240 Bnotes
240 Bnotes
240 Bnotes
Riemannian Geometry
Simon Rubinstein–Salzedo
Winter 2006
0.1 Introduction
These notes are based on a graduate course on differentiable manifolds and Rieman-
nian geometry I took from Professor Doug Moore in the Winter of 2006. The textbook
was Riemannian Geometry by Manfredo Perdigão do Carmo. Many other books are
also mentioned in the notes. Since the professor handed out very good notes, I have
made very few changes to these notes.
1
Chapter 1
h·, ·ip : Tp M × Tp M → R
The last condition means that if (U, (x1 , . . . , xn )) is a smooth coordinate system on
M , then each of the functions gij : U → R defined by
* +
∂ ∂ ∂ ∂
gij (p) = , or gij = ,
∂xi p ∂xj p ∂xi ∂xj
2
1.1 Geometry of a Riemannian Manifold
One can use the Riemannian manifold to define lengths of vectors and lengths of
curves. Thus if v ∈ Tp M , the length of v is given by the formula
q
|v| = hv, vip .
hv, wip
cos θ = .
|v| |w|
= (a1 , . . . , an ) · (b1 , . . . , bn ),
3
An Important Class of Examples. Suppose M n is a smooth manifold and F :
M n → RN is an imbedding. We then define the induced Riemannian metric on M
by
hv, wip = F∗p (v) · F∗p (w), for v, w ∈ Tp M.
We can identify TF (p) RN with RN . Then
!
∂ ∂F
F∗p i
= (p),
∂x p ∂xi
In the case of Euclidean space Rn , with coordinates (x1 , . . . , xn ) and the Riemannian
metric
h , i = dx1 ⊗ dx1 + · · · + dxn ⊗ dxn ,
the curves of smallest length joining p and q are straight lines.
4
This question can be formulated with the calculus of variations.
Let Ω(M ; p, q) = {smooth paths γ : [0, 1] → M | γ(0) = p and γ(1) = q}. Define
L : Ω(M ; p, q) → R by L(γ) =length of γ. The function L : Ω(M ; p, q) → R is difficult
to deal with because
2. If γ
e is a reparametrization of γ, L(e
γ ) = L(γ).
1 1 0
Z
J(γ) = hγ (t), γ 0 (t)i dt
2 0
with equality holding if and only if the functions 1 and hγ 0 (t), γ 0 (t)i are linearly inde-
pendent.
Proof. If λ ∈ Ω,
2J(γ) = [L(γ)]2 ≤ [L(λ)]2 ≤ 2L(γ).
5
Proposition. If γ ∈ Ω minimizes J, it minimizes L and has constant speed.
If γ minimizes J, we must have equality, and by the first proposition, γ has constant
speed. Suppose that λ ∈ Ω and L(λ) < L(γ). If λ has a unit speed reparametrization
λ,
e then
2J(λ) e 2
e = (L(λ))
= (L(λ))2
< (L(γ))2
= 2J(γ),
There is only one problem with this sketch: it assumes we can reparametrize γ and
λ to have constant speed. This can always be done, so long as γ 0 (t) and λ0 (t) never
vanish.
But there is also an easier argument we will be able to give later. The upshot is that
γ ∈ Ω minimizes J iff γ ∈ Ω minimizes L and has constant speed. This motivates
looking at the calculus of variations problem for
J : Ω(M ; p, q) → R.
6
1. ᾱ(0) = γ, and
2. the map α : (−ε, ε) × [0, 1] → M defined by α(s, t) = ᾱ(s)(t) is smooth.
[γ 00 (t)]T = 0,
α : (−ε, ε) × [0, 1] → M ⊆ RN
7
is a smooth map. Dropping F , we can regard α as a smooth map
α : (−ε, ε) × [0, 1] → RN
1 1 ∂ ∂α
Z
d ∂α
(J(ᾱ(s))) = (s, t) · (s, t) dt
ds 2 0 ∂s ∂t ∂t
Z 1 2
∂ α ∂α
= (s, t) · (s, t) dt
0 ∂s∂t ∂t
Z 1
∂α ∂2α
=− (s, t) · 2 (s, t) dt.
0 ∂s ∂t
The last step follows by integration by parts, since
∂α ∂α
(s, 0) ≡ 0 = (s, 1).
∂s ∂s
Thus
Z 1
∂2α
d ∂α
(J(ᾱ(s)))
=− (0, t) 2 (0, t) dt
ds s=0 0 ∂s ∂t
Z 1
=− V (t) · γ 00 (t) dt,
0
8
Chapter 2
Motivated by the theorem we proved last time, we say that a smooth map γ : (a, b) →
M n ⊆ RN is a constant speed geodesic if it satisfies the condition
where (·)T is the orthogonal projection to the tangent space Tγ(t) (M ) ⊆ Tγ(t) RN .
Thus γ : (a, b) → M n ⊆ RN is a constant speed geodesic if the tangential component
of its acceleration is zero.
Thus, for example, if (e1 , e2 , e3 ) is an orthonormal basis for R3 , S 2 is the unit sphere
in R3 , then the great circle γ : R → S 2 ⊆ R3 ,
which is perpendicular to Tγ(t) S 2 . Indeed, it has unit speed (γ 0 (t) · γ 0 (t) = 1) and is
a smooth closed geodesic, since γ(t + 2π) = γ(t) for all t ∈ R.
M 2 = {(x, y, z) ∈ R3 : x2 + y 2 = 1},
9
the right circular cylinder, then the curve
is a constant speed geodesic for all choices of constants a and b. Thus the geodesics
on the right circular cylinder are helices and degenerate cases of these (circles and
lines).
