂x ∂x
∂ ∂ ∂
= −x1 1
− x2 2 = −r 6= 0.
∂x ∂x ∂r
We see that the idea of simply differentiating the coefficient functions was to naive.
Since we do not know how to come up with a better definition we follow an axiomatic
approach similar to the concept of derivations, except that this time we differentiate
vector fields rather than functions.
Notation 2.3.2. Let M be a differentiable manifold and let k ∈ N∪{∞}. For any open
subset U ⊂ M we put
For p ∈ M we set [
Ξp := C ∞ (U, T M ).
U ⊂M open
with p∈U
47
2 Semi-Riemannian Geometry
Now we list the properties that the derivative of vector fields should have. Differenti-
ation takes a tangent vector ξ ∈ Tp M and a smooth vector field η defined near p and
gives us a tangent vector in Tp M as a result. Hence it is a map Tp M × Ξp → Tp M .
∇ : Tp M × Ξp → Tp M
∇ξ η = ∇ξ (η|Ũ ).
∇ξ (η1 + η2 ) = ∇ξ η1 + ∇ξ η2 .
∇ξ (f · η) = ∂ξ f · η|p + f (p) · ∇ξ η.
(vi) Torsion-freeness
For all charts x : U → V of M with p ∈ U we have:
∂ ∂
∇ ∂ j
=∇ ∂ i
∂xi p ∂x ∂xj p ∂x
48
2.3 Differentiation of vector fields
Remark 2.3.4
(1) From (iii) and (iv) we get the R-linearity in the second argument. Let α, β ∈ R:
(iii)
∇ξ (α η1 + β η2 ) = ∇ξ (α η1 ) + ∇ξ (β η2 )
(iv)
= ∂ξ α ·η1 |p + α∇ξ (η1 ) + ∂ξ β ·η2 |p + β∇ξ (η2 )
|{z} |{z}
=0 =0
= α∇ξ (η1 ) + β∇ξ (η2 ).
(2) If (vi) holds in a chart x, then it also holds in every other chart y containing p.
n
!
∂ X ∂xk ∂
∇ ∂ j
= ∇ ∂
∂y i p ∂y ∂y i p ∂y j p ∂xk
k=1
(iii) n
!
(iv) X ∂ 2 xk ∂ ∂xk ∂
= · + ∇
∂y i ∂y j y(p) ∂xk ∂y j y(p) ∂y∂ i p ∂xk
k=1
n n
(ii) X ∂ 2 xk ∂ X ∂xk ∂xl ∂
= · k+ · ∇ ∂
∂y i ∂y j y(p) ∂x ∂y j y(p) ∂y i y(p) ∂xl p ∂xk
k=1 k,l=1
49
2 Semi-Riemannian Geometry
Pn i ∂
Remark 2.3.6. The Christoffel symbols determine ∇. Namely, let ξ = i=1 ξ ∂xi p
∈
P
Tp M and η = nj=1 η j ∂x∂ j ∈ Ξp . Then we compute:
n (ii) n
X
j ∂ (iii) X i j ∂
∇Pn ξ i ∂ η = ξ∇ ∂ η
i=1 ∂xi p ∂xj ∂xi p ∂xj
j=1 i,j=1
n
!
(iv) X ∂η j ∂ ∂
= ξi · + η j |x(p) · ∇ ∂
∂xi x(p) ∂x
j
p ∂xi p
∂xj
i,j=1
n n
!
X
i ∂η j ∂ j
X ∂
= ξ · + η |x(p) · Γkij (x(p)) ·
∂xi x(p) ∂x j
p ∂xk p
i,j=1 k=1
n n
X ∂η k X ∂
= ξi + η j |x(p) · Γkij (x(p)) (2.5)
∂xi x(p) ∂xk p
i,k=1 j=1
Remark 2.3.7. Torsion freeness is equivalent to the Christoffel symbols being symmet-
ric in the two lower indices:
Theorem 2.3.8
Let (M, g) be a semi-Riemannian manifold and let p ∈ M . Then there is exactly one Levi-
Civita connection at p.
50
2.3 Differentiation of vector fields
∂gij
= Γlki · glj + Γlkj · gil (2.6)
∂xk
i→i ∂gik
j →k = Γlji · glk + Γljk · gil (2.7)
k→j ∂xj
i →k ∂gkj
j →j = Γlik · glj + Γlij · gkl (2.8)
k→ i ∂xi
Equation (2.6) − (2.7) + (2.8) together with the symmetry of the Christoffel symbols in
the lower indices yields:
g ij · gjk = δki .
Therefore
∂gij ∂gik ∂gkj
k
− j
+ gjm = 2 Γlki · glj · gjm = 2 Γlki · δlm = 2 Γm
ki
∂x ∂x ∂xi
and hence
1 ∂gij ∂gik ∂gkj
Γm
ki = k
− j
+ gjm .
2 ∂x ∂x ∂xi
Renaming indices (k → j, m → k, j → m) we obtain:
n
1 X mk ∂gim ∂gjm ∂gij
Γkij = g + − m (2.9)
2 ∂xj ∂xi ∂x
m=1
51
2 Semi-Riemannian Geometry
We check the second product rule (v), using the Einstein summation convention and
occasional renaming of indices:
Remark 2.3.9. For any chart x : U → V on (M, g) the Christoffel symbols are smooth
functions
n
k 1 X mk ∂gim ∂gjm ∂gij
Γij = g · + − m : V → R.
2 ∂xj ∂xi ∂x
m=1
Remark 2.3.10. Our naive ansatz to differentiate vector fields by simply differentiat-
ing the coefficient functions corresponds to formula (2.5) with Γkij = 0. The problem
was that this depends on the choice of coordinates. When we use formula (2.5) with
the correct definition (2.9) for the Christoffel symbols, then we get the uniquely deter-
mined Levi-Civita connection. In particular, this kind of differentiating vector fields is
independent of the choice of chart.
Note however, that the Levi-Civita connection depends on the semi-Riemannian met-
ric. This cannot only seen from (2.9) but also from the second product rule (v) in Defi-
nition 2.3.3 which involves the metric. There is nothing we can do about this; different
semi-Riemannian metrics will in general lead to different Levi-Civita connections.
So the situation is somewhat curious: Differentiability and the derivative of a function
are well defined on a differentiable manifold. Differentiability of a vector field is also
well defined on a differentiable manifold. But in order to define the derivative of a
vector field we need a semi-Riemannian metric.
52
2.3 Differentiation of vector fields
Definition 2.3.11. Let (M, g) be a semi-Riemannian manifold and let ∇ be its Levi-
Civita connection. Let p ∈ M , let ξ ∈ Tp M and let η ∈ Ξp . Then
∇ξ η ∈ Tp M
Example 2.3.12. Let (M, g) = (R2 , geucl ) be the 2-dimensional Euclidean space. In
Cartesian coordinates x1 , x2 the gij = δij are constant. Therefore Γkij = 0. In this case,
covariant differentiation is indeed given by differentiation of the coordinate functions.
For example,
∂ 2 ∂ 1 ∂
∇∂ = ∇−x2 ∂ +x1 ∂ −x +x
∂ϕ ∂ϕ ∂x1 ∂x2 ∂x1 ∂x2
2 ∂ 1 ∂ 2 ∂ 2 ∂ 1 ∂ ∂
= −x 1
+x 2
(−x ) 1 + −x 1
+x 2
(x1 ) 2
∂x ∂x ∂x ∂x ∂x ∂x
∂ ∂ ∂
= −x1 1 − x2 2 = −r
∂x ∂x ∂r
In polar coordinates r, ϕ we have
1 0 ij 1 0
(gij )(r, ϕ) = and (g )(r, ϕ) = 1 .
0 r2 0 r2
1
Γ111 = 1 · (0 + 0 − 0) + 0 · . . . = 0.
2
and similarly
Γ111 = Γ211 = Γ112 = Γ121 = Γ222 = 0.
Moreover:
1 1 ∂g12 ∂g22 ∂g12 1
Γ212 = Γ221 = + − + 0 · ... = and Γ122 = −r.
2 r2 ∂ϕ ∂r ∂ϕ r
Thus
∂ ∂ ∂ ∂
∇ ∂ = Γ122 + Γ222 = −r .
∂ϕ ∂ϕ ∂r ∂ϕ ∂r
Indeed, we obtained the same result for both computations, one in Cartesian and one
in polar coordinates.
53
2 Semi-Riemannian Geometry
Remark 2.3.14. To compute ∇ξ η with ξ = ċ(0) we only need to know η along the curve
c. Namely,
n n
X ∂ ∂ X
∇ċ(0) ηj = ∇Pn ċi (0) ∂ ηj j
∂xj i=1 ∂xi ∂x
j=1 j=1
n
j ∂
X
i
= ċ (0) ∇ ∂ η
∂xi ∂xj
i,j=1
n n
!
X ∂η j ∂ X ∂
= ċi (0) i j
+ η j Γkij
∂x ∂x ∂xk
i,j=1 k=1
n n
X d j ∂ X ∂
= η ◦ c |t=0 j + ċi (0) η j (c(0))Γkij (c(0)) k .
dt ∂x ∂x
j=1 i,j,k=1
πM ◦ ξ = ϕ.
54
2.4 Vector fields along maps
Example 2.4.2
(1) Vector fields along curves. Let N = I ⊂ R be an open interval and c = ϕ : N = I → M
be a curve.
c
ξ
N
M
An important special case is given by ξ(t) = ċ(t) := ċt (0) where ct (s) := c(t + s).
This is the velocity field of c.
c
ξ
N
M
(2) If N = M and ϕ = id then a vector field along id is just a vector field in the usual
sense.
(3) Let ϕ be constant, i.e., ϕ(x) = p for all x ∈ N . Then a vector field along ϕ is a map
N → Tp M .
(4) Let ϕ be differentiable and let ξ be a vector field on N . Then
p 7→ dϕ|p ξ(p) ∈ Tϕ(p) M
is a vector field along ϕ.
(5) If ξ is a vector field on M then
p 7→ ξ ϕ(p)
is a vector field along ϕ.
55
2 Semi-Riemannian Geometry
Proposition 2.4.4
Let N be a differentiable manifold, (M, g) a semi-Riemannian manifold and ϕ : N → M a
differentiable map. Let η, η1 , η2 be differentiable vector fields along ϕ. Let α1 , α2 ∈ R and
f : N → R be a differentiable function. Furthermore, let p ∈ N and ξ, ξ1 , ξ2 ∈ Tp N .
Then the covariant derivative ∇ξ η is defined independently of the choice of chart x and the
choice of curve c with ċ(0) = ξ and we have:
(i) If η is the form η = ζ ◦ ϕ where ζ is a differentiable vector field on M , then we have
∇ξ η = ∇dϕ|p (ξ) ζ.
(vii) Torsion freeness: For all charts y of N and all i, j = 1, . . . , dim(N ) we have:
∂ ∂
∇ ∂ dϕ = ∇ ∂ dϕ .
∂y i ∂y j ∂y j ∂y i
56
2.4 Vector fields along maps
Proof. The assertions follow directly from the definition and the corresponding state-
ments for the Levi-Civita connection.
Remark 2.4.6. For a vector field along a curve c : I → M we have the following formula
in local coordinates on M :
n n
∇η X
η̇ k (t) +
X ∂
(t) = ċi (t) · η j (t) · Γkij x c(t) .
dt ∂xk c(t)
k=1 i,j
Example 2.4.7. Let (M, g) = (Rn , geucl ) or (M, g) = (Rn , gMink ). Then the gij are con-
stant in Cartesian coordinates. Consequently, the Christoffel symbols with respect to
Cartesian coordinates vanish, Γkij = 0.
P
For a C 1 -curve c : I → M and a C 1 -vector field ξ along c with ξ(t) = nj=1 ξ j (t) ∂x∂ j |c(t)
we have:
n
∇ X ∂
ξ(t) = ξ̇ j (t) .
dt ∂xj c(t)
j=1
Hence, in this case, covariant differentiation just consists of differentiation of the coef-
ficient functions. Note however, that this is no longer true in other coordinate systems
such as polar coordinates.
57
2 Semi-Riemannian Geometry
Example 2.4.8. In the Euclidean plane (M, g) = (R2 , geucl ) we consider the circle line
c(t) = (cos(t), sin(t)) and its velocity field
∂ ∂
ξ(t) = ċ(t) = − sin(t) + cos(t) .
∂x1 c(t) ∂x2 c(t)
∇ ∇ ∂ ∂ ∂
ξ(t) = ċ(t) = − cos(t) − sin(t) =− .
dt dt ∂x1 c(t) ∂x2 c(t) ∂r c(t)
For the fun of it, let us also carry out the calculation in polar coordinates (r, ϕ). Now
∂
c1 (t) = r(t) = 1, c2 (t) = ϕ(t) = t and ξ(t) = ∂ϕ c(t) , i.e., ξ 1 (t) = 0 and ξ 2 (t) = 1. This
time there are no derivatives of the coefficients of ξ but we have to take the Christoffel
symbols into account. Recall from Example 2.3.12 that there are three non-vanishing
Christoffel symbols for polar coordinates,
1
Γ212 = Γ221 = , Γ122 = −r.
r
Therefore we get
2 2
∇ X ∂ X ∂
ξ(t) = i
ċ (t) ξ j
(t) Γ1ij
r(t), ϕ(t) + ċi (t) ξ j (t) Γ2ij r(t), ϕ(t)
dt ∂r c(t) ∂ϕ c(t)
ij=1 ij=1
∂ 1 1 ∂
= ċ2 (t) ξ 2 (t) − r(t) + ċ1 (t) ξ 2 (t) + ċ2 (t) ξ 1 (t)
∂r c(t) r(t) r(t) ∂ϕ c(t)
∂ ∂
= 1 · 1 · (−1) + (0 · 1 · 1 + 1 · 0 · 1)
∂r c(t) ∂ϕ c(t)
∂
= − .
∂r c(t)
∇
ξ ≡ 0.
dt
58
2.5 Parallel transport
Pn Let j(M, g)
Example 2.5.2. = (Rn , geucl ) or (Rn , gMink ). In Cartesian coordinates, a vector
∂
field ξ(t) = j=1 ξ (t) ∂xj c(t) along a curve c is parallel if and only if ξ̇ j (t) = 0 for all
t ∈ I, i.e., if and only if the ξ j are constant.