This is a very famous theorem by the mathematician John Nash who received a Nobel
Prize in Economics. (See the movie A Beautiful Mind.) However, given (M n , h , i),
it is usually quite difficult to construct F , and in any case, F is by no means unique!
Thus we would like an intrinsic definition of geodesics, one that does not require
an “isometric” imbedding F : M n → RN .
we can set
n n 2 i
d2 (xi ◦ γ)
X ∂ X dx ∂
acceleration of γ at t = (t) i = (t). (∗)
i=1
dt2 ∂x γ(t) i=1
dt2 ∂xi
10
However, under a change to curvilinear coordinates,
d2 xi X ∂xi d2 y j X ∂ 2 xj dy j dy k
= + ,
dt2 ∂y j dt2 ∂y j ∂y k dt dt
so (∗) does not hold in curvilinear coordinates. We need a more subtle definition of
acceleration, based upon the notion of connection.
Definition. Let X(M n ) be the set of smooth vector fields on M and F (M n ) the set
of smooth real-valued functions on M . A connection (in the tangent bundle T M to
M ) is a map
∇ : X(M n ) × X(M n ) → X(M n )
(we write ∇X Y for ∇(X, Y )) such that
1. ∇f X+gY Z = f ∇X Z + g∇Y Z,
2. ∇Z (f X + gY ) = (Zf )X + f ∇Z X + (Zg)Y + g∇Z Y
for X, Y, Z ∈ X(M n ), f, g ∈ F (M n ).
11
The lemma can be restated as
If (U, (x1 , . . . , xn )) is such a local coordinate chart, we can define smooth functions
Γkij : U → R by
n
∂ X ∂
∇ ∂i j = Γkij k .
∂x ∂x ∂x
k=1
n
X
j ∂
∇X Y = ∇f i ∂ g
i,j=1
∂xi ∂xj
n j
X
i j ∂ i ∂g ∂
= f g ∇ ∂i j + f (∗)
i,j=1
∂x ∂x ∂xi ∂xj
n
" n #
X X ∂g i X ∂
= fj j + Γijk f j g k .
i=1 j=1
∂x ∂xi
Lemma. (∇X Y )(p) depends only on X(p) and on Y on some curve tangent to X(p).
12
In terms of local coordinates,
n
d2 (xi ◦ γ)
0
X ∂
(∇γ 0 γ )(t) = (t) i
i=1
dt2 ∂x γ(t)
n i
◦ γ) d(xi ◦ γ)
X d(x ∂
+ Γkij (γ(t)) (t) (t) k .
i,j,k=1
dt dt ∂x γ(t)
A. ∇ is symmetric: ∇X Y − ∇Y X = [X, Y ].
We will prove this next time and show that for this connection, if γ : (a, b) → M n ,
13
Chapter 3
14
Fundamental Theorem. If (M, h , i) is a Riemannian manifold, there is a unique
connection ∇ on T M which satisfies the conditions
A. ∇X Y − ∇Y X = [X, Y ],
for X, Y, Z ∈ X(M n ).
A0 . Γkij = Γkji .
∂gij Pk Pn
B0 . ∂xk
= `=1 gi` Γ`jk + `=1 gj` Γ`ik .
and so
n
∂gjk ∂gik ∂gij X
+ − = 2 g`k Γ`ij .
∂xi ∂xj ∂xk `=1
15
If (g mk ) = (gmk )−1 , we conclude that
n
1 X mk ∂gjk ∂gik ∂gij
Γm
ij = g + − k . (∗)
2 `=1 ∂xi ∂xj ∂x
For existence, use (∗) to define Γkij and hence ∇ on a coordinate chart (U, (x1 , . . . , xn )).
Check that ∇ satisfies A and B.
Local uniqueness implies that the locally defined ∇’s fit together to give a globally
defined connection ∇ on T M which satisfies A and B.
Pn Pn
If X = i=1 f i ∂x∂ i and Y = i=1 g i ∂x∂ i , then
n n
X ∂g j ∂ X ∂
∇X Y = fi
i i
+ f i g j Γkij k
i,j=1
∂x ∂x i,j,k=1
∂x
n
" n
#
X
k
X
i j k ∂
= X(g ) + f g Γij .
k=1 i,j=1
∂xk
This formula shows that (∇X Y )(p) depends only on X(p) and on Y along some
curve tangent to X(p). Thus if γ : (a, b) → M is a smooth curve, we can define
∇γ 0 γ 0 (t) ∈ Tγ(t) M .
16
We need to check that this agrees with the definition previously given when M n ⊆ RN
has the induced metric. To do this, we first note that the Levi-Civita connection for
Euclidean space Rn with Euclidean coordinates (x1 , . . . , xn ) and Riemannian metric
n
(
X
i j 1 if i = j,
gij dx ⊗ dx , gij = δij =
i,j=1
0 if i 6= j.
Pn
is given by Γkij = 0. (This follows from (∗).) Thus if X and Y = i=1 g i ∂x∂ i are vector
fields on Rn , the Levi-Civita connection for Euclidean space is
n
X ∂
DX Y = X(g i ) .
i=1
∂xi
where (·)T is the orthogonal projection into the tangent space. One verifies that
(∗∗) defines a connection on T M which satisfies A0 and B0 . Thus (∗∗) defines the
Levi-Civita connection on T M . In particular,
17
Chapter 4
Last time we saw that there is a unique connection ∇, the so-called Levi-Civita
connection, on T M such that
A. ∇X Y − ∇Y X = [X, Y ],
where n
1 X k` ∂gi` ∂gj` ∂gij
Γkij = g + − .
2 `=1 ∂xj ∂xi ∂x`
In the special case where M n ⊆ RN and h , i is the metric induced by the Euclidean
dot product on RN ,
∇X Y = (DX Y )T ,
18
where (·)T is the orthogonal projection to T M and D is the usual directional derivative
on RN . In terms of Euclidean coordinates (u1 , . . . , uN ) on RN ,
N
! N
I ∂ ∂
X X
DX f I
= (Xf I ) I .
I=1
∂u I=1
∂u
We write xi (t) for (x◦ γ)(t). Then γ : (a, b) → M is a geodesic if and only if
n
d2 xi X
i dxj dxk
(t) + Γjk (γ(t)) (t) (t) = 0. (∗)
dt2 j,k=1
dt dt
dxi
If we let ẋi (t) = dt
(t), we can rewrite (∗) as
( i
dx
dt
= ẋi ,
i (∗∗)
dẋ
= − nj,k=1 Γijk ẋj ẋk .
P
dt
19
Suppose p ∈ M , v ∈ Tp M . It follows from the fundamental existence and uniqueness
theorem for ordinary differential equations that for some ε > 0, there is a unique
solution xi (t), −ε < t < ε to (∗) such that
dxi
xi (0) = p, (0) = ẋi (0) = dxi |p (v).
dt
Thus there is a unique geodesic γv : (−ε, ε) → M such that
for any p ∈ M , v ∈ Tp M .
It is usually quite difficult to find explicit solutions to (∗). The explicit solutions
usually occur only when M has lots of isometries.
Examples.
20
just the constant speed straight lines.
Define F : R2 → R2 by
1 1 1
x ◦F cos θ − sin θ x c
= + 2 ,
x2 ◦ F sin θ cos θ x2 c
where θ, c1 , and c2 are constants. Then F is an isometry. One can construct an
isometry which takes any straight line into any other straight line.
Suppose next that H2 = {(x, y) ∈ R2 : y > 0}, the so-called Poincaré upper
half-plane, with the Riemannian metric
1
h , i = 2 (dx2 + dy 2 ).
y
We can calculate the geodesics by brute force. Let x1 = x, x2 = y. Then a straight-
forward calculation gives
Γ111 = Γ122 = Γ121 = Γ112 = 0
1
Γ222 = Γ121 = Γ112 = −
y
2
Γ11 = y.
21
Hence the equations for geodesics are
( 2
d x dy
dt2
− y2 dx
dt dt
=0
d2 y 1 dy 2 dx 2
− y dt + y1
dt2 dt
= 0.
It dx
dt
= 0 at one point, it follows from the first equation that dx
dt
≡ 0 and from the
second that 2
d2 y 1 dy
− = 0.
dt2 y dt
dy
Reduce the order, setting z = dt
, so
(
dy
dt
=z dz z
dz
⇒ = .
dt
= y1 z 2 dy y
dz dy dy
Hence z
= y
, log |z| = log |y| + C, z = (±ec )y = by, dt
= by, y = aebt . Thus
x = 0, y = aebt
dx
If dt
6= 0 and y is a function of x,
d2 y
d dy dx
=
dt2 dt dx dt
d dy dx dy d2 x
= +
dt dx dt dx dt2
2
d2 y dx dy d2 x
= 2 + ,
dx dt dx dt2
and hence
2 2 2
d2 y d2 x
1 dy 1 dx dx dy
− = 2 +
y dt y dt dx dt dx dt2
2
d2 y
dx 2 dy dx dy
= 2 + .
dx dt y dx dt dt
22
dx 2
Dividing by dt
allows us to eliminate t:
2 2
d2 y 2 dy
1 dy 1
− = 2+ ,
y dx y dx y dx
2
d2 y dy
y 2+ = −1,
dx dx
d dy
y = −1,
dx dx
dy
y = x + 2a,
dx
y 2 = −x2 + ax + b.
y 2 + x2 + ax + b = 0.
Of course, we still need to find the constant speed parametrizations of the semicircles.
These could be determined by exploiting the three-dimensional group of isometries
which operates on H2 . Let
a b
SL(2, R) = : ad − bc = 1 ,
c d
a b
the so-called special linear group. If A = ∈ SL(2, R), define FA : H2 → H2
c d
by
az + b
x ◦ F1 + iy ◦ Fa = , z = x + iy.
cz + d
It is an exercise to show that each FA is an isometry. Moreover, given any semicircle,
there is an A ∈ SL(2, R) such that FA takes a vertical line to that semicircle.
23
The three Riemannian manifolds
have three-dimensional groups of isometries. Along with RP 2 , they are the only Rie-
mannian manifolds with three-dimensional groups of isometries.
24
Chapter 5
h , i = dt ⊗ dt + ds ⊗ ds,
But there are many other imbeddings from R2 into R3 which induce exactly the same
metric. Consider, for example, the catenary z = cosh y. We can parametrize the
catenary by arc length; one can check that
√ √
y = s1 + 1, z = log(s + s2 + 1)
We say that the parametrized surfaces F0 and F1 have the same intrinsic geometry
because the Riemannian metrics induced on R2 are the same, but different extrinsic
25
geometry because the shapes of the images F0 (R2 ) and F1 (R2 ) are different.
How can we recognize when the intrinsic geometry of a surface is not that of Euclidean
space?
The invariant that answers this question is the Riemann-Christoffel curvature tensor.
is a Levi-Civita connection ∇,
n n
∂ X 1 X k` ∂gi` ∂gj` ∂gij
∇ ∂i j = Γkij , Γkij = g + − .