Example 2.5.3. Let (M, g) = (R2 , geucl ). Recall from Example 2.3.12 that the Christoffel
symbols in polar coordinates (r, ϕ) are given by:
1
Γ111 = Γ211 = Γ112 = Γ121 = Γ222 = 0, Γ212 = Γ221 = , Γ122 = −r.
r
∂ ∂
Thus ξ = ξ 1 ∂r + ξ 2 ∂ϕ is parallel along a curve c if and only if
∇
0 = ξ
dt
∂ ∂ ∂ ∂
= ξ̇ 1 + ξ 1 ∇ċ1 ∂ +ċ2 ∂ + ξ˙2 + ξ 2 ∇ċ1 ∂ +ċ2 ∂
∂r ∂r ∂ϕ ∂r ∂ϕ ∂r ∂ϕ ∂ϕ
1 ∂ 1 1 2 1 ∂ ˙2 ∂ 2 1 1 ∂ 2 1 ∂
= ξ̇ + ξ ċ · 0 + ċ · 1 +ξ + ξ ċ 1 + ċ (−c )
∂r c ∂ϕ ∂ϕ c ∂ϕ ∂r
∂ 2 1
ċ ċ ∂
= ξ̇ 1 − c1 ċ2 ξ 2 + ξ̇ 2 + 1 ξ 1 + 1 ξ 2 .
∂r c c ∂ϕ
This is equivalent to:
ċ2 ċ1
ξ˙1 − c1 ċ2 ξ 2 = 0, ξ˙2 + 1 ξ 1 + 1 ξ 2 = 0,
c c
that is 1 1
ξ˙ 0 c1 ċ2 ξ
= 2 1 .
ξ˙2 − ċ1 c
− cċ1 ξ2
This is a system of linear first order ordinary differential equations for (ξ 1 , ξ 2 ).
M
Proposition 2.5.4
ξ0
Let (M, g) be a semi-Riemannian manifold
and c : I → M be a C 1 -curve and t0 ∈ I.
For any ξ0 ∈ Tc(t0 ) M there exists exactly one ξ
b
c
c(t0 )
parallel vector field ξ along c with ξ(t0 ) = ξ0 .
59
2 Semi-Riemannian Geometry
Proof. Case 1: Let c(I) be contained in one chart and let x : U → V be such a chart.
∇
Then the condition dt ξ = 0 is equivalent to
n
X
ξ˙k = − Γkij ◦ x ◦ c ċi · ξ j ,
i,j=1
which is a system of linear ordinary equations of first order. Hence there exists a unique
solution with initial condition
ξ 1 (t0 ), . . . , ξ n (t0 ) = (ξ01 , . . . , ξ0n ).
Since the system is linear, the solution is defined on all of I.
ξ0
Ui bc
bc
bc
c(t1 ) c(b1 )
c(t0 )
c
bc
c(a1 )
Not something like this!
∇
W.l.o.g. let c(t0 ) ∈ U1 , otherwise renumber the charts. We solve the equation ξ=0
dt
as in Case 1 with ξ(t0 ) = ξ0 in U1 .
If the solution is not defined on the whole of [a1 , b1 ], we choose t1 ∈ (a1 , b1 ) with
c(t1 ) ∈ U1 ∩ U2 . Then we solve the equation in the chart x2 with the initial condition
ξ(t1 ), given by the previous solution.
Due to uniqueness in Case 1 both parallel vector fields coincide on U1 ∩U2 . After finitely
many steps we get a parallel vector field which is defined on [a1 , b1 ].
The same holds true for the next compact subinterval [a2 , b2 ] and we obtain a parallel
vector field on [a2 , b2 ] which extends the one on [a1 , b1 ]. By induction, we then find a
parallel vector field on every [ai , bi ] extending the one on the smaller interval[ai−1 , bi−1 ].
S
Since N i=1 [ai , bi ] = (a, b) we obtain a parallel vector field ξ on (a, b) with ξ(t0 ) = ξ0 .
˜ 0 ) = ξ0 .
Uniqueness: Let ξ and ξ̃ be two parallel vector fields along c with ξ(t0 ) = ξ(t
Write I = Igood ⊔ Ibad where
˜
Igood = t ∈ I | ξ(t) = ξ(t)
˜
Ibad = t ∈ I | ξ(t) 6= ξ(t)
60
2.5 Parallel transport
Since ξ and ξ˜ are continuous, Igood is closed in I. For t1 ∈ Igood choose a chart x : U → V
which contains c(t1 ). By uniqueness in Case 1 we then have ξ(t) = ξ̃(t) for all t ∈ I with
c(t) ∈ U . Therefore a neighborhood of t1 is contained in Igood . Hence Igood is open in I.
We have seen that Igood is open and closed in I. It is also non-empty because t0 ∈ Igood .
Since I is connected, we have I = Igood and therefore ξ(t) = ξ(t) ˜ for all t ∈ I.
is called parallel transport along c. Here ξ(t) is the parallel vector field along c with
ξ(t0 ) = ξ0 .
Proposition 2.5.6
Let M , c, t0 , and t1 as in Definition 2.5.5 and let t2 ∈ I. Then we have:
Proof. (a) Let ξ0 , η0 ∈ Tc(t0 ) M . Let ξ, η the corresponding parallel vector fields along c.
Then
d ∇ ∇
g(ξ, η) = g ξ , η + g ξ, η = 0.
dt dt
|{z} dt
|{z}
=0 =0
Therefore g(ξ, η) is constant, hence
g Pc,t0 ,t1 (ξ0 ), Pc,t0 ,t1 (η0 ) = g ξ(t1 ), η(t1 )
= g ξ(t0 ), η(t0 )
= g(ξ0 , η0 ).
This proves that parallel transport is a linear isometry.
61
2 Semi-Riemannian Geometry
(b) is obvious.
Remark 2.5.7. For ξ0 ∈ Tc(t0 ) M the parallel vector field ξ with ξ(t0 ) = ξ0 is given by
Proposition 2.5.8
Let (M, g) be a semi-Riemannian manifold, let c : I → M be a C 1 -curve, and let t0 ∈ I.
Then for every C 1 -vector field ξ along c we get:
∇ Pc,t,t0 ξ(t) − ξ(t0 )
ξ = lim .
dt t0 t→t0 t − t0
Proof. Let e1 (t0 ), . . . , en (t0 ) be a basis of Tc(t0 ) M . Let e1 (t), . . . , en (t) be the correspond-
ing parallel vector fields along c.
By Proposition 2.5.6 P(a), we know that e1 (t), . . . , en (t) form a basis of Tc(t) M for every
t ∈ I. Write ξ(t) = nj=1 ξ j (t)ej (t). Then
=ej (t0 )
Pn z }| { P
Pc,t,t0 (ξ(t)) − ξ(t0 ) j=1 ξ j (t) P
c,t,t0 (ej (t)) − nj=1 ξ j (t0 )ej (t0 )
=
t − t0 t − t0
n
X ξ j (t) − ξ j (t0 )
= ej (t0 )
t − t0
j=1
n
X
t→t0
−→ ξ̇ j (t0 )ej (t0 ).
j=1
62
2.6 Geodesics
n
X
= ξ˙j (t0 )ej (t0 ).
j=1
dψ|c(t0 ) dψ|c(t1 )
Pc̃,t0 ,t1
Tc̃(t0 ) M̃ Tc̃(t1 ) M̃
Remark 2.5.10. In general, parallel transport depends on the curve joining two given
points. This means, in general we have Pc,t0 ,t1 6= Pĉ,s0,s1 if c and ĉ are two curves in M
with c(t0 ) = ĉ(s0 ) and c(t1 ) = ĉ(s1 ). In this respect, Euclidean space is not typical.
2.6 Geodesics
the energy of c.
63
2 Semi-Riemannian Geometry
Remark 2.6.2. If (M, g) is Riemannian, then g(ċ, ċ) ≥ 0 and therefore E[c] ≥ 0 (and
equal to 0 if and only if c is constant).
Question. Are there curves with minimal energy joining two given endpoints? More
generally, are there curves with “stationary energy”?
c : (−ε, ε) × [a, b] → M
with c(0, t) = c(t) for all t ∈ [a, b]. If c(s, a) = c(a) and c(s, b) = c(b) for all s ∈ (−ε, ε)
then we call c(s, t) a variation with fixed endpoints.
c(s, t) b
ξ
c(t) b b
c b
c
M M
∂c
The vector field ξ(t) := (0, t) is called the variational vector field.
∂s
Remark 2.6.4. The variational vector field ξ of a variation with fixed endpoints satisfies
ξ(a) = 0 and ξ(b) = 0.
64
2.6 Geodesics
Proof. We compute:
Zb
d 1 d
E[cs ] s=0
= g ċs (t), ċs (t) dt
ds 2 ds s=0
a
Zb
1 ∂ ∂c ∂c
= g (s, t), (s, t) dt
2 ∂s s=0 ∂t ∂t
a
Zb
1 ∇ ∂c ∂c ∂c ∇ ∂c
= g (0, t), (0, t) + g (0, t), (0, t) dt
2 ∂s ∂t ∂t ∂t ∂s ∂t
a
Zb
∇ ∂c ∂c
= g (0, t), (0, t) dt
∂s ∂t ∂t
a
Zb
(∗) ∇ ∂c ∂c
= g (0, t), (0, t) dt
∂t ∂s ∂t
a
Zb
∇
= g ξ(t), ċ(t) dt
dt
a
Zb
d ∇
= g(ξ(t), ċ(t)) − g ξ(t), ċ(t) dt
dt dt
a
Zb
∇
= g ξ(b), ċ(b) − g ξ(a), ċ(a) − g ξ(t), ċ(t) dt.
dt
a
Corollary 2.6.6
If the variation has fixed endpoints then
Zb
d ∇
E[cs ] s=0
= − g ξ(t), ċ(t) dt.
ds dt
a
65
2 Semi-Riemannian Geometry
Lemma 2.6.7
Let c : [a, b] → M be a smooth curve and ξ a smooth vector field along c. Then there exists a
variation c of c with variational vector field ξ. If ξ(a) = 0 and ξ(b) = 0, then we can choose
the variation with fixed endpoints.
Proof. a) We first consider the case that supp(ξ) is contained in a chart x : U → V , i.e.,
c(t) ∈ U whenever ξ(t) 6= 0.
ξ
x
U (c1 , . . . , cn )
c M
Rn ⊃ V
n
X ∂
We write ξ(t) = ξ j (t) and we set
∂xj c(t)
j=1
(
x−1 c1 (t), . . . , cn (t) + s ξ 1 (t), . . . , ξ n (t) , c(t) ∈ U
c(s, t) :=
c(t), c(t) ∈
6 U
Then we have for the corresponding variational vector field:
j
∂c j ∂c
(0, t) = dx (0, t)
∂s ∂s
∂(xj ◦ c)
= (0, t)
∂s
∂ cj (t) + sξ j (t)
=
∂s
s=0
= ξ j (t).
Hence the variation c has the variational vector field ξ. Moreover, if ξ vanishes at the
endpoints, then c has fixed endpoints.
b) In the general case, cover the compact set c([a, b]) with finitely many charts and
construct the variation piecewise.
Remark 2.6.8. Later, when we have the Riemannian exponential map at our disposal,
we will be able to directly write down a suitable variation without usage of charts.
66
2.6 Geodesics
Corollary 2.6.10
Let (M, g) be a semi-Riemannian manifold and c ∈ Ωp,q (M ). Then the following are equiv-
alent:
d
E[cs ] s=0
=0
ds
for all variations cs of c with fixed endpoints;
Proof. The implication “(ii)⇒(i)” is directly clear by Corollary 2.6.6. We show “(i)⇒(ii)”.
Let [a, b] be the parameter interval of c. Assume there exists a t0 ∈ (a, b) with
∇
dt ċ(t0 ) 6= 0. Then there exists a ξ0 ∈ Tc(t0 ) M with
∇
g ξ0 , ċ(t0 ) > 0
dt
because g is non-degenerate. Let ξ˜ be the parallel vector field along c with ξ(t
˜ 0 ) = ξ0 .
By continuity there exists an ε > 0 such that (t0 − ε, t0 + ε) ⊂ (a, b) and
∇
g ξ̃(t), ċ(t) > 0
dt
67
2 Semi-Riemannian Geometry
By Lemma 2.6.7 we can choose a variation of c with fixed endpoints and variational
vector field ξ. Then we have for this variation
Zb
d ∇
E[cs ] s=0
= − g ξ(t), ċ(t) dt < 0
ds dt
a
∇
which contradicts the assumption. Hence we have dt ċ = 0 on (a, b) and by continuity
also on the whole of [a, b].
∇
Definition 2.6.11. A smooth curve c with dt ċ = 0 is called a geodesic.
Example 2.6.12. Let (M, g) = (Rn , geucl ) or (Rn , gMink ). In Cartesian coordinates
x1 , . . . , xn we have:
∇
ċ = 0 ⇐⇒ c̈1 = 0, . . . , c̈n = 0
dt
⇐⇒ cj (t) = pj + tv j
⇐⇒ c(t) = p + tv.
Lemma 2.6.13
For any geodesic c the quantity g ċ, ċ is constant.
Proof. We compute
d
∇
dt g ċ, ċ = 2 · g dt ċ , ċ = 0.
|{z}
=0
68
2.6 Geodesics
If c : I → M and c̃ : I˜ → M are two such geodesics with c(0) = c̃(0) and ċ(0) = c̃(0),
˙ then
˜
c and c̃ coincide on their common domain I ∩ I.
69
2 Semi-Riemannian Geometry
Example 2.6.18. Let M = (R2 \ {0}, geucl ) be the Euclidean plane with the origin re-
moved and let M̃ = {(x, y, z) ∈ R3 | x2 + y 2 = z 2 /3, z < 0)} be a cone with the cone
tip removed and equipped with √ the first fundamental form g̃. Now ψ : M → M̃ ,
ψ(u, v) = 2√u12 +v2 (u2 − v 2 , 2uv, − 3(u2 + v 2 )), can be checked to be a local isometry.
Hence ψ maps straight lines in M onto geodesics in M̃ .
bc
bc
ψ
R2 \ {0}
Fix(ψ) := {p ∈ M | ψ(p) = p}
Proposition 2.6.20
Let (M, g) be a semi-Riemannian manifold and ψ ∈ Isom(M, g).
Then for any p ∈ Fix(ψ) and any ξ ∈ Tp M with dψ|p (ξ) = ξ the geodesic c : I → M with
70
2.6 Geodesics
have:
c̃(0) = ψ c(0) = ψ(p) = p = c(0) and
˙
c̃(0) = dψ|c(0) ċ(0) = dψ|p (ξ) = ξ = ċ(0).