∂x ∂x
k=1
2 `=1 ∂xj ∂xi ∂x`
26
Just like ∇, R is a local operator and it is clearly linear over the field R of real num-
bers. However, it is also linear over the ring F (M ) = {smooth real-valued functions
f : M → R}.
Proposition.
R(X, Y )f Z = ∇X ∇Y (f Z) − ∇Y ∇X (f Z) − ∇[X,Y ] (f Z)
= ∇X ((Y f )Z + f ∇Y Z) − ∇Y ((Xf )Z + f ∇X Z) − ([X, Y ](f )Z + f ∇[X,Y ] Z)
= XY (f ) + (Y f )∇X Z + (Xf )∇Y Z + f ∇X ∇Y Z
− Y X(f ) − (Xf )∇Y Z − (Y f )∇X Z − f ∇Y ∇X Z
− [X, Y ](f )Z − f ∇[X,Y ] Z
= f (∇X ∇Y Z − ∇Y ∇X Z − ∇[X,Y ] Z)
= f R(X, Y )Z,
as desired.
`
The Proposition justifies defining the components Rkij of R by
n
∂ ∂ ∂ X
` ∂
R , = R kij .
∂xj ∂xj ∂xk `=1
∂x `
Proposition.
n
` ∂ ` ∂ `
X
Rkij = (Γ ) − j (Γki ) + (Γm ` m `
jk Γim − Γik Γjm ).
∂xi kj ∂x m=1
27
Proof.
n
X
` ∂ ∂ ∂ ∂
Rkij =R , j
`=1
∂x` ∂x ∂x ∂xk
i
∂ ∂
= ∇ ∂ i ∇ ∂j k − ∇ ∂j ∇ ∂ i k
∂x ∂x ∂x ∂x ∂x ∂x
m
! m
!
X ∂ X ∂
= ∇ ∂i Γ`jk ` − ∇ ∂ j Γ`ik `
∂x
`=1
∂x ∂x
`=1
∂x
n
X ∂ ` ∂ `
X
` m ` m ∂
= i
(Γjk ) − j (Γik ) + (Γim Γjk − Γjm Γik ) ,
`=1
∂x ∂x ∂x`
as desired.
If M is Euclidean space,
(
1 if i = j,
gij = δij = ⇒ Γ`ij = 0 ⇒ Rkij
`
= 0.
0 if i 6= j
∂ ∂
Note. It suffices to prove these symmetries for coordinate fields X = ∂xi
, Y = ∂xj
,
Z = ∂x∂ k , and W = ∂x∂ ` .
28
Proof of 1.
∂ ∂ ∂ ∂ ∂
R i
, j k
= ∇ ∂ i ∇ ∂j k − ∇ ∂j ∇ ∂ i k
∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x
∂ ∂ ∂
= −R , i .
∂x ∂x ∂xk
j
Proof of 2.
∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂ ∂
R i
, j k
+R j
, k i
+R , i
∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂xj
k
∂ ∂
= ∇ ∂ i ∇ ∂j k − ∇ ∂j ∇ ∂ i k
∂x ∂x ∂x ∂x ∂x ∂x
∂ ∂
+ ∇ ∂j ∇ ∂ i
− ∇ ∂ ∇ ∂j i
∂x ∂x ∂x
k ∂x k ∂x ∂x
∂ ∂
+ ∇ ∂ ∇ ∂i j − ∇ ∂i ∇ ∂ = 0,
∂x k ∂x ∂x ∂x ∂x ∂xj
k
∂ ∂
where we have used the fact that ∇ ∂
∂xj
=∇ ∂
∂xi
.
∂xi ∂xj
But
∂2 ∂2
∂ ∂
0= − j i ,
∂xi ∂xj ∂x ∂x ∂xk ∂xk
∂ ∂ ∂ ∂ ∂ ∂
= 2 i ∇ ∂j k , k − 2 j ∇ ∂ i k , k
∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x
∂ ∂ ∂ ∂
= 2 ∇ ∂ i ∇ ∂j k , k + 2 ∇ ∂j k , ∇ ∂ i k
∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x
∂ ∂ ∂ ∂
− 2 ∇ ∂j ∇ ∂ i k , k − 2 ∇ ∂ i k , ∇ ∂j k
∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x
∂ ∂ ∂ ∂
=2 R , , .
∂xi ∂xj ∂xk ∂xk
29
Proof of 4.
hR(X, Y )W, Z) = −hR(Y, X)W, Zi
= hR(X, W )Y, Zi + hR(W, Y )X, Zi
= −hR(X, Y )Z, W i
= hR(Y, Z)X, W i + hR(Z, X)Y, W i,
so
2hR(X, Y )W, Zi = hR(X, W )Y, Zi+hR(W, Y )X, Zi+hR(Y, Z)X, W i+hR(Z, X)Y, W i.
Interchanging (X, Y ) and (W, Z) gives
2hR(W, Z)X, Y i = hR(W, X)Z, Y i+hR(X, Z)W, Y i+hR(Z, Y )W, Xi+hR(Y, W )Z, Xi.
The terms on the right sides are equal in pairs, so
2hR(X, Y )W, Zi = 2hR(W, Z)X, Y i.
The first of the above propositions shows that R restricts to a multilinear map
R : Tp M × Tp M × Tp M → Tp M
for each p ∈ M . We can define
R : Tp M × Tp M × Tp M × Tp M → R
by R(x, y, z, w) = hR(x, y)w, zi. Then the last of the above propositions implies that
R satisfies the symmetries
R(x, y, z, w) = −R(y, x, z, w) = −R(x, y, w, z) = R(z, w, x, y)
and
R(x, y, z, w) + R(y, z, x, w) + R(z, x, y, w) = 0.
30
Chapter 6
h , ip : Tp M × Tp M → R.
A. ∇X Y − ∇Y X = [X, Y ],
∂
Pn k ∂ k 1
Pn k` ∂gi` ∂gj` ∂gij
In local coordinate, ∇ ∂
∂xj
= k=1 Γij ∂xk , where Γij = 2 `=1 g ∂xj
+ ∂xi
− ∂x`
.
∂xi
R(X, Y )Z = ∇X ∇Y Z = ∇Y ∇X Z − ∇[X,Y ] Z.
31
Examples. First, suppose M N = RN with coordinates (x1 , . . . , xN ) and metric
N
! N
X ∂ X ∂
DX fI I = X(f I ) .
I=1
∂x I=1
∂xI
Next, suppose that M n ⊆ RN , a submanifold with the induced metric. Vector fields
X, Y on M n can be extended to vector fields X, Y on RN . Note that if p ∈ M n ,
(DX Y )(p) depends only on X(p) and on Y along some curve tangent to X(p). Thus
(DX Y )(p) depends only on the values of X and Y along M n itself.
and check that it satisfies the axioms for a connection on T M . Moreover, ∇ satisfies
axioms A and B. It suffices to check A when X = ∂x∂ i and Y = ∂x∂ j , where (x1 , . . . , xn )
are curvilinear coordinates on M which extend to curvilinear coordinates (x1 , . . . , xN )
on RN , and then
T
∂ ∂ ∂ ∂
∇ ∂ i j − ∇ ∂j i = D ∂ i j − D ∂j i = 0.
∂x ∂x ∂x ∂x ∂x ∂x ∂x ∂x
32
Similarly, axiom B holds: if X, Y, Z are tangent to M n ,
XhY, Zi = X(Y · Z)
= (DX Y ) · Z + Y · (DX Z)
= h∇X Y, Zi + hY, ∇X Zi.
Proof of (a).
α(f X, Y ) = (Df X Y )⊥
= f (DX Y )⊥
= f α(X, Y ).
α(X, f Y ) = (DX (f Y ))⊥
= ((Xf )Y + f DX Y )⊥
= f (DX Y )⊥
= f α(X, Y ).
33
∂ ∂
Proof of (b). We can take X = ∂xi
and Y = ∂xj
. Then
⊥
∂ ∂
α(X, Y ) − α(Y, X) = D ∂ i j − D ∂j i = 0,
∂x ∂x ∂x ∂x
as desired.
We set
κg (s) = (γ 00 (s))T
= (∇γ 0 γ 0 )(s)
= geodesic curvature of γ at s,
κn (s) = (γ 00 (s))⊥
= α(γ 0 (s), γ 0 (s))
= normal curvature of γ at s.
α : Tp M × Tp M → Np M
with the following interpretation: If v ∈ Tp M has unit length, α(v, v) is the normal
curvature of a unit-speed curve γ : (−ε, ε) → M with γ(0) = p and γ 0 (0) = v.
for x, y, z, w ∈ Tp M .
34
This is known as the Gauß equation. It provides an often quite effective means of
calculating the curvature of M n .
0 = (DX DY W − DY DX W − D[X,Y ] W ) · Z
= X(DY X · Z) − (DY W ) · (DX Z)
− Y (DX W · Z) + (DX W · DY Z) − h∇[X,Y ] W, Zi
= Xh∇Y W, Zi − h∇Y W, ∇X Zi − α(Y, W ) · α(X, Z)
− Y h∇X W, Zi + h∇X W, ∇Y Zi + α(X, W ) · α(Y, Z) − h∇[X,Y ] W, Zi
= h∇X ∇Y W, Zi − α(Y, W ) · α(X, Z)
h∇Y ∇X W, Zi + α(X, W ) · α(Y, Z) − h∇[X,Y ] W, Zi
= hR(X, Y )W, Zi − α(Y, W ) · α(X, Z) + α(X, W ) · α(Y, Z),
as desired.
00 1 s 1 s 1
γ (s) = − cos e1 − sin = − N (γ(s)),
a a a a a
where N (p) is the outward-pointing unit normal to S n (a) at p. Then
1
α(γ 0 (s), γ 0 (s)) = − N (γ(s)),
a
and
1
α(v, v) = − N (p), when v ∈ Tp M, hv, vi = 1.
a
35
By polarization,
1
α(x, y) = − hx, yiN (p) for x, y ∈ Tp M.
a
It therefore follows from the Gauß equation that
1
hR(x, y)w, zi = [hx, zihy, wi − hx, wihy, zi]
a2
for x, y, z, w ∈ Tp M .
36
Chapter 7
There is one case in which the curvature of a Riemannian manifold is easy to calcu-
late:
for x, y, z, w ∈ Tp M , the dot on the right denoting the usual dot product in RN .
Hence if Z ∈ X(M ),
0 = (DX DY W − DY DX W − D[X,Y ] W ) · Z
= X(DY W · Z) − (DY W ) · (DX Z) − Y (DX W · Z) + (DX W ) · (DY Z) − (D[X,Y ] W ) · Z
= Xh∇Y W, Zi − h∇Y W, ∇X Zi − α(Y, W ) · α(X, Z)
− Y h∇X W, Zi + h∇X W, ∇Y Zi + α(X, W ) · α(Y, Z)
− hD[X,Y ] W, Zi
= h∇X ∇Y W, Zi − α(Y, W ) · α(X, Z)−i∇Y ∇X W, Zi
+ α(X, W ) · α(Y, Z) − h∇[X,Y ] W, Zi
= hR(X, Y )W, Zi − α(X, Z) · α(Y, W ) + α(X, W ) · α(Y, Z),
37
as desired.
where
−1 0 ··· 0
0 1 · · · 0
J = .. .. .