Φp (ξ)
Example 2.6.21
We use Proposition 2.6.20 to determine the geodesics
of the sphere (S n , gstd ). Let p ∈ S n and ξ ∈ Tp S n . Let p
E ⊂ Rn+1 be the two-dimensional vector subspace
b
0
reflection about E. Then A ∈ O(n + 1). Hence
c(0) = p is satisfied.
d Φp (ξ)
c(0) = · α and therefore α = ||Φp (ξ)|| = ||ξ|| .
dt ||Φp (ξ)||
d
Then we get dt c(0) = Φp (ξ), i.e., ċ(0) = ξ. Thus the geodesic c with initial conditions
c(0) = p and ċ(0) = ξ is given by
Φp (ξ)
c(t) = p · cos ||ξ|| t + · sin ||ξ|| t .
||ξ||
71
2 Semi-Riemannian Geometry
c̃(0) = cξ (0) = p,
˙
c̃(0) = α · ċξ (0) = αξ,
Then we call expp : Dp → M the Riemannian exponential map (at the point p).
Remark 2.6.24
(1) By Remark 2.6.22 we know expp (t · ξ) = ctξ (1) = cξ (t). Thus t 7→ expp (tξ) is the
geodesic with initial values p and ξ.
(2) For any p ∈ M we have expp (0) = p because c0 is the constant curve c0 (t) = p.
72
2.6 Geodesics
Example 2.6.25
(1) Let (M, g) = (Rn , geucl ) or (Rn , gMink ). Then we have:
expp (ξ) = p + 1 · Φp (ξ) = p + Φp (ξ).
Here Dp = Tp Rn .
p
Dp = Tp M \ − t · Φp −1 (p) | t ≥ 1 . 0 bc
R2
Tp M
Dp
(3) Let (M, g) = (S n , gstd ). Then we have Dp = ξ
b b
Tp M and M p
Φp (ξ)
expp (ξ) = p · cos (||ξ||) + · sin (||ξ||).
||ξ||
Lemma 2.6.26
The differential of the map expp : Dp → M at 0 is given by the canonical isomorphism
d expp |0 = Φ0 : T0 Dp = T0 Tp M → Tp M.
Corollary 2.6.27
For p ∈ M there exists an open neighborhood Vp ⊂ Dp ⊂ Tp M of 0, such that
73
2 Semi-Riemannian Geometry
is a diffeomorphism.
Proof. By Lemma 2.6.26 d expp |0 is invertible. The inverse function theorem yields the
claim.
Φp (ξ) Tp S n
expp (ξ) = p · cos ||ξ|| + · sin ||ξ|| . p
||ξ|| b
b
{ξ ∈ Tp M | ||ξ|| = π}
In particular, for any ξ ∈ Tp M with ||ξ|| = π
we have Sn
b
−p
expp (ξ) = p · cos(π) = −p.
For ξ ∈ Tp M with ||ξ|| = π the differential d expp |ξ has the (n − 1)-dimensional kernel
{η ∈ Tξ Tp S n | Φξ (η) ⊥ ξ}.
Now we construct coordinates which are well adapted to the geometry and to this end
we choose a generalized orthonormal basis E1 , . . . , En of Tp M regarding g|p , that is
74
2.6 Geodesics
Definition 2.6.30. The coordinates we just defined are called Riemannian normal
coordinates around the point p.
Proposition 2.6.31
Let (M, g) be a semi-Riemannian manifold and p ∈ M . Let gij : Vp → R be the metric
coefficients and Γkij : Vp → R be the Christoffel symbols in Riemannian normal coordinates
around p. Then we have:
Proof. a) Clearly, we have x(p) = A−1 expp −1 (p) = A−1 (0) = 0.
Here ck (t) = xk (c(t)) = tv k , ċk (t) = v k and c̈k (t) = 0. For t = 0 we get
n
X
0=0+ Γkij (0, . . . , 0) · v i · v j .
i,j=1
75
2 Semi-Riemannian Geometry
Pn
For each k we define a bilinear form β k on Rn by β k (y, z) := k i j
i,j=1 Γi,j (0) y z .
These bilinear forms are symmetric because:
n
X n
X n
X
k
β (z, y) = Γkij (0) z i j
y = Γkji (0) z j i
y= Γkij (0) y i z j = β k (y, z).
i,j=1 j,i=1 i,j=1
Exchanging ∇ free of
indices torsion
Since we know that β k (v, v) = 0 for all v ∈ Rn , polarization yields β k (y, z) = 0 for
all y, z ∈ Rn . This means Γkij (0) = 0 for all i, j, k.
Tp M
M p b
Geodesic
in M
through p
Example 2.6.32. Let (M, g) = (Rn , geucl ) or (Rn , gMink ) and p ∈ M . Choose
A = Φp = canonical isomorphism Rn → Tp Rn .
Then we have expp (Av) = p + v, thus Riemannian normal coordinates around p are
given by
x : Rn → Rn , x(q) = q − p.
Corollary 2.6.33
In Riemannian normal coordinates we have for the Taylor expansion of gij : Vp → R
around 0:
gij (x) = εi δij + O(||x||2 ).
76
2.6 Geodesics
77
3 Curvature
We now come to one of the central concepts of differential geometry, that of curvature.
We will see that there are various inequivalent notions of curvature. We start with the
most basic one.
∇2ξ,η ζ := ∇ξ ∇η ζ − ∇∇ξ η ζ ∈ Tp M
Lemma 3.1.2
The second covariant derivative ∇2ξ,η ζ depends on η only via η|p , i.e., if η, η̃ ∈ Ξp (M ) with
η|p = η̃|p then
∇2ξ,η ζ = ∇2ξ,η̃ ζ.
79
3 Curvature
and therefore
j
i ∂η k ∂ ∂η j ∂ζ k ∂
∇∇ξ η ζ = ξ i
(0)∇ ∂ ζ = ξi (0) (0) . (3.1)
∂x ∂xj p ∂xk ∂x i ∂x j ∂xk p
Moreover,
k ∂ j ∂ζ k ∂ ∂
∇η ζ = ∇η j ∂ ζ =η + ζ k Γm
jk
∂xj ∂xk j
∂x ∂x k ∂xm
and hence (again using that the Christoffel symbols vanish)
k ∂
j ∂ζ j k m ∂
∇ξ ∇η ζ = ∇ξ i ∂ η + η ζ Γjk m
∂xi p ∂xj ∂xk ∂x
∂η j ∂ζ k ∂ ∂2ζ k ∂ ∂Γm
jk ∂
= ξi i
(0) j
(0) + ξ i η j (0) (0) + ξ i η j (0)ζ k (0) .
∂x ∂x ∂xk p
j
∂x ∂x i ∂xk p ∂xi ∂xm p
(3.2)
Subtracting (3.1) from (3.2) we see that the terms containing a derivative of the η j cancel
and we are left with
2 i j ∂2ζ k i j m ∂Γkim ∂
∇ξ,η ζ = ξ η (0) i j (0) + ξ η (0)ζ (0) i
(0) . (3.3)
∂x ∂x ∂x ∂xk p
This expression depends on η only via the η j (0) which are the coefficients of η|p .
Lemma 3.1.4
For ξ, η ∈ Tp M and ζ ∈ Ξp (M )
depends only on ζ via ζ|p . Thus R(ξ, η)ζ ∈ Tp M is well defined for ξ, η, ζ ∈ Tp M .
Proof. Again we choose Riemann normal coordinates around p and recall (3.3):
!
2 i j ∂2ζ k m
∂Γkjm ∂
∇ξ,η ζ = ξ (0)η (0) i j
(0) + ζ (0) i
(0) .
∂x ∂x ∂x ∂xk p
80
3.1 The Riemannian curvature tensor
R(ξ, η)ζ
!
k
∂2ζ k ∂Γ jm ∂
= ξ i (0)η j (0) − ξ j (0)η i (0) i j
(0) + ζ m (0) i
(0)
∂x ∂x ∂x ∂xk p
!
i j ∂2ζ k ∂2ζ k ∂Γkjm ∂Γkim ∂
= ξ (0)η (0) (0) − (0) + ζ m (0) (0) m
− ζ (0) (0)
∂xi ∂xj ∂xj ∂xi ∂xi ∂x j ∂xk p
!
∂Γkjm ∂Γkim ∂
= ξ i (0)η j (0)ζ m (0) (0) − (0) .
∂xi ∂xj ∂xk p
R : Tp M × Tp M × Tp M → Tp M
(ξ, η, ζ) 7→ R(ξ, η)ζ
l
∂Γljk ∂Γlik
Rkij (0) = (0) − (0)
∂xi ∂xj
Remark 3.1.6. One can check (not difficult but tedious) that we have in arbitrary coor-
dinates
n
l
∂Γljk ∂Γlik X
Rkij = − + (Γm l m l
kj Γmi − Γki Γmj )
∂xi ∂xj
m=1
81
3 Curvature
(1) R : Tp M × Tp M × Tp M → Tp M is trilinear;
Proof.
(1) is obvious because ∇2ξ,η ζ is already R-linear in ξ, η and ζ.
(3) We choose Riemannian normal coordinates around p and consider the special case
∂ ∂ ∂ ∂
ξ= , η= , ζ= , ν= .
∂xi p ∂xj p ∂xk p ∂xl p
Then we find
! !
∂ ∂ ∂ ∂
g|p R(ξ, η)ζ, ν = g|p R , ,
∂xi p ∂x
j
p ∂xk p ∂x
l
p
n
!
X
m ∂ ∂
= g|p Rkij (0) ,
∂xm p ∂xl p
m=1
n
!
X
m ∂ ∂
= Rkij (0) · g|p ,
∂xm p ∂x
l
p
m=1
Xn
m
= gml (0) · Rkij (0).
m=1
82
3.1 The Riemannian curvature tensor
n
∂ 2 gij X ∂Γm
ki
∂Γmkj
(0) = gmj (0) (0) + gmi (0) (0) .
∂xk ∂xl m=1
∂xl ∂xl
Thus
∂ 2 gij ∂ 2 gij
0= (0) − (0)
∂xk ∂xl ∂xl ∂xk
n
X ∂Γmki
∂Γmkj ∂Γm
li
∂Γm
lj
= gmj (0) (0) + gmi (0) (0) − gmj (0) k (0) − gmi (0) k (0)
m=1
∂xl ∂xl ∂x ∂x
n
X
m m
= gmj (0)Rilk (0) + gmi (0)Rjlk (0)
m=1
n
X
m m
0= gml (0)Rkij (0) + gmk (0)Rlij (0)
m=1
and therefore
n
X n
X
m m
gml (0)Rkij (0) = − gmk (0)Rlij (0).
m=1 m=1
This proves the assertion for coordinate fields ξ, η, ζ, ν of Riemannian normal coor-
dinates.
l l l
Rkij + Rijk + Rjki = 0.
l l l
Rkij (0) + Rijk (0) + Rjki (0)
= 0.
83
3 Curvature
Example 3.1.8. Let (M, g) = (Rn , geucl ) or (Rn , gMink ). In Cartesian coordinates we have
Γkij = 0. Thus we get Rkij
l = 0 for all i, j, k, l and therefore R ≡ 0.
Warning. In the literature there are two different sign conventions for R: For example,
our R is the negative of the curvature tensor as defined in [ON83].
Lemma 3.1.11
Let (M, g) and (M̃ , g̃) be semi-Riemannian manifolds and ψ : M → M̃ a local isometry. Let
p ∈ M . Then the curvature tensors R of M at p and R̃ of M̃ at ψ(p) are related by:
dψ|p R(ξ, η)ζ = R̃ dψ|p (ξ), dψ|p (η) dψ|p (ζ)
for all ξ, η, ζ ∈ Tp M .
84
3.1 The Riemannian curvature tensor
Since ψ is a local isometry, it follows that gij = g̃ij : V → R, where the gij are the
components of g w.r.t. x and the g̃ij are the components of g̃ w.r.t. x̃. Therefore the
Christoffel symbols coincide, Γkij = Γ̃kij , hence so do the components of the curvature
l
tensors, Rkij l . From (3.4) we conclude
= R̃kij
X n ∂
∂ ∂ ∂ l
dψ R , = R kij dψ
∂xi ∂xj ∂xk ∂xl
l=1
n
X
l ∂
= Rkij
∂ x̃l
l=1
n
X
l ∂
= R̃kij
∂ x̃l
l=1
∂ ∂ ∂
= R̃ ,
∂ x̃i ∂ x̃j ∂ x̃k
∂ ∂
∂
= R̃ dψ , dψ dψ .
∂xi ∂xj ∂xk
∂
This proves the lemma for the coordinate basis tangent vectors ∂xi
. By trilinearity of R
it follows for all tangent vectors.
Alternatively one can define the curvature tensor as a multilinear map R : Tp M ×Tp M ×
Tp M × Tp M → R by
R(ξ, η, ζ, ν) = g R(ξ, η)ζ, ν .
In this version, R is known as the Riemannian (4, 0)-curvature tensor. In local coordinates
x : U → V around p, we define Rijkl : V → R by
!
∂ ∂ ∂ ∂
Rijkl (x(p)) := R , , , .
∂xi p ∂xj p ∂xk p ∂xl p
Then we have
Prop.
3.1.7(5)
Rijkl = Rklij
∂ ∂ ∂ ∂
= g R , ,
∂xk ∂xl ∂xi ∂xj
n
!
X
m ∂ ∂
= g Rikl ,
m=1
∂xm ∂xj
n
X
m ∂ ∂
= Rikl g ,
∂xm ∂xj
m=1
Xn
m
= gmj Rikl .
m=1
85
3 Curvature
Hence we have
n
X
m
Rijkl = gmj Rikl
m=1
hence
n
X
l
Rkij = gal Rkaij
a=1
Proposition 3.1.12
Let (M, g) be a semi-Riemannian manifold. In Riemannian normal coordinates we have:
n
1 X
gij (x) = εi δij + Rikjl (0)xk xl + O ||x||3 .
3
k,l=1
Proof. We already know that gij (x) = εi δij + O (||x||2 ) by Corollary 2.6.33. In the fol-
lowing we will use the Einstein summation convention and the following abbreviations
∂f ∂2f
f,k := and f,kℓ =
∂xk ∂xk ∂xℓ
for the first and the second partial derivatives. In the proof of Theorem 2.3.8 we have
seen that
gij,k = Γm m
ki gmj + Γkj gmi .
We differentiate this equation with respect to xℓ , evaluate at 0 and use that the Christof-
fel symbols vanish at 0:
gij,kℓ (0) = Γm m
ki,ℓ (0) · gmj (0) + Γkj,ℓ (0) gmi (0) . (3.5)
Claim:
Γkij,ℓ (0) + Γkℓi,j (0) + Γkjℓ,i (0) = 0 . (3.6)
86
3.1 The Riemannian curvature tensor
Proof of the claim: In normal coordinates the straight lines t 7→ t · x give geodesics. The
equation for geodesics then looks like:
0 = Γkij (t · x) xi xj .
d
0= Γkij (tx) xi xj = Γkij,ℓ (0) xℓ xi xj .
dt t=0
Γkij,ℓ (0) + Γkℓi,j (0) + Γkjℓ,i (0) + Γkji,ℓ (0) + Γkiℓ,j (0) + Γkℓj,i (0) = 0.