..
. . .
0 0 · 1
Let O(1, n) = {A ∈ GL(n + 1, R) : AT JA = J}, a subgroup of the general linear
group called the Lorentz group. If
t t
e
x1 x 1
e
= A .. ,
..
. .
n
x en
x
h , i = −de
t ⊗ de x1 ⊗ de
t + de x1 + · · · + de
xn ⊗ de
xn .
38
Elements of O(1, n) are called Lorentz transformations.
We will sometimes identify Rn+1 with its tangent space and denote the metric h , i
on Minkowski space-time by a dot: X · Y = hX, Y i.
The submanifold H n (a) inherits an induced metric from Rn+1 , and we claim that it
is positive definite. Indeed, in terms of the coordinates (x1 , . . . , xn ), we can write
p
t = a2 + (x1 )2 + · · · + (xn )2 ,
and hence Pn i i
i=1 x dx
dt = p ,
a2 + (x1 )2 + · · · + (xn )2
and hence the induced metric is
Pn i j i j
i,j=1 x x dx ⊗ dx
h, i=− 2 + dx1 ⊗ dx1 + · · · + dxn ⊗ dxn .
a + (x1 )2 + · · · + (xn )2
Thus
xi xj
gij = δij − ,
a2 + (x1 )2 + · · · + (xn )2
and one readily checks that the matrix (gij ) is positive definite.
39
Lemma. If p ∈ H n (a), there is an element A ∈ O(1, n) such that
a
0
p = A .. .
.
0
Proof. Suppose
t a cosh α
x1 a sinh α u1
p = .. = ,
..
. .
xn a sinh α un
where
u1
..
.
un
is a unit-length vector in Rn . Let B be a rotation matrix of Rn such that
1 1
u 0
..
. = B .. .
.
un
0
Then
cosh α sinh α 0 · · · 0
1 0 ··· 0 a cosh α
sinh α cosh α 0 · · · 0 a
0
.. a sinh α u1
..
0
0 1 ··· 0 0 . = .. ,
. B .. .. .. .. .
. . . . 0
0 a sinh α un
0 0 0 ··· 1
so we can take
cosh α sinh α 0 · · · 0
1 0 ··· 0
0 sinh α cosh α 0 · · · 0
A = ..
0
0 1 ··· 0 ,
. B .. .. .. ..
. . . .
0
0 0 0 ··· 1
40
as desired.
Proof. Exercise.
γ
e(s) = Aγ(s),
where A ∈ O(1, n) and A(H n (a)) ⊆ H n (a). These are just parametrizations of inter-
sections of H n (a) with planes passing through the origin.
If γ is defined by (∗),
s s
γ 0 (s) = sinh , cosh , 0, . . . , 0
a a
41
and
1
α(γ 0 (s), γ 0 (s)) = (γ 00 (s))⊥ = γ(s).
a2
Thus if x ∈ Tp H n (a) and hx, xi = 1,
1
α(x, x) = N (p),
a
where N is the future-pointing normal vector at p with N (p) · N (p) = −1. By
polarization,
1
α(x, y) = hx, yiN (p) for x, y ∈ Tp H n (a).
a
Since N (p) · N (p) = −1, it follows from the theorem presented at the beginning of
the lecture that
1
hR(x, y)w, zi = − [hx, zihw, yi − hx, wihy, zi].
a2
E n has K(σ) ≡ 0 for all 2-planes. S n (a) has K(σ) ≡ a12 for all 2-planes. H n (a) has
K(σ) ≡ − a12 for all 2-planes. These are the spaces of constant curvature, the
Euclidean and non-Euclidean geometries in n dimensions.
Proposition. Let
R, S : Tp M × Tp M × Tp M × Tp M → R
then R = S.
42
Thus the sectional curvatures determine the full curvature tensor.
Hence
0 = T (x, y + z, x, y + z)
= T (x, y, x, y) + T (x, y, x, z) + T (x, z, x, y) + T (x, z, x, z)
= 2T (x, y, x, z),
and
Thus
as desired.
43
Chapter 8
February 2, 2006
Note that h∇X Z, Y i(p) depends only on X(p) and Y (p). Thus to show that Z is
Killing it suffices to show that
h∇X Z, Y i(p) + hX, ∇Y Zi(p) = 0
for all vector fields X and Y such that hX, Xi, hY, Y i, and hX, Y i are constant.
Remark. The converse is also true, but we will not prove it.
44
Proof. We can assume that hX, Y i is constant. Then hϕt∗ (X), ϕt∗ i is constant, and
so
d
0 = h(ϕt∗ X)(p), (ϕt∗ Y )(p)i
dt
t=0
d d
= (ϕt∗ X(p)) Y (p) + X(p), (ϕt∗ Y )(p)
dt t=0 dt t=0
= −h[Z, X], Y i(p) − hX, [Z, Y ]i(p),
h∇X Z, Y i + hX, ∇Y Zi = 0.
Proof.
(a)
d 0 0
hγ , γ i = 2h∇γ 0 γ 0 , γ 0 i = 0;
dt
45
(b)
d
hZ, γ 0 i = h∇γ 0 Z, γ 0 i + hZ, ∇γ 0 γ 0 i = 0.
dt
46
dθ
Eliminating dt
yields
2
L2
1 du
+ = E. (∗)
2 dt 2(r(u))2
This is first order equation for u in terms of t, and the solution can be reduced to an
integration.