The symmetry of the Christoffel symbols in their two lower indices implies the claim.✓
ℓ (0) = Γℓ ℓ
From Rkij jk,i (0) − Γik,j (0) we conclude:
Rkℓij (0) = Γm m
jk,i (0) − Γik,j (0) gmℓ (0)
(3.6)
= − Γm m m
ij,k (0) + Γki,j (0) + Γik,j (0) gmℓ (0)
= − Γm m
ij,k (0) + 2Γki,j (0) gmℓ (0) . (3.7)
Thus we have:
Prop. 3.1.7
2Rikjℓ (0) xk xℓ = (−Rkijℓ (0) − Rℓjik (0)) xk xℓ
(3.7)
= Γm m k ℓ
jℓ,k (0) + 2Γkj,ℓ (0) gmi (0) x x
+ Γm m k ℓ
ik,ℓ (0) + 2Γℓi,k (0) gmj (0) x x
(∗)
= Γm m k ℓ
jℓ,k (0) + 2Γkj,ℓ (0) gmi (0) x x
+ Γm m ℓ k
iℓ,k (0) + 2Γki,ℓ (0) gmj (0) x x
(3.5)
= gij,ℓk (0) + 2gij,kℓ (0) · xk xℓ
= 3gij,kℓ (0) · xk xℓ .
At equality (∗) we renamed the summation parameter k to ℓ and vice versa. Thus we
get for the second term in the Taylor expansion
1 1
gij,kℓ (0) xk xℓ = Rikjℓ (0) · xk xℓ .
2 3
87
3 Curvature
Lemma 3.2.2
For two-dimensional subvector spaces E ⊂ V the following assertions are equivalent:
Proof. With respect to any basis ξ, η of E, the bilinear form h·, ·i |E×E is represented by
the matrix
hξ, ξi hη, ξi
Aξ,η := .
hξ, ηi hη, ηi
Then we have Q(ξ, η) = det Aξ,η which proves the lemma.
88
3.2 Sectional curvature
Remark 3.2.3
(a) If h·, ·i is positive definite, then
p
Q(ξ, η) = area of the parallelogram spanned by ξ and η.
(b) The two-dimensional subvector space E ⊂ V is degenerate if and only if there exists
a basis ξ, η of E with hξ, ξi = hξ, ηi = 0. Namely,
“⇐”: Q(ξ, η) = hξ, ξi hη, ηi − hξ, ηi2 = 0.
| {z } | {z }
=0 =0
“⇒”: Let E be degenerate, i.e., h·, ·i |E×E is degenerate. Then there exists a
ξ ∈ E \ {0} with hξ, ζi = 0 for all ζ ∈ E. Now complete ξ by some η to
a basis of E. ✓
Example 3.2.4
Let V = R3 with the Minkowski product
x0
hhξ, ηii = −ξ 0 η 0 + ξ 1 η 1 + ξ 2 η 2 .
b
d
⇒ 0= hhc(t), c(t)ii = 2 hhċ(0), c(0)ii = 2 hhξ, pii
dt t=0
⇒ Tp C = p⊥
⇒ Tp C is degenerate.
89
3 Curvature
p
b b b
p⊥
Lemma 3.2.5
Let V be a finite-dimensional real vector space with non-degenerate symmetric bilinear
form h·, ·i. Let R : V × V × V × V → R be multilinear with
for all ξ, η, ζ, ν ∈ V . Then for E ∈ G2 (V, h·, ·i) and any basis ξ, η of E the expression
R(ξ, η, η, ξ)
K(E) :=
Q(ξ, η)
does not depend on the choice of the basis ξ, η of E, but only on E itself.
90
3.2 Sectional curvature
[
Set G2 (M, g) := G2 (Tp M, g|p ).
p∈M
R(ξ, η, η, ξ)
K(E) := ,
Q(ξ, η)
where ξ, η is a basis of E, is called sectional curvature of (M, g). Here R is the Rie-
mannian (4, 0)-curvature tensor.
Remark 3.2.7. The sectional curvature is only defined for manifolds of dimension at
least 2. If dim(M ) = 1, then R(ξ, η, ζ, ν) = 0 for all ξ, η, ζ, ν ∈ Tp M due to the skew-
symmetry in ξ and η.
Remark 3.2.9. The sectional curvature determines the curvature tensor, as can be seen
by
91
3 Curvature
Special case: If K(E) only depends on p but not on the particular plane E ⊂ Tp M (satis-
fied automatically if dim(M ) = 2, but not in general if dim(M ) ≥ 3), then:
R(ξ, η, ζ, ν) = K(p) hη, ζi hξ, νi − hξ, ζi hη, νi .
92
3.3 Ricci- and scalar curvature
The local description of Ricci curvature is similar to that of the semi-Riemannian metric
itself: For any chart x : U → V of M we define the functions
∂ ∂
ricij : V → R, ricij (x(p)) := ric , .
∂xi p ∂xj p
n
X
k
(iii) We have: ricij = Rikj .
k=1
Proof. (i) Bilinearity of ric follows directly from trilinearity of R. We show symmetry
of ric:
n
X
ric(η, ξ) = εi · g R(η, Ei )Ei , ξ
i=1
Prop.
Xn
3.1.7(5)
= εi · g R(Ei , ξ)η, Ei
i=1
Prop.
3.1.7 n
X
(2),(3)
= εi · g R(ξ, Ei )Ei , η
i=1
= ric(ξ, η).
93
3 Curvature
∂ ∂
∂
∂
(iii) We fix i and j and we have ricij = ric ∂x i , ∂xj = tr ζ 7→ −R ∂x i , ζ ∂xj . W.r.t.
∂
the basis ∂x∂ 1 , . . . , ∂x∂ n the endomorphism ζ 7→ −R ∂x ∂
i , ζ ∂xj has the matrix rep-
resentation
l l
(−Rjik )kl = (Rjki )kl .
Pn k
Thus we get that ricij = k=1 Rjki and because of (i) we have ricij = ricji , which
yields the assertion.
We defined Ricci curvature using the Riemann curvature tensor. Since the curvature
tensor and sectional curvature contain the same information, Ricci curvature should
also be computable in terms of sectional curvature. Indeed, Ricci curvature can can be
computed by averaging the sectional curvature of certain planes.
Lemma 3.3.4
Let (M, g) be a semi-Riemannian manifold and p ∈ M . If ξ ∈ Tp M with g(ξ, ξ) 6= 0 and if
E2 , . . . , En is a generalized orthonormal basis of ξ ⊥ , then
n
X E3
ric(ξ, ξ) = g(ξ, ξ) · K(span{ξ, Ej })
j=2
| {z } ξ
This is essentially the
mean value of K on all
planes containing ξ.
E2
94
3.3 Ricci- and scalar curvature
Remark 3.3.5. Lemma 3.3.4 expresses ric(ξ, ξ) in terms of sectional curvatures provided
g(ξ, ξ) 6= 0. Since g is non-degenerate the set of vectors ξ ∈ Tp M with g(ξ, ξ) 6= 0 is
dense in Tp M . By continuity, ric(ξ, ξ) is determined for all ξ ∈ Tp M . By polarization,
this determines the values of ric(ξ, η) for all ξ, η ∈ Tp M via
1
ric(ξ, η) = 2 ric(ξ +η, ξ +η) − ric(ξ, ξ) − ric(η, η) .
for all ξ, η ∈ Tp M .
We compute:
∂ ∂
ricij = ric ,
∂xi ∂xj
∂ ∂
= g Ric , j
∂xi ∂x
n
!
k ∂ ∂
X
= g Rici k , j
∂x ∂x
k=1
n
X
k ∂ ∂
= Rici · g , .
∂xk ∂xj
k=1
We have shown:
n
X
ricij = Ricki · gkj
k=1
The functions ricij are obtained from the functions Ricki by lowering the upper index.
Similarly, the Ricki can be obtained from the ricij by raising one index.
95
3 Curvature
scal(p) := tr(Ric|p )
Lemma 3.3.8
(i) In local coordinates we have
n
X n
X
scal(p) = Ricii x(p) = ricij x(p) · gij x(p) .
i=1 i,j=1
Proof. Clear.
Remark 3.3.9. Let us consider the special case when dim(M ) = 2. Let K be the Gauß
curvature, i.e., K(p) = K(Tp M ). Then the curvature tensor is given by
R(ξ, η, ζ, ν) = K(p) g(η, ζ)g(ξ, ν) − g(ξ, ζ)g(η, ν) .
2
X
ric(ξ, η) = εi · R(ξ, Ei , Ei , η)
i=1
2
X
= K(p) εi g(Ei , Ei )g(ξ, η) − g(ξ, Ei )g(Ei , η)
i=1
= K(p) 2g(ξ, η) − g(ξ, η)
= K(p) · g(ξ, η).
96
3.4 Jacobi fields
This shows
ric = K · g
and
scal = 2K.
In the case of surfaces the Riemann curvature tensor, sectional curvature (Gauß curva-
ture), Ricci curvature and scalar curvature all determine each other. In higher dimen-
sions this is no longer so.
Remark 3.3.10. The following table shows how the different notions of curvature de-
pend on each other:
dim M 2 3 ≥4
R
m m m
K
m m ⇓
ric
m ⇓ ⇓
scal
Remark 3.3.11. In the physics literature the following notation in local coordinates is
often used:
l
• for R and R ones writes: Rijk and Rijkl (as here),
• for Ric and ric one write: ricij = Rij and Ricji = Rij ,
t 7→ cs (t) := c(s, t)
is a geodesic.
97
3 Curvature
∂
Let ξ(t) := c(0, t) be the corresponding variational vector field. Then we have:
∂s
2
∇ ∇∇ ∂
ξ(t) = c(s, t)|s=0
dt ∂t ∂t ∂s
∇∇ ∂
= c(s, t)|s=0
∂t ∂s ∂t
∇ ∇ ∂ ∂c ∂c ∂c
= c(s, t) |s=0 + R (0, t), (0, t) (0, t)
∂s |∂t ∂t{z } ∂t ∂s ∂t
≡0 since cs geodesic
The above computation shows that the variational vector field of a geodesic variation
is a Jacobi field.
Proposition 3.4.3
Let M be a n-dimensional semi-Riemannian manifold, c : I → M a geodesic and t0 ∈ I.
For all ξ, η ∈ Tc(t0 ) M there exists a unique Jacobi field J along c with
∇
J(t0 ) = ξ and J(t0 ) = η.
dt
The set of Jacobi fields along c forms a 2n-dimensional vector space.
98
3.4 Jacobi fields
∇ 2
Pn j (t)E (t)
Then dt J(t) = j=1 v̈ j and
n
X
R(ċ(t), J(t))ċ(t) = v j (t)R(ċ(t), Ej (t))ċ(t).
j=1
Pn k
Write R(ċ(t), Ej (t))ċ(t) = k=1 aj (t)Ek (t). Then J is a Jacobi field if and only if
n
X n
X
k
v̈ Ek = akj v j Ek ,
k=1 j,k=1
This is a linear system of ordinary differential equations of second order. Thus solutions
exist (on all of I) and are uniquely determined by the initial data v k (t0 ) and v̇ k (t0 ), i.e.,
∇
by J(t0 ) and dt J(t0 ).
The linearity of the Jacobi equation implies that its solution space forms a vector space.
∇
The map {Jacobi fields} → Tc(t0 ) M ⊕ Tc(t0 ) M , J 7→ (J(t0 ), dt J(t0 )) is a vector space
isomorphism. In particular, the dimension of the space of Jacobi fields along c equals
2n.
Example 3.4.4. If M is flat then the equation for Jacobi fields is simply given by
2
∇
J ≡ 0.
dt
Hence
{Jacobi fields} = ξ(t) + t η(t) | ξ, η parallel .
99
3 Curvature
∇
implies dt J, ċ ≡ 0 and from
d ∇
hJ, ċi = J, ċ ≡ 0
dt dt
we see that hJ, ċi ≡ 0.
Consequence. Let c be non light-like. In this case we have Tc(t) M = Rċ(t)⊕ ċ(t)⊥ . Then
∇
{Jacobi fields along c} = R · ċ ⊕ R · t ċ ⊕ {Jacobi fields J along c | J ⊥ ċ, dt J ⊥ ċ} .
| {z } | {z }
uninteresting interesting Jacobi fields
Jacobi fields
Remark 3.4.8. For light-like geodesics c this is not true because ċ ⊥ ċ.
Example 3.4.9
c
Let (M, g) = (R2 , gMink ), let c be a light-like geodesic and
let ξ be a light-like parallel vector field along c which is
linearly independent of ċ. ξ
Since ξ is parallel and R = 0, the vector field ξ is also a C
Jacobi field and we have:
{Jacobi field along c} = R · ċ ⊕ R(tċ) ⊕ R ξ ⊕ R(tξ)
| {z }
∇
= Jacobi field J along c | J ⊥ ċ, dt J ⊥ ċ
100
3.4 Jacobi fields
respectively.
κ s2κ + c2κ = 1,
s′κ = cκ and sκ (0) = 0,
c′κ = −κ sκ and cκ (0) = 1.
Example 3.4.11. Let (M, g) be a Riemannian manifold with constant sectional curvature
K ≡ κ. Let c be a geodesic, parametrized by arc-length. Let ξ be a parallel vector field
along c with ξ ⊥ ċ. Set
J(t) := a sκ (t) + b cκ (t) ξ(t) with a, b ∈ R.
Then
2
∇
J = (a s̈κ + b c̈κ )ξ = −κ(a sκ + b cκ )ξ = −κJ.
dt
For the curvature tensor we here have R(ξ, η)ζ = κ(hη, ζi ξ − hξ, ζi η). Thus
R(ċ, J)ċ = (a sκ + b cκ ) · κ hξ, ċi ċ − hċ, ċi ξ = −κ a sκ + b cκ ξ = −κJ.
| {z } | {z }
=0 =1
n ∇ o
Jacobi fields along c | J ⊥ ċ, J ⊥ ċ
dt
= (a sκ + b cκ ) ξ | a, b ∈ R, ξ parallel along c, ξ ⊥ ċ .
101
3 Curvature
κ>0 J
b b b b
κ=0
J
J
b b
κ<0
Proposition 3.4.12
Let M be a semi-Riemannian manifold and c : [a, b] → M a geodesic. Let ξ be a smooth
vector field along c. Then
102
3.4 Jacobi fields
c(0, t) = expγ(0) (t − t0 )η(0) = expc(t0 ) (t − t0 )ċ(t0 ) = c(t)
∂c
Hence c(s, t) is a geodesic variation of c(t). Let J(t) := ∂s (0, t) be the corresponding
variational field. Then J is a Jacobi field. We show:
∇ ∇
ξ(t0 ) = J(t0 ) and ξ(t0 ) = J(t0 ).
dt dt
Then we get ξ = J because Jacobi fields are uniquely determined by their initial data
and hence ξ is the variational field of the geodesic variation c(s, t).
We calculate
∂c d d
J(t0 ) = (0, t0 ) = expγ(s) (0) = γ(s) = γ̇(0) = ξ(t0 )
∂s ds s=0 ds s=0
and
∇ ∇ ∂c ∇ ∂c ∇ ∇
J(t0 ) = (0, t0 ) = (0, t0 ) = η(0) = η2 (0) = ξ(t0 ).
dt ∂t ∂s ∂s ∂t ds dt
We are now able to generalize Lemma 2.6.26 and can identify the differential of the
exponential at arbitrary points in its domain.
Proposition 3.4.13
Let M be a semi-Riemannian manifold, p ∈ M and ξ ∈ Tp M . We assume that the geodesic
γ(t) := expp (tξ) is defined on [0, 1], i.e., ξ lies in the domain of expp .
For η ∈ Tp M (∼ ∇
= Ttξ Tp M ) let J be the Jacobi field along γ with J(0) = 0 and dt J(0) = η.
Then we have for all t ∈ (0, 1]:
J(t)
d expp |tξ (η) = .
t
∂c
Let ζ := ∂s |s=0 be the corresponding variational Jacobi field. Then we have
∂c d
ζ(0) = (0, 0) = expp |s=0 (0) = 0 = J(0)
∂s ds
and
103
3 Curvature
Tp M
0
b
η
ξ
expp
p ξ
M
∇ ∇ ∂c ∇ ∂c ∇ ∇
ζ(0) = (0, 0) = (0, 0) = (ξ + sη)|s=0 = η = J(0)
dt dt ∂s ∂s ∂t ds dt
Hence ζ = J. Now we compute for fixed t ∈ (0, 1]:
∂ ∂ s 1 1
d expp |tξ (η) = expp (tξ + s η)|s=0 = expp t ξ + η = ζ(t) = J(t).
∂s ∂s t s=0 t t
Corollary 3.4.14
Let M be a semi-Riemannian manifold, let p ∈ M and let ξ be in the domain of expp . Then
ker(d expp |ξ ) ∼
= Jacobi field along γ(t) = expp (tξ) | J(0) = 0, J(1) = 0 .
104
3.4 Jacobi fields
Case 2: κ > 0 .
For a geodesic parametrized by arc-length, the ker d expp |ξ
conjugate points belonging to t0 are the points pb
ξ
t0 + m √πκ for m ∈ Z \ {0}. Considering the case ξ′
m = 1 we have Tp S n
expp {ξ ∈ Tp S n | ||ξ|| = π} = {−p}.
−pb
Sn
For ||ξ|| = π we obtain
ker d expp |ξ = ξ ⊥ .
Proposition 3.4.18
Let M be a semi-Riemannian manifold and let c : [t0 , t1 ] → M be a geodesic. Let t0 and t1
be not conjugated with each other along c.
Then for ξ ∈ Tc(t0 ) M and η ∈ Tc(t1 ) M there exists exactly one Jacobi field J along c with
J(t0 ) = ξ and J(t1 ) = η.
ξ
J
c(t1 )
b b b
c(t0 )
η
105
3 Curvature
2n-dimensional (n + n)-dimensional
z }| { z }| {
{Jacobi field along c} → Tc(t0 ) M ⊕ Tc(t1 ) M ,
J 7→ J(t0 ), J(t1 ) ,
is injective since t0 and t1 are not conjugated to each other along c. For dimensional
reasons, this map is an isomorphism.
Proposition 3.4.18 means that in the non-conjugate case we can also characterize Jacobi
fields by the boundary values J(t0 ) and J(t1 ) instead of the initial values J(t0 ) and
∇
dt J(t0 ). In the conjugate case this is certainly wrong.
{η = α · ċ(π) | α ∈ R}.
b −p = c(t1 )
106
4 Submanifolds
4.1 Submanifold of differentiable manifolds
x {0} × Rm−n
N V ⊂ Rn
U
x(p)
p b
b
Rn × {0}
M
Example 4.1.2
(1) Codimension n = 0: A subset N ⊂ M is a submanifold of codimension 0 if and
only if N is open subset of M .
is a submanifold chart.
107
4 Submanifolds
U1 × U2 p Γx2 ◦f ◦x1 −1
b
U2 x 1 × x 2 b
Ψ b
M2
M1 U1 V1 × V2
Theorem 4.1.3
Let M be an m-dimensional differentiable manifold. Let N ⊂ M be a subset. Then the
following assertions are equivalent:
(ii) For every p ∈ N there exists an open neighborhood U of p and smooth functions
f1 , . . . , fm−n : U → R such that
(a) N ∩ U = {q ∈ U | f1 (q) = . . . = fm−n (q) = 0};
(b) The differentials df1 |p , . . . , dfm−n |p ∈ Tp∗ M are linearly independent.
108
4.1 Submanifold of differentiable manifolds
ϕ̃(q)
ϕ̃ ◦ f ◦ ϕ−1
V
Choose a chart ϕ : U → V of M around p and a chart ϕ̃ : Ũ → Ṽ of R around q := f (p).
W.l.o.g. we assume that f (U ) ⊂ Ũ . Since ϕ and ϕ̃ are diffeomorphisms, we have
The implicit function theorem yields: One can shrink V and U to smaller neighbor-
hoods of q and p, respectively, such that V = V1 × V2 and one can find a smooth map
g : V1 → V2 such that
If we compose ϕ with a submanifold chart for graphs as in the Example 4.1.2 (4) then
we get a submanifold chart for N in M around p.
109
4 Submanifolds
4
are f (0) = 0 and f ( 32 ) = − 27 .
The examples indicate that there may be many critical points but there are always only
few critical values. This is true in general:
Corollary 4.1.9
If f : M → R is smooth and if q ∈ R is a regular value of f , then N = f −1 (q) is empty or a
submanifold of M with codim(N ) = dim(R).
110
4.1 Submanifold of differentiable manifolds
Df |x = (2x0 , . . . , 2xn ).
1, x 6= 0
⇒ rank(Df |x ) =
0, x = 0
Remark 4.1.11. In this example all q ∈ R \ {0} are regular values. We have
n √
−1 S ( q), q > 0
f (q) =
∅ ,q<0
For the critical value q = 0 we have that f −1 (0) = {0} is also (by coincidence) a subma-
nifold, but of the wrong codimension n + 1. In general, the preimage of a critical value
is not a submanifold.
Remark 4.1.12. Sometimes the set f −1 (q) is a submanifold with codimension dim R
even if q is a critical value.
Example 4.1.13. Let g : Rn+1 → R, g(x) = (||x||2 − 1)2 . Then 0 is critical value of g but
S n = g−1 (0) is a submanifold of codimension 1.
(x1 , . . . , xn ) : U ∩ N → V ∩ Rn
111
4 Submanifolds
Theorem 4.1.15
Let N ⊂ M be a submanifold. Let ι : N ֒→ M be the inclusion map, ι(p) = p. Then we
have:
(ii) The function f |N = f ◦ ι is the composition of two smooth maps and therefore
again smooth.
Remark 4.1.16. One identifies Tp N with dι|p (Tp N ) and thinks of it as a vector subspace
of Tp M .
Tp M
b Tp N
p
N
M
112
4.2 Semi-Riemannian submanifolds
(ḡ|p ) Tp M ×Tp M
=: g|p
is non-degenerate.
Example 4.2.3. Let (M̄ , ḡ) = (R2 , gMink ) be 2-dimensional Minkowski space, i.e.,
gMink = −dx0 ⊗ dx0 + dx1 ⊗ dx1 . Then
N1 = {(x0 , 0) | x0 ∈ R} is semi-Riemannian (with negative-definite metric).
x0 M̄ + M̄
N4 b b
−
N2 b
x1 b
− b b
+
N3 N1
113
4 Submanifolds
Tp M̄
b Tp M
p
Np M
M
M̄
tan :Tp M̄ → Tp M,
nor :Tp M̄ → Np M,
be the orthogonal projections. Both M and also M̄ have, when seen as semi-Riemannian
¯ respectively. Now we
manifolds in their own rights, a Levi-Civita connection ∇ and ∇,
want to investigate, how we can determine ∇ directly from ∇.¯
114
4.2 Semi-Riemannian submanifolds
¯ ξ η := ∇
∇ ¯ ξ η̄.
Example 4.2.6 η
Let M̄ = (R2 , geucl ) and M = S1. Set
η(x , x ) = (−x , x ). For c : R → S 1 with c(t) =
1 2 2 1
ċ(t) = η(c(t)).
¯ ηη
∇
Then we get:
¯
¯ η η = ∇ ċ = c̈ = − cos(t), − sin(t)
∇
dt
which is not tangent to S 1 .
¯ ξ η).
We set ∇ξ η := tan(∇
Theorem 4.2.7
Let (M̄ , ḡ) be a semi-Riemannian manifold and M ⊂ M̄ a semi-Riemannian submanifold
with induced semi-Riemannian metric g. Let ∇ ¯ be the Levi-Civita connection of (M̄ , ḡ).
Then
∇ξ η = tan ∇¯ ξη
115
4 Submanifolds
Proof. We check that ∇ satisfies the axioms of the Levi-Civita connection for (M, g). By
the uniqueness statement in Theorem 2.3.8, ∇ must then be the Levi-Civita connection
of (M, g).
¯ is local.
(i) Locality is clear because ∇
¯ is linear in ξ.
(ii) Linearity in ξ is clear because tan is linear and ∇
(v) Product rule II: Let ξ ∈ Tp M and η1 , η2 ∈ C ∞ (U, T M ). Choose smooth extensions
η̄1 , η̄2 ∈ C ∞ (Ū , T M̄ ). Then
∂ξ g(η1 , η2 ) = ∂ξ ḡ η̄1 , η̄2
= ḡ|p ∇¯ ξ η̄1 , η̄2 |p + ḡ|p η̄1 |p , ∇ ¯ ξ η̄2
|{z}
=η2 |p , in particular tangent to M
= g|p tan ∇ ¯ ξ η̄1 , η2 |p + g|p η1 |p , tan ∇ ¯ ξ η̄2
= g|p ∇ξ η1 , η2 |p + g|p η1 |p , ∇ξ η2 .
116
4.2 Semi-Riemannian submanifolds
Lemma 4.2.9
Let ξ ∈ Tp M and η ∈ C ∞ (U, T M ), where U ⊂ M is an open neighborhood of p. Then
nor(∇¯ ξ η) ∈ Np M only depends η via η|p .
¯ ξ η) = nor(∇
nor(∇ ¯ ξ η̄)
¯ ξ η̄ − ∇ξ η
= ∇
m m̄ k m̄
X X ∂ η̄ X ∂
= ξi
i
+ Γ̄kij |x(p) · η̄ j |x(p)
∂x x(p) ∂xk p
i=1 k=1 j=1
m m k m
∂η ∂
X X X
− ξi i
+ Γkij |x(p) · η j |x(p)
∂x x(p) ∂xk p
i=1 k=1 j=1
m m m̄ m
!
X X X ∂ X ∂
= ξi η j |x(p) Γ̄kij |x(p) − Γkij |x(p) .
∂xk p ∂xk p
i=1 j=1 k=1 k=1
117
4 Submanifolds
Lemma 4.2.11
The second fundamental form II is bilinear and symmetric.
Clearly, II is bilinear. By the symmetry of the Christoffel symbols in the lower indices,
II is also symmetric.
Notation 4.2.13. For better readability we will from now on write hξ, ηi instead of
g(ξ, η) or ḡ(ξ, η).
Since one can compute the Levi-Civita connection ∇ of the submanifold M from the
Levi-Civita connection ∇¯ of M̄ , one should also be able to compute the curvature tensor
R of M from that of M̄ . Indeed this is possible.
118
4.2 Semi-Riemannian submanifolds
* +
¯ ζ∇
¯ ∂ ∂ ¯¯ ∂ ¯ ¯ ∂ ¯¯ ∂
R̄(ζ, ν)ξ, η = ∇ i
−∇ ∇ ∂ | ∂ l ∂xi − ∇ν ∇ ∂ k ∂xi + ∇∇ ∂
i
,η
∂xl ∂x p ∂xk ∂x
∂ |
k p ∂x ∂x
∂x ∂xl
¯ ¯ ∂ ¯ ¯ ∂
torsion freeness = ∇ζ ∇ ∂ − ∇ν ∇ ∂ ,η
∂xl ∂x
i ∂xk ∂x
i
¯ ∂ ¯ ∂ ∂ ¯ ∂ ¯ ∂ ∂
= ∇ζ ∇ ∂ i
+ ∇ζ II , − ∇ν ∇ ∂ − ∇ν II , ,η
∂xl ∂x ∂xl ∂xi ∂xk ∂x
i ∂xk ∂xi
∂ ∂
= ∇ζ ∇ ∂ i
− ∇ν ∇ ∂ i
,η
∂xl ∂x ∂xk ∂x
≡0
z
}| {
∂ ∂ ∂ ∂ ∂ ¯ ∂
+ ∂ζ II , , j − II , , ∇ζ j
∂xl ∂xi ∂x ∂xl ∂xi ∂x
≡0
z
}| {
∂ ∂ ∂ ∂ ∂ ¯ ∂
− ∂ν II , , j + II , , ∇ν j
∂xk ∂xi ∂x ∂xk ∂xi ∂x
= hR(ζ, ν)ξ, ηi + hII(ζ, ξ), II(ν, η)i − hII(ν, ξ), II(ζ, η)i .
Corollary 4.2.15
If E ⊂ Tp M is a non-degenerate plane with basis ξ, η, then we have
Proof. This follows directly from the definition of sectional curvature and the Gauß
formula.