L2
VL (u) = .
2(2 + cos u)2
L2 L2
If 18
<E< 2
, the motion will be confined to one of the potential wells.
One can also calculate the curvature of surfaces of revolution quite easily.
47
where (r0 (u))2 + (z 0 (u))2 = 1, a direct calculation gives
∂F
= (r0 (u) cos θ, r0 (u) sin θ, z 0 (u)),
∂u
∂F
= (−r(u) sin θ, r(u) cos θ, 0),
∂θ
N (u, θ) = unit normal
∂F ∂F
∂u
× ∂θ
= ∂F ∂F
∂u
× ∂θ
= (−z 0 (u) cos θ, −z 0 (u) sin θ, r0 (u)).
Thus
∂ ∂ ∂
α , ·N =D∂ ·N
∂u ∂u ∂u ∂u
∂2F
·N
=
∂u2
= −r00 (u)z 0 (u) + r0 (u)z 00 (u),
∂2F
∂ ∂
α , ·N = ·N
∂u ∂θ ∂u∂θ
= 0,
∂2F
∂ ∂
α , ·N = ·N
∂θ ∂θ ∂θ2
= r(u)z 0 (u).
48
Chapter 9
February 7, 2006
Definition. A Lie group is a group G which is also a manifold such that the maps
and
ν:G→G defined by ν(σ) = σ −1
are smooth.
If G is a Lie group, a vector field X ∈ X(G) is said to be left invariant if (Lσ )∗ (X) =
X for all σ ∈ G.
49
Let g = {X ∈ X(G) : (Lσ )∗ (X) = X for all σ ∈ G}. g is called the Lie algebra of
G. If X, Y ∈ g, then
so [X, Y ] ∈ g.
α : g → Te G, β : Te G → g
by α(X) = X(e), β(x)(σ) = (Lσ )∗e (x). One easily checks that α ◦ β = id, β ◦ α = id,
so α and β are real vector space isomorphisms.
F∗ (X)(F (σ)) = (LF (σ) )∗ (F∗ (X)(e)) = F∗σ (Lσ∗ (X(e))) = F∗ (X(σ)),
More generally, a Lie algebra is a real vector space g together with an operation
[ , ] : g × g → g such that
50
A Lie algebra homomorphism between Lie algebras g and h is a linear map
f : g → h such that
f ([X, Y ]) = [f (X), f (Y )].
We have defined a “functor” which assigns to each Lie group G its Lie algebra g and
to each Lie group homomorphism F : G → H the corresponding Lie algebra homo-
morphism F∗ : g → h. This is called the Lie group–Lie algebra correspondence.
{ϕt = Rθ(t) : t ∈ R}
Proof. Let θX : (a, b) → G be the maximal integral curve for X such that θX (0) = e.
Assume b < ∞. If σ ∈ G,
51
The one-parameter group of diffeomorphisms corresponding to X is now seen to be
{ϕt = RθX (t) : t ∈ R}, and
so θX : R → G is a one-parameter subgroup of G.
O(n) = {A ∈ GL(n, R) : AT A = I}
2
is a submanifold of GL(n, R) ⊆ Rn . It inherits a Riemannian metric h , i which we
claim is biinvariant. Indeed, if we define
a11 · · · a1n
xij : O(n) → R by xij ... .. = ai ,
. j
a1 · · · ann
n
then n
X
h, i= dxij ⊗ dxij .
i,j=1
Moreover,
n
X X
xij (AB) = aik xkj (B) ⇒ xij ◦ La = aik xkj
k=1
X X
⇒ L∗A (xij ) = aik xkj ⇒ L∗A (dxij ) = aik dxkj .
52
Hence
n
X
L∗A h , i = L∗A (dxij ) ⊗ L∗A (dxij )
i,j=1
Xn
= aik ai` dxkj ⊗ dx`j
i,j,k,`=1
X n
= dxkj ⊗ dxkj
j,k=1
= h , i,
(a) geodesics passing through the origin are just the one-parameter subgroup of G.
But since hX, Xi is constant, 0 = Y hX, Xi = h2∇Y X, Xi. Hence it follows from (∗)
that h∇X X, Y i = 0, which implies ∇X X = 0. But then
0 = ∇X+Y (X + Y ) = ∇X Y + ∇Y X,
53
and
1
∇X Y − ∇Y X = [X, Y ] ⇒ ∇X Y = [X, Y ].
2
0
This gives (b), and ∇X X = 0 ⇒ ∇θX0 (t) θX (t) = 0, so θX : R → G is a geodesic,
yielding (a). Finally
R(X, Y )W = ∇X ∇Y W − ∇Y ∇X W − ∇[X,Y ] W
1 1 1
= [X, [Y, W ]] − [Y, [X, W ]] − [[X, Y ], W ]
4 4 2
1 1 1
= [X, [Y, W ]] + [Y, [W, X]] − [[X, Y ], W ]
4 4 2
1 1
= − [W, [X, Y ]] − [[X, Y ], W ]
4 2
1
= − [[X, Y ], W ].
4
If X, Y, Z ∈ g,
and hence
1
hR(X, Y )W, Zi = − h[[X, Y ], W ], Zi
4
1
= h[W, [X, Y ]], Zi
4
1
= − h[X, Y ]i, [W, Z]i
4
1
= h[X, Y ], [Z, W ]i,
4
finishing (c).