Lemma 4.2.16
Let M ⊂ M̄ be a semi-Riemannian submanifold. Let ϕ : M̄ → N̄ be a local isometry. Set
119
4 Submanifolds
¯
∇ ∇
ξ= ξ + II(ξ, ċ).
dt dt
In particular, we have for ξ = ċ
¯
∇ ∇
ċ = ċ + II(ċ, ċ).
dt dt
Therefore the curve c is a geodesic in M if and only if
¯
∇ ¯
∇
ċ = II(ċ, ċ), i.e., if ċ(t) ∈ Nc(t) M for all t ∈ I.
dt dt
¯
∇
ċ(t) = c̈(t) = − cos(t) · p − sin(t) · ξ = −c(t) ∈ Nc(t) S n .
dt
Hence c is a geodesic in S n .
120
4.3 Totally geodesic submanifolds
Theorem 4.3.3
For a semi-Riemannian submanifold M ⊂ M̄ the following statements are equivalent:
(iii) For any p ∈ M and ξ ∈ Tp M there exists an ε > 0 such that the M̄ -geodesic
c : (−ε, ε) → M̄ with c(0) = p and ċ(0) = ξ lies in M , i.e., c(t) ∈ M for all
t ∈ (−ε, ε).
(iv) Let c : I → M be a smooth curve. Then the parallel transport along c w.r.t. M and
w.r.t. M̄ coincide (for tangent vectors of M ).
Proof. “(ii)⇒(iii)”: Let p ∈ M and ξ ∈ Tp M . Let c be the M̄ -geodesic with c(0) = p and
ċ(0) = ξ. Let c̃ be the M -geodesic with c̃(0) = p and c̃(0)˙ = ξ. By (ii), c̃ is also geodesic
˙
in M̄ . Since we have c̃(0) = c(0) and c̃(0) = ċ(0), the two M̄ -geodesics must coincide,
c = c̃ on (−ε, ε) for a ε > 0. In particular, c lies in M .
“(iii)⇒(i)”: Let p ∈ M and ξ ∈ Tp M . Let cξ be the M̄ -geodesic with cξ (0) = p and
ċξ (0) = ξ. By (iii), cξ lies in M for t ∈ (−ε, ε) with suitable ε > 0. On (−ε, ε) we get:
¯
∇ ∇
0= ċξ = ċξ + II(ċξ , ċξ )
dt dt
|{z} | {z }
normal
tangential
In particular, we have
II ċξ (t), ċξ (t) = 0 for all t ∈ (−ε, ε).
121
4 Submanifolds
⇒ c ist geodesic in M̄ .
Remark 4.3.6. Most semi-Riemannian manifolds M̄ do not have totally geodesic sub-
manifolds of dimension m ∈ {2, . . . , m̄ − 1}.
Theorem 4.3.7
Let M ⊂ M̄ be a semi-Riemannian submanifold. Assume that there exists an isometry
ϕ ∈ Isom(M̄ ), such that M is a connected component of
Proof. We check Criterion (iii) in Theorem 4.3.3. Let p ∈ M and ξ ∈ Tp M . We first show
that
dϕ|p (ξ) = ξ.
Namely, let γ : J → M be a smooth curve with γ(0) = p and γ̇(0) = ξ. Then
d
dϕ|p (ξ) = dϕ|p (γ̇(0)) = (ϕ ◦ γ) |t=0 = γ̇(0) = ξ. ✓
dt | {z }
= γ , since
M ⊂Fix(ϕ)
122
4.4 Hypersurfaces
⇒ ϕ := A|S n ∈ Isom(S n )
⇒ Fix(ϕ) = W ∩ S n is totally geodesic
The Gauß Formula (Theorem 4.2.14) tells us that if M ⊂ M̄ is totally geodesic, then
4.4 Hypersurfaces
Remark 4.4.2. For ε = +1 we have Index(M̄ , ḡ) = Index(M, g) while for ε = −1 we get
Index(M̄ , ḡ) = Index(M, g) + 1.
123
4 Submanifolds
Caution! This does not define a norm unless h·, ·i is definite. In particular, it can occur
that |ξ| = 0 even if ξ 6= 0.
n
X ∂f
df = dxi .
∂xi
i=1
n n
!
∂f X ∂f i ∂ ∂ X
i ∂ ∂
= dx = df = g α ,
∂xj ∂xi ∂xj ∂xj ∂xi ∂xj
i=1 i=1
n n
X ∂ ∂ X
= αi g , = αi gij .
∂xi ∂xj
i=1 i=1
n
i
X ∂f
Matrix multiplication with (gij )ij yields α = gij , thus
∂xj
j=1
n
X ∂f ij ∂
gradf = g
∂xj ∂xi
i,j=1
Lemma 4.4.4
Let M̄ be a semi-Riemannian manifold and f : M̄ → R smooth and c ∈ R be a regular value
of f . Then M := f −1 (c) ⊂ M̄ is a semi-Riemannian hypersurface of signature ε, if
gradf |p
Moreover, we have ν := ∈ Np M and hν, νi = ε.
|gradf |p |
124
4.4 Hypersurfaces
Proof. Since c is a regular value, M is a hypersurface. The lemma follows once we show
gradf |p ⊥ Tp M.
d
0= f γ(t) |t=0 = df |p (ξ) = hgradf |p , ξi .
dt | {z }
≡c
Definition 4.4.5 Np M
Let M ⊂ M̄ be a semi-Riemannian hypersurface ν
and p ∈ M . Let ν ∈ Np M with |ν| = 1.
bc
p
M
The linear map Sν : Tp M → Tp M , characterized
by
hSν (ξ), ηi = hII(ξ, η), νi for all ξ, η ∈ Tp M,
Lemma 4.4.6
The Weingarten map Sν is self-adjoint.
Remark 4.4.7. We have S−ν = −Sν . Without specifying the choice of ν, the Weingarten
map is only determined up to a sign.
125
4 Submanifolds
Lemma 4.4.8
Let M ⊂ M̄ be a semi-Riemannian hypersurface and ν
p ∈ M . Let U ⊂ M be an open neighborhood of p and bc
Gauß formula:
Let M ⊂ M̄ be a semi-Riemannian hypersurface with signature ε. Let ξ, η, ζ ∈ Tp M .
Then:
R(ξ, η)ζ = R̄(ξ, η)ζ + ε hSν (η), ζi Sν (ξ) − hSν (ξ), ζi Sν (η) .
In particular, all ḡij are constant. Hence all Christoffel symbols vanish in Cartesian
coordinates. Therefore the curvature vanishes:
R̄ ≡ 0, K̄ ≡ 0, ric ≡ 0 and scal ≡ 0.
126
4.4 Hypersurfaces
of (Rn+1 , ḡ) is called the pseudo-sphere of radius r and with index k. The semi-
Riemannian hypersurface
n
Hk−1 (r) := f −1 (−r 2 ),
127
4 Submanifolds
Example 4.4.10. Let k = 0 and ḡ = geucl . Example 4.4.11. The case k = 1 and ḡ =
Then S0n (r) = S n (r) is the standard gMink is also of great interest.
sphere of radius r.
f −1 (0)
bc
H0n (r)
r
S1n (r)
S n (r)
Definition 4.4.12. The hypersurface H n := {x ∈ H0n (1) | x0 > 0} together with the
induced Riemannian metric ghyp is called the n-dimensional hyperbolic space.
Definition 4.4.13. The hypersurface S14 (r) together with the induced Lorentzian met-
ric is called deSitter spacetime and H14 (r) is called anti-deSitter spacetime.
Remark 4.4.14. The pseudo-sphere Skn (r) is diffeomorphic to Rk × S n−k while the
pseudo-hyperbolic space Hkn (r) is diffeomorphic to S k × Rn−k . See the exercises or
[ON83, page 111] for a proof of this fact.
hence
n
gradf |x gradf |x 1X i ∂
ν|x := p = = x .
|4c| 2r r ∂xi
i=0
128
4.4 Hypersurfaces
ε ε
R(ξ, η)ζ = (hη, ζi ξ − hξ, ζi η) and K≡
r2 r2
We compute
n
X
ric(ξ, η) = εi hR(ξ, ei )ei , ηi
i=1
n
ε X
= εi hei , ei i ξ − hξ, ei i ei , η
r2 | {z }
i=1 =εi
ε
= (n hξ, ηi − hξ, ηi),
r2
thus
ε(n − 1) ε(n − 1)n
ric = g and scal =
r2 r2
Remark 4.4.15. For the Einstein tensor of S14 (r) or H14 (r) we get
1 3ε 1ε·3·4 ε
G = ric − scal · g = 2 g − g = −3 2 g.
2 r 2 r2 r
Thus deSitter and anti-deSitter spacetime are vacuum solutions of the Einstein field
equations with cosmological constant Λ = r32 and Λ = − r32 , respectively.
129
4 Submanifolds
E
E
M M M
Pk−1 j j Pn
Here hx, yi = − j=0 x y + j=k xj y j . We have O(n + 1, 0) = O(n + 1) and O(n, 1) is
the Lorentz group. For any A ∈ O(n + 1 − k, k) we have
O(n + 1 − k, k) → Isom(M ),
A 7→ A|M .
Proposition 4.4.16
Let M be a semi-Riemannian manifold, let p ∈ M and ϕ, ψ ∈ Isom(M ) with ϕ(p) = ψ(p)
and dϕ|p = dψ|p .
Then ϕ and ψ coincide on the set of all points which can be joined with p by a geodesic.
130
4.4 Hypersurfaces
ϕ(p)
ċ(0)
˙
and c̃(0) ˙
= dϕ|p (ċ(0)) = dψ|p (ċ(0)) = ĉ(0). Therefore b p
c̃ = ĉ. In particular, ϕ(q) = c̃(1) = ĉ(1) = ψ(q).
M
Corollary 4.4.17
If all points of M can be joined by geodesics with p, then every isometry ϕ of M is uniquely
determined by ϕ(p) and dϕ|p .
Remark 4.4.19. The assumption that the points can be joined with p by geodesics is
necessary for the statement of Corollary 4.4.17.
131
4 Submanifolds
p
Put
(x, y) on M +
ϕ1 := id, and ϕ2 (x, y) :=
(−x, y) on M − M−
Both ϕ1 and ϕ2 are isometries. Now ϕ1 (p) = ϕ2 (p) and
dϕ1 |p = dϕ2 |p but ϕ1 6= ϕ2 .
Lemma 4.4.22
On M = Skn (r) (where 0 ≤ k ≤ n − 1), on M = Hkn (r) (where 1 ≤ k ≤ n) and on
M = H n (r) any two points can be joined by a geodesic.
Theorem 4.4.23
Restriction yields isomorphisms
Isom(Skn (r)) ∼
= O(n + 1 − k, k) for 0 ≤ k ≤ n − 1
Isom(H (r)) ∼
n
k = O(n − k, k + 1) for 1 ≤ k ≤ n
Isom(H (r)) ∼
n
= SO(n, 1) := {A ∈ O(n, 1) | A00 > 0}.
132
4.5 Trigonometry in spaces of constant curvature
Proof. Put M = Skn (r), M = Hkn (r) or M = H n (r) and G = O(n + 1 − k, k), G =
O(n − k, k + 1), or G = SO(n, 1), respectively. We need to show: Every isometry of M
is of the form
ϕ = A|M with A ∈ G.
a) We first show that G acts transitively on M . This means that for all p, q ∈ M there
exists an A ∈ G with Ap = q.
Namely: W.l.o.g. let p = r · e0 = (r, 0, . . . , 0)T . From hq, qi = ±r 2 we see that b0 := 1r q
satisfies hb0 , b0 i = ±1. We extend b0 to a generalized eigenbasis b0 , b1 , . . . , bn of Rn+1 .
Now A := (b0 , b1 , . . . , bn ) ∈ G and Ap = r Ae0 = rb0 = q.
b) Next we show: For any linear isometry B : Tp0 M → Tp0 M where p0 = re0 , there exists
an A ∈ G such that ϕ = A|M satisfies ϕ(p0 ) = p0 and dϕ|p0 = B. Namely:
1 0...0
0
A := ...
B
0
does the job.
c) Let now ϕ ∈ Isom(M ). Put p1 := ϕ(p0 ) where p0 = re0 . By a) there exists an A1 ∈ G
with A1 p0 = p1 . Hence ψ := A−1
1 |M ◦ ϕ is an isometry of M with ψ(p0 ) = p0 .
Moreover, B := dψ|p0 : Tp0 M → Tp0 M is a linear isometry. By b) there is an A2 ∈ G
such that χ := A2 |M satisfies dχ|p0 = B. Lemma 4.4.22 and Corollary 4.4.17 imply
χ = ψ. Thus
ϕ = A1 |M ◦ ψ = A1 |M ◦ χ = A1 |M ◦ A2 |M = (A1 ◦ A2 ) |M .
| {z }
∈G
133
4 Submanifolds
Thus Mnκ is an n-dimensional Riemannian manifold with the constant sectional cur-
vature κ.
Remark 4.5.2. Since for any given three points there exists a two-dimensional subma-
nifold of Mnκ which contains these points, it suffices to consider the case n = 2.
Lemma 4.5.3
For every κ ∈ R, the isometry group of Mκ contains the subgroup
Gκ := ϕ | ϕ = A|Mκ where A ∈ GL(3) with hAx, Ayiκ = hx, yiκ ,
1 1
hAx, Ayiκ = hx, yiκ ∀ x, y ∈ R3 and A(Mκ ) = Mκ .
κ κ
Remark 4.5.4. In the case κ 6= 0 the conditions hAx, Ayiκ = hx, yiκ and κ1 hAx, Ayiκ =
1
κ hx, yiκ are of course equivalent and we could omit one of them. But in the case κ = 0
we need both of them.
134
4.5 Trigonometry in spaces of constant curvature
From hAx, Ayiκ = hx, yiκ it already follows that A(M̂κ ) = M̂κ . In the case κ ≤ 0, A
could possibly exchange the two connected components of M̂κ . This is ruled out by the
condition A(Mκ ) = Mκ . In the case κ > 0 we could omit this condition.
Proof of the Lemma. Let A ∈ Gκ . Since ϕ = A|Mκ is the restriction of the linear map A,
we get that for p ∈ Mκ the differential dϕ(p) : Tp Mκ → Tϕ(p) Mκ also is the restriction of
A,
dϕ(p) = A|Tp Mκ .
Here, the tangent spaces of Mκ are viewed as subvector spaces of R3 . Since A respects
the bilinear form κ1 h·, ·iκ , the differential dϕ(p) is a linear isometry for every p ∈ Mκ .
Thus ϕ is an isometry of Riemannian manifolds.
Remark 4.5.5. Indeed, we have Isom(Mκ ) = Gκ but we will not need this fact.
For κ = 1 we have
Gκ ∼
= {A ∈ GL(3) | hAx, Ayi = hx, yi ∀x, y ∈ R3 } = O(3)
In case κ = 0, we have:
G0 = {A ∈ GL(3) | hAx, Ayi0 = hx, yi0 , 01 hAx, Ayi0 = 1
0 hx, yi0 ∀ x, y ∈ R3 , AM0 = M0 }
1 0 0
1 1 2
= A= b b , b ∈ R, B ∈ O(2)
B
b2
This holds true since for any A ∈ G0 ,
x0 y 0 = (Ax)0 (Ay)0 = (A00 x0 + A01 x1 + A02 x2 )(A00 y 0 + A01 y 1 + A02 y 2 )
Thus
0 2 A(M0 )=M0
For x = y =e0 : 1 = (A0 )
⇒ A00 = ±1 ⇒ A00 =1.
135
4 Submanifolds
Mκ ∩ E, b
0
E
where E ⊂ R3 is a two-dimensional subvector space.
Lemma 4.5.6
The geodesics parametrized by arc-length γ : R → Mκ with γ(0) = e0 are then given by
cκ (r)
γ(r) = sκ (r) · sin(ϕ)
sκ (r) · cos(ϕ)
where ϕ ∈ R is fixed.
136
4.5 Trigonometry in spaces of constant curvature
1
= κ κ2 sκ (r)2 + κ cκ (r)2 sin(ϕ)2 + cκ (r)2 cos(ϕ)2
= κ sκ (r)2 + cκ (r)2
= 1.
The generalized sine and cosine functions allow us to explicitly write down many
isometries in Gκ .
Example 4.5.7. Rotations about the e0 -Axis are isometries,
1 0 0
0 cos(ϕ) − sin(ϕ) ∈ Gκ
0 sin(ϕ) cos(ϕ)
for any ϕ and any κ ∈ R. Using κ s2κ + c2κ = 1 one easily checks that
cκ (r) −κ sκ (r) 0
sκ (r) cκ (r) 0 ∈ Gκ
0 0 1
for all r ∈ R. In the case κ = 1 this is a rotation about the e2 -axis. For κ = 0 this is
the identity, hence uninteresting. In the case κ = −1 such isometries are called Lorentz
boosts. Similarly, one sees that
cκ (r) 0 κ sκ (r)
Lr := 0 1 0 ∈ Gκ .
sκ (r) 0 −cκ (r)
Before using these isometries we observe that
cκ (r)
Lr e0 = 0
sκ (r)
and
cκ (r) cκ (r) 0 κsκ (r) cκ (r)
Lr 0 = 0 1 0 0
sκ (r) sκ (r) 0 −cκ (r) sκ (r)
2 2
cκ (r) + κsκ (r)
= 0
sκ (r)cκ (r) − cκ (r)sκ (r)
1
= 0
0
= e0 .
137
4 Submanifolds
Definition 4.5.8.
C
b
Let M be a Riemannian manifold. A geodesic triangle γA
is a 6-tupel γB
(A, B, C, γA , γB , γC ),
b
B
where A, B, C ∈ M are pairwise disjoint points, γA , γB b
γC
and γC geodesic segments with endpoints B and C, C A
and A or A and B, respectively.
The points A, B and C are the vertices, the geodesic segments γA , γB and γC are the
sides of the geodesic triangle. The angle at a vertex is defined to be the angle of the
tangent vectors of the sides at that vertex.
Cb
γ a
Let (A, B, C, γA , γB , γC ) a geodesic triangle in Mκ . The sides b
have the lengths a, b and c, respectively, and the angles are
denoted by α, β and γ, respectively. β b
α B
b
c
A
Here the length of a geodesic segment γ is defined as the length of the parameter in-
terval × the norm of γ̇, which is constant. A more general definition of the length of a
differentiable curve in a Riemannian manifold will be introduced later. Since the isom-
etry group of Mκ acts transitively, we can assume w.l.o.g. that
1
A = e0 = 0 .
0
Applying an isometry of the form
1 0 0
0 cos(ϕ) − sin(ϕ)
0 sin(ϕ) cos(ϕ)
(rotation about the e0 -axis) we can rotate B in the e0 -e2 -plane without moving A = e0 .
The formula from Lemma 4.5.6 for the geodesic γC with ϕ = 0 and r = c then tells us
cκ (c)
B= 0
sκ (c)
138
4.5 Trigonometry in spaces of constant curvature
γ b
B = Lc A
Hence the isometry Lc interchanges the points A and
a
B and we obtain a new geodesic triangle. On the one
hand one can compute Lc C similarly as C itself and α b
one obtains β b
c
A = Lc B
cκ (a)
Lc C = sκ (a) sin(β) .
On the other hand sκ (a) cos(β)
cκ (c) 0 κsκ (c) cκ (b)
Lc C = 0 1 0 sκ (b) sin(α)
sκ (c) 0 −cκ (c) sκ (b) cos(α)
cκ (c)cκ (b) + κsκ (c)sκ (b) cos(α)
= sκ (b) sin(α)
sκ (c)cκ (b) − cκ (c)sκ (b) cos(α)
Thus we obtain the equations:
cκ (a) = cκ (c)cκ (b) + κ sκ (c)sκ (b) cos(α) (Law of Cosines) (1)
sκ (a) sin(β) = sκ (b) sin(α)
sκ (a) sκ (b)
= (Law of Sines) (2)
sin(α) sin(β)
sκ (a) cos(β) = sκ (c)cκ (b) − cκ (c)sκ (b) cos(α) (3)
139
4 Submanifolds
cos(γ) = cκ (c) sin(α) sin(β) − cos(α) cos(β) (Cosine Rule for Angles).
We have proved
Theorem 4.5.9
Let κ ∈ R. For a geodesic triangle in Mκ with the side lengths a, b, c and the angles α, β, γ
we have
Now we analyze the sum of angles in the model space of constant curvature.
Theorem 4.5.10
Let κ ∈ R. For the sum of angles α + β + γ of a geodesic triangle in Mκ with the inner angles
0 < α, β, γ < π we have
> π, if κ> 0
α+β+γ = π, if κ= 0
< π, if κ< 0
140
4.5 Trigonometry in spaces of constant curvature
b b b
b b b
b b b
Proof. W.l.o.g. we assume that α ≥ β. For this proof we will use the notation “⋚” for
“<”, if κ > 0, for “=”, if κ = 0, and for “>”, if κ < 0. We have −κ ⋚ 0, for instance.
If is κ > 0, then Mκ is the sphere of radius √1κ . Thus in this case the side lengths have
2π
to be < √
κ
. In the case κ ≤ 0, we do not have any bounds on the side lengths. We use
the convention √1 = ∞, if κ ≤ 0. With this convention we have in all cases
κ
cκ ⋚ 1
2π
in the interval (0, √κ
). Since sin is positive on (0, π) the Cosine Rule for Angles yields
First case: π − (β + γ) ≥ 0.
Since cos is strictly monotonically decreasing on [0, π], the relation
cos(α) ⋚ cos(π − (β + γ)) yields π − (β + γ) ⋚ α and thus π ⋚ α + β + γ. This
is what we wanted to show.
Remark 4.5.11. Since the inner angles are < π, we always have for the sum of angles in
a geodesic triangle α + β + γ < 3π. It is easy to see that for Mκ with κ > 0 the sum of
141
4 Submanifolds
angles of a geodesic triangle can take all values in (π, 3π). For Mκ with κ < 0 all values
of the interval (0, π) occur.
142
5 Riemannian Geometry
From now on we concentrate on Riemannian geometry, that is, on semi-Riemannian
manifolds whose metric is positive definite and hence defines a Euclidean scalar prod-
uct on each tangent space. One special feature of the Riemannian case is that each
connected Riemannian manifold naturally becomes a metric space.
Remark 5.1.2. The length does not depend on the parametrization of the curve.
Namely, if ϕ : [a, b] → [α, β] is a parameter transformation, then we have
Zb
d
L[c ◦ ϕ] = (c ◦ ϕ)(t) dt
dt
a
Zb
= ||ċ(ϕ(t))|| · |ϕ̇(t)| dt
a
Zβ
Substitution
s = ϕ(t) = ||ċ(s)|| ds
α
= L[c].
143
5 Riemannian Geometry
Remark 5.1.4. The infimum is, in general, not a minimum. In other words, there need
not exist a shortest curve connection p and q.
Example 5.1.5 Rn
b
p
M = Rn \ {0} and p = −q. We have d(p, q) = 2 ||p||, but 0 bc
0 b
tξ
{tξ | 0 ≤ t ≤ b} ⊂ Tp M and we have
expp
(i) d expp |tξ (ξ) = γ̇(t).
d expp |tξ (η)
(ii) For η ∈ Ttξ Tp M ∼
= Tp M we have b
γ b
γ̇(t)b
p γ(t)
d expp |tξ (η), γ̇(t) = hη, ξi .
M
In particular, d expp |tξ (η) ⊥ γ̇(t), if η ⊥ ξ.
d d
Proof. (i) We compute d expp |tξ (ξ) = expp (t ξ + s ξ)|s=0 = γ(t + s)|s=0 = γ̇(t).
ds ds
(ii) By (i) it suffices to consider the case η ⊥ ξ. Let J be the Jacobi field along γ with
∇
J(0) = 0 and dt J(0) = η. Proposition 3.4.13 yields
J(t)
d expp |tξ (η) = for t > 0.
t
144
5.1 The Riemannian distance function
∇
Since both J and dt J are perpendicular to γ̇ at t = 0, this holds for all t. We conclude
J(t)
d expp |tξ (η), γ̇(t) = , γ̇(t) = 0 = hη, ξi .
t
x
Φ : Tp M \ {0} −→ (0, ∞) × S n−1 , x = t · y 7→ (t, y) = (kxk, ),
kxk
where S n−1 ⊂ Tp M is the unit sphere in the tangent space. There exists an r > 0, such
that expp maps B(0, r) ⊂ Tp M diffeomorphically onto a neighborhood U of p in M .
Then the map
Tp M
0 S n−1
b
Φ
B(0, r) ≈
expp U1
p
b
0 r
U
M
The Gauß lemma says that in such coordinates the Riemannian metric takes the form
1 0 ··· 0
0
...
(gij ) =
∗
0
145
5 Riemannian Geometry
Corollary 5.1.7
Let r > 0 so small that expp |B̄(0,r) is a diffeomorphism on its image. Let c : [a, b] → M be a
piecewise C 1 -curve with c(a) = p and c(b) 6∈ expp (B(0, r)). Then L[c] ≥ r.
Proof. Let β ∈ (a, b) be minimal such that c(β) ∈ ∂ expp (B(0, r)) = expp (S n−1 (r)). Let
α ∈ [a, β) maximal such that c(α) = p. Now it is ensured that for τ ∈ (α, β) the curve
c(τ ) lies in expp (B(0, r)) \ {p}. For τ ∈ (α, β] we write
ξ(q) := d expp ξ˜ expp −1 (q) .
d dt dy
c̃(τ ) = · y(τ ) +t(τ ) · (τ )
dτ dτ |{z} |dτ{z }
= ξ̃(c̃(τ ))
⊥ ξ̃(c̃(τ ))
D E D E
˙ ) = dt .
ξ c(τ ) , ċ(τ ) = d expp ξ˜ c̃(τ ) , d expp c̃(τ
˙ ) = ξ̃ c̃(τ ) , c̃(τ
dτ
146
5.1 The Riemannian distance function
Thus we get
L[c] ≥ L[c|[α,β] ] Tp M
Zβ 0 ξ̃
b
= ||ċ(τ )|| dτ
B(0, r)
α
Cauchy
Zβ expp
Schwarz ≥ ξ c(τ ) , ċ(τ ) dτ
inequality
α ξ
Zβ p
b
dt U
= dτ
dτ
α M
= t(β) − t(α)
= r − 0 = r.
Theorem 5.1.8
(M, d) is a metric space.
Proof. a) Obviously we have d(p, q) ≥ 0 and d(p, p) = 0 because the constant curve has
length 0. Now let p 6= q. We have to show d(p, q) > 0. Choose r > 0 such that expp |B(0,r)
is a diffeomorphism and q 6∈ expp (B(0, r)). Then by Corollary 5.1.7 every curve from p
to q has length r at least. Hence d(p, q) ≥ r > 0.
b) Symmetry d(p, q) = d(q, p) is clear. Simply traverse the curves in the opposite direc-
tion.
c) It remains to show the triangle inequality d(p, q) ≤ d(p, r) + rd(r, q). b
Let ε > 0. Choose a continuous piecewise C 1 -
curves c1 from p to r with L[c1 ] ≤ d(p, r) + ε and c2
c2 from r to q with L[c2 ] ≤ d(r, q) + ε. Now concate- c1
nate c1 and c2 to a continuous piecewise C 1 -curve c b
q
from p to q. Then we have b
p
d(p, q) ≤ L[c] = L[c1 ] + L[c2 ] ≤ d(p, r) + ε + d(r, q) + ε.
147
5 Riemannian Geometry
Remark 5.1.12. For 0 < r < injrad(p) we have expp (B(0, r)) = B(p, r). Namely:
“⊂”: Let q = expp (ξ) with ||ξ|| < r. Then t 7→ expp (tξ), t ∈ [0, 1], is a curve from p to q
with length ||ξ|| < r. Hence d(q, p) < r, i.e., q ∈ B(p, r).
“⊃”: Corollary 5.1.7.
Theorem 5.1.13
The metric d induces the original topology on M .
148
5.1 The Riemannian distance function
Proof. For the moment we denote the open subsets w.r.t. d of M as “d-open”. We have
to show: d-open = open.
a) Claim: Every d-open set is open.
Let U ⊂ M be d-open. For every p ∈ U there exists a r(p) > 0, such that B(p, r(p)) ⊂ U .
W.l.o.g. let r(p) < injrad(p). Then B(p, r(p)) = expp (B(0, r(p))) is the diffeomorphic
| {z }
open in Tp M
S
image of an open subset of Tp M . Hence it is open itself. Therefore U = B p, r(p)
p∈M
is the union of open subsets of M and thus open.
b) Claim: Every open set is d-open. The proof is similar.
Corollary 5.1.14
The map d : M × M → R is continuous.
If Φ ∈ Isom(M ).
Remark 5.1.15. Then we have L[Φ ◦ c] = L[c] and thus also
d Φ(p), Φ(q) = d(p, q).
Rb
Recall that E[c] = 1
2 ||ċ(t)||2 dt is the energy of c.
a
Proposition 5.1.16
Let M be a Riemannian manifold and c : [a, b] → M be a continuous, piecewise C 1 -curve.
Then we have
L[c]2 ≤ 2(b − a) · E[c].
Proof. With the Cauchy Schwarz inequality for the L2 -scalar product we obtain:
b 2
Z Zb Zb
2
L[c]2 = ||ċ(t)|| · 1 dt ≤ ||ċ(t)|| dt · 12 dt = 2E[c] · (b − a).
a a a
149
5 Riemannian Geometry
Equality holds if and only if ||ċ|| and 1 (as a function) are linearly dependent. This means
that ||ċ|| is constant, i.e., that c is parametrized proportionally to arc-length.
Corollary 5.1.17
A curve c minimizes the energy in the set of all continuous piecewise C 1 -curves connecting p
and q if and only if c minimizes the length and is parametrized proportionally to arc-length.
Corollary 5.1.19
Every shortest curve from p to q is a geodesic up to parametrization. It is a geodesic if and
only if it is parametrized proportionally to arc-length.
Caution! The converse is not true. Not every geodesic is a shortest curve connecting its
endpoints.
Example 5.1.20. Great circles on S n are geodesics connecting points to themselves. But
the only shortest curves connecting a point to itself are constant curves which have
length 0.
Definition 5.1.21. A geodesic γ : [a, b] → M with L[γ] = d γ(a), γ(b) is called mini-
mal.
150
5.2 Completeness
5.2 Completeness
General Assumption. Throughout this section let M be a connected Riemannian ma-
nifold.
(3) The closed balls B̄(p, r) are compact for all r > 0.
(4) The closed balls B̄(q, r) are compact for all r > 0 and all q ∈ M .
(5) (M, d) is complete as a metric space, i.e., all d-Cauchy sequences converge.
Definition 5.2.5. If the equivalent conditions (1)–(5) in Theorem 5.2.2 hold, then one
calls M a complete Riemannian manifold.
151
5 Riemannian Geometry
Corollary 5.2.6
Every compact Riemannian manifold is complete.
Proof of Corollary 5.2.6. We check condition (3) in the Hopf-Rinow theorem. Indeed,
B̄(p, r) ⊂ M is a closed subset of the compact space M and thus compact itself.
Proof of Theorem 5.2.2. We will prove this theorem in five steps. The structure of the
proof is as follows:
(a) trivial (e)
(5) (2) (1) (6)
| {z }
(b) (d)
(4) (3)
(c)
1. Claim: The limit point q does not depend on the choice of the sequence (ti )i∈N with
i→∞
ti −→ β.
Proof. If (t′i )i∈N is another such sequence with q ′ = limi→∞ γ(t′i ), then also (t′′i )i∈N is
such a sequence, where
′′ tj , i = 2j
ti :=
t′j , i = 2j + 1
The sequence (γ(t′′i ))i∈N is a d-Cauchy-sequence with accumulation points q and q ′ .
We thus have q = q ′ . This proves the first claim. ✓
152
5.2 Completeness
Proof. Let x : U → V be a chart of M around q with x(q) = 0. Choose r > 0 such that
B̄(0, r) ⊂ V . Since B̄(0, r) is compact, there exist constants C1 , C2 , C4 > 0 with
n
P
• ||a||max ≤ C2 aj ∂x∂ j x−1 (y) for all a = (a1 , . . . , an ) ∈ Rn and y ∈ B̄(0, r).
j=1
g
∂Γkij
• ∂xl
(y) ≤ C4 for all y ∈ B̄(0, r).
Choose ε > 0 small enough so that γ(t) ∈ x−1 (B̄(0, r)) for t ∈ (β − ε, β). Write
γ k := xk ◦ γ and ak := γ̇ k . By the equations for geodesics we obtain:
n
X n
X
ȧk = γ̈ k = − Γkij (γ 1 , . . . , γ n ) · γ̇ i γ̇ j = − Γkij (γ 1 , . . . , γ n ) ai aj
i,j=1 i,j=1
and hence
ȧk ≤ n2 · C1 · ||a||2max .
This implies
ȧ max
≤ n2 C1 · ||a||2max ≤ n2 C1 · C2 2 ||γ̇||g 2 = n2 C1 C2 2 =: C3 .
| {z }
=1
We get
Ztj Ztj
||a(ti ) − a(tj )||max = ȧ(t) dt ≤ ||ȧ(t)||max dt ≤ C3 |ti − tj | .
ti max ti
Thus the a(ti ) form a Cauchy sequence in Rn and hence converge to some A ∈ Rn .
i→∞
As before A is independent of the special choice of the sequence (ti )i∈N with ti −→
β. Thus we obtain a continuous extension of a by
(
a(t), t ∈ (β − ε, β)
ā(t) :=
A, t=β
Hence the velocity field γ̇ is extended continuously to t = β. This shows that the
extension γ̄ of γ is C 1 . ✓
n n
!
X X ∂Γkij
äk = − al ai aj + 2 Γkij ȧi aj
∂xl
i,j=1 l=1
153
5 Riemannian Geometry
This implies
As before we see that (ȧ(ti ))i∈N forms a d-Cauchy-sequence in Rn . This shows that
the extension γ̄ is even a C 2 -curve. By continuity it satisfies the geodesic equation
also at t = β.
˙
Now let γ̂ : (β − δ, β + δ) → M be the geodesic with γ̂(β) = γ̄(β) and γ̂(β) ˙
= γ̄(β).
Since geodesics are uniquely determined by their initial values, γ̂ and γ̄ coincide on
their common domain of definition. This yields a continuation of γ as a geodesic on
(α, β + δ). This contradicts the maximality of β and thus shows (2).
b) Let all closed balls in M be compact. Let (pi )i∈N be a Cauchy sequence in M . Since
Cauchy sequences are bounded, there exists a R > 0 such that pi ∈ B̄(p, R) for all i ∈
N. Since B̄(p, R) is compact, the Cauchy sequence (pi )i∈N has an accumulation point
q ∈ B̄(p, R). Since accumulation points of Cauchy sequences are unique, (pi )i∈N
converges to q.
c) Let all B̄(p, r) be compact for all r > 0. Let q ∈ M
and let R > 0. Set r := R + d(p, q). Then
q R b
p
because for x ∈ B̄(q, R) we have
b
b
γ̇i (0) b
b
compact we have, after passing to a suitable subsequence, b b
..
p .
X b
i→∞ q
γ̇i (0) −→ X ∈ Sn − 1(1) ⊂ Tp M.
B̄(p, r)
i→∞
The ti lie in the compact interval [0, r]. After passing to a subsequence again, ti −→
T ∈ [0, r] converges too. Set q := expp (T · X). This definition is possible because of
154
5.2 Completeness
εk ց 0. b
b
b
q̄ γ
erwise we are finished. Choose 0 < r0 < b
q
injrad(p). Then B̄(p, r0 )
S(p, r0 ) = expp S n−1 (r0 )
is compact. Let qk be the first intersection point of ck with S(p, r0 ). After passing to
a subsequence, qk possesses a limit q̄ ∈ S(p, r0 ). We have
d(p, q) ≤ d(p, qk ) + d(qk , q) ≤ L[ck ] ≤ d(p, q) + εk
k→∞
⇒ d(p, q) ≤ d(p, q̄) + d(q̄, q) ≤ d(p, q)
⇒ d(p, q) = d(p, q̄) + d(q̄, q)
Let γ be the unique minimal geodesic that connects p with q̄. We parametrize γ by
arc-length. With (1) we can extend γ to [0, d(p, q)].
It remains to show that γ : [0, d(p, q)] → M is a minimal geodesic from p to q. Set
I := t ∈ [0, d(p, q)] | d(p, γ(t)) = t and d(p, γ(t)) + d(γ(t), q) = d(p, q) .
We have seen that [0, r0 ] ⊂ I. Set t0 := sup(I). We have to show that t0 = d(p, q)
because then
d γ(t0 ), q = d(p, q) − d γ(t0 ), p = d(p, q) − t0 = 0,
γ b
q
exists a q̄ ′ ∈ ∂B(q ′ , r1 ) with
b
γ(t ) = q ′ 0
155
5 Riemannian Geometry
2.6.5) says:
b
c
Rb cs
d ∇
ds E[cs ]|s=0 = − ξ, dt ċ dt + hξ, ċi |ba .
a
ċ(t+
c(ti+1 )
If cs is continuous and only piecewise that is, C 2, i )
there exists a partition a = t0 < t1 < · · · < tN = b, c
such that (s, t) 7→ cs (t) is continuous on (−ε, ε)×[a, b] b
c(ti )
b
N
d d X
E[cs ]|s=0 = E[cs |[ti−1 ,ti ] ]|s=0
ds ds
i=1
N Z ti !
X ∇ − +
= − ξ, ċ dt + ξ(ti ), ċ(ti ) − ξ(ti−1 ), ċ(ti )
ti−1
dt
i=1
Z b N
∇ X
= − ξ, ċ dt + ξ(b), ċ(b− ) − ξ(a), ċ(a+ ) + ξ(ti ), ċ(t− +
i ) − ċ(ti )
a dt
i=1
d
Question: If c is a continuous and only piecewise C 2 -curve with ds E[cs ]|s=0 = 0 for all
continuous, piecewise C 2 -variations cs with fixed endpoints, does c then have to be a
156
5.3 The second variation of the energy
If ċ(t− +
i ) 6= ċ(ti ) for an i ∈ {1, . . . , N − 1} then we can choose an η ∈ Tc(ti ) M with
η, ċ(t− +
i ) − ċ(ti ) > 0.
d
0= E[cs ]|s=0 = ξ(ti ), ċ(t− + − +
i ) − ċ(ti ) = η, ċ(ti ) − ċ(ti ) > 0.
ds
This is a contradiction. We summarize:
Theorem 5.3.1
Let M be a semi-Riemannian manifold and c : [a, b] → M be a continuous, piecewise C 2 -
curve. Then for every continuous piecewise C 2 -variation cs of c with variational field ξ we
have
Zb b N −1
d ∇ X
E[cs ] =− ξ, ċ dt + hξ, ċi + ξ(ti ), ċ(t− +
i ) − ċ(ti ) ,
ds s=0 dt a
a i=1
where a = t0 < t1 < · · · < tN = b is a partition for which both c and cs are C 2 on the
intervals [ti−1 , ti ], i = 1, . . . N .
The curve c is a geodesic if and only if for all such variations with fixed endpoints we have
d
E[cs ] = 0.
ds s=0
To investigate the minima of the energy, we have to consider the second variation of
the energy.
157
5 Riemannian Geometry
Zb
d2 ∇ ∇
E[cs ] = ξ, ξ − hR(ξ, ċ)ċ, ξi dt.
ds2 s=0 dt dt
a
the diameter of M .
158
5.4 The Bonnet-Myers theorem
Example 5.4.2. For M = S n equipped with the standard metric g = gstd we have
diam(S n ) = π. For M = Rn with the Euclidean metric g = geucl and for hyperbolic
space M = H n with g = ghyp we have diam(Rn ) = diam(H n ) = ∞.
Namely:
Example 5.4.5
(1) Let M = S n with g = α2 · gstd where α is a positive constant. Then we have
1 n−1
diam(M ) = α π, K≡ , ric ≡ g
α2 α2
π 1
⇒ diam(M ) = √ and ric = κ(n − 1)g with κ = .
κ α2
This shows that the estimate in the Bonnet-Myers theorem is optimal and cannot be
improved.
(2) Now let M = RPn with g = gstd . Since RPn is locally isometric to S n , we have as for
the sphere ric = (n − 1)g. But diam(RPn ) = π2 . Here we find diam(M ) < √πκ where
κ = 1.
159
5 Riemannian Geometry
Proof of Theorem 5.4.4. Let p, q ∈ M with p 6= q. Set δ := d(p, q). Since M is complete,
there exists a minimal geodesic from p to q by the Hopf-Rinow theorem. W.l.o.g. let
γ : [0, δ] → M be parametrized by arc-length with γ(0) = p and γ(δ) = q.
Let e ∈ Tp M with e ⊥ γ̇(0) and ||e|| = 1. Let e(t) be the vector field along γ obtained
from e by parallel transport. Set
π
ξ(t) := sin t · e(t). ξ
δ
Let γs (t) be a variation of γ with fixed endpoints and vari- b
p
Since γ is a minimal geodesic, we have
d
0= E[γs ]|s=0
ds
and d2
0 ≤ 2 E[γs ]|s=0
ds
Zδ !
∇ 2
= ξ − hR(ξ, γ̇)γ̇, ξi dt
dt
0
Zδ π π 2
π 2
= cos t e(t) − sin t hR(e, γ̇)γ̇, ei dt
δ δ δ
0
Zδ
π2 π 2 π 2
= cos t · 1 − sin t K(e, γ̇) dt.
δ2 δ δ
0
Zδ π 2
π2 π 2
≤ (n − 1) cos t − sin t · κ dt
δ2 δ δ
0
1 π 2 − κδ2
= (n − 1) · .
2 δ
Therefore 0 ≤ π 2 − κδ2 and hence δ ≤ √πκ . Since this holds for all choices of p and q we
conclude
π
diam(M ) ≤ √ .
κ
160
5.4 The Bonnet-Myers theorem
The theorem tells us that the larger the Ricci curvature of a Riemannian manifold, the
smaller the manifold.
Note that the following general implications hold:
Thus the Bonnet-Myers theorem also holds if the sectional curvature is bounded from
below by a postive constant, K ≥ κ > 0. Does the Bonnet-Myers theorem also hold
under the weaker condition scal ≥ n(n − 1)κ?
The answer is “no” as we see by the following counterex- M2
ample. If M1 and M2 are Riemannian manifolds and if b
p
(p1 , p2 ) 2
b
RM1 (ξ1 , η1 ) 0
⇒ RM (ξ 1 + ξ2 , η1 + η2 ) = M
0 R (ξ2 , η2 )
2
M ricM1 0
⇒ ric =
0 ricM2
⇒ scalM = scalM1 + scalM2 .
For n ≥ 3 we obtain with M = S n−1 × R that
but diam(M ) = ∞. Thus the Bonnet-Myers theorem does not hold under the weaker
condition scal ≥ n(n − 1)κ if n ≥ 3.
For n = 2 on the other hand, the three conditions in (1) are equivalent.
161
Bibliography
[M65] J. M ILNOR : Topology from the differential viewpoint. University Press of Virginia,
Charlottesville 1965.
[ON83] B. O’N EILL : Semi-Riemannian geometry. Academic Press, New York 1983.
163
Index
165
Index
166
Index
167
Index
tangent bundle, 31
tangent space, 18
tangent vector, 17
topological space, 1
topology, 1
torsion freeness, 50, 56
torsion-free, 48
totally geodesic submanifold, 121
transformation of principal axes, 36
triangle, geodesic, 138
variation of a curve, 64
variation of the energy, second, 157
variation of the energy, first, 64
variation, geodesic, 97
variational vector field, 64
vector field, 33
vector field along a map, 54
168