54
Chapter 10
February 9, 2006
Thus if b ∈ Rn ,
dx
x(t) = etA · b is a solution to = Ax.
dt
The matrix exponential θA (t) = etA thus yields the solutions to systems of homoge-
neous linear first order ordinary differential equations with constant coefficients.
55
Proposition. etA esA = e(t+s)A .
Proof. X1 (t) = etA · esA and X2 (t) = e(t+s)A are both solutions to the initial value
problem
d(X(t))
= A · X(t), X(0) = esA ,
dt
so they must be equal by uniqueness of solutions to systems of ordinary differential
equations.
a11 · · · a1n b11 · · · b1n
If A = ... .. and B = .. .. , then
. . .
an1 · · · an n
b1 · · · bnn
n
n
X
xij (B · θA (t)) = xik (B)xkj (θA (t)),
k=1
n
d i X
(xj (B · θA (t))) =
xik (B)xkj (A)
dt t=0 k=1
n
X
= xik (B)akj .
k=1
56
is n
X ∂
X |B = xik (B)akj .
i,j,k=1
∂xij B
If n
X ∂
XB = xik bkj ,
i,j,k=1
∂xij
then a direct calculation yields
XX
i k ∂ λ ν ∂ λ ν ∂ i k ∂
XX
[XA , XB ] = x k aj i x ν b µ µ − x ν b µ µ x k aj i
i,j,k λ,µ,ν
∂xj ∂xλ λ,µ,ν i,j,k
∂xλ ∂xj
X ∂ ∂
= xik akj bj` ` − xik bkj aj` `
i,j,k,`
∂xi ∂xi
X ∂
= xik (xkj (AB) − xkj (BA)) i
i,j,k
∂xj
= X[A,B] ,
where [A, B] = AB −BA. Hence we see that the Lie algebra of GL(n, R) is gl(n, R) =
{n × n real matrices} with bracket [A, B] = AB − BA. Most important Lie groups
are subgroups of GL(n, R), and their Lie algebras are subalgebras of gl(n, R).
T
Thus θA (t) = etA takes values in O(n) iff etA · etA = I. Differentiating and setting
t = 0 yields the relation
d tA tAT
(e · e ) = A + AT = 0.
dt t=0
57
Proposition. Suppose (M, h , i) is a pseudo-Riemannian manifold, p ∈ M n . Then
there is an open neighborhood V of 0 in Tp M such that if v ∈ Tp M , the unique
geodesic γv which satisfies the initial conditions
Define expp : V → M by expp (v) = γv (1). expp is called the exponential map.
Note that if v ∈ V , t 7→ expp (tv) is a geodesic (because expp (tv) = γtV (1) = γv (t)),
and hence expp takes line segments through the origin in Tp M to geodesic segments
in M . For example, suppose M is a Lie subgroup G of GL(n, R) with a biinvariant
metric. (For example, G = O(n).) Let p = I, the identity matrix and v = A ∈ g ⊆
gl(n, R). Then the geodesic γA satisfying the initial conditions
t2 2
γA (t) = θA (t) = etA = I + tA + A + ··· .
2!
Thus expI (A) = eA in this case.
58
Proposition. There is an open neighborhood U e of 0 ∈ Tp M which expp maps dif-
feomorphically onto an open neighborhood U of p in M .
59
Chapter 11
Idea. Find coordinates near a point p which make the gij ’s as simple as possible.
expp : Tp M → M
60
(ẋ1 , . . . , ẋn ) is a global coordinate system on Tp M .
p
Let r : U → R be defined by r = (x1 )2 + · · · + (xn )2 , and define a radial vector
field R on U − {p} by
n
X xi ∂
R= .
i=1
r ∂xi
To prove the Gauß Lemma, we make use of the rotation vector field
∂ ∂
Eij = xi j
− xj i .
∂x ∂x
Lemma 1 [Eij , R] = 0.
Proof. The radial vector field R is invariant under rotation. Thus if {ϕt : t ∈ R} is
the one-parameter group corresponding to Eij ,
d
[Eij , R] = − (ϕt∗ (R)) = 0.
dt t=0
61
Lemma 2 If ∇ is the Levi-Civita connection on M , ∇R R = 0.
Pn
Proof. If (a1 , . . . , an ) are constants with i=1 (a
i 2
) = 1, then the curve γ defined by
xi (γ(t)) = ai t
is an integral curve for R as one verifies by direct calculation. On the other hand,
!
X
i ∂
γ(t) = expp a t i ,
∂x p
Then hR, Eij i is constant along the geodesic rays emanating from p. Let kXk =
p
hX, Xi. Then
|hR, Eij i| ≤ kRk kEij k ≤ kEij k → 0
62
as r → 0. It follows that hR, Eij i ≡ 0.
Proof of Gauß Lemma. First note that R(r) = 1, Eij (r) = 0, by direct calculation.
Suppose that X is a smooth vector field on part of U − {p} of the form
n
∂ X
X=f + fij Eij ,
∂r i,j=1
Let q = expp (v). If λ : [0, 1] → M n is any smooth curve such that λ(0) = p and
λ(1) = q, then
Proof. We use normal coordinates (x1 , . . . , xn ) defined on U . Note that L(γ) =length
of γ = r(q). Suppose that λ : [0, 1] → M is a smooth curve with λ(0) = p and
λ(1) = q. We need to show that L(λ) ≥ L(γ).
63
Case I. λ does not leave U . Then
Z 1 p
L(λ) = hλ0 (t), λ0 (t)i dt
0
Z 1
= kλ0 (t)k dt
Z0 1
≥ hλ0 (t), R(λ(t))i dt
Z0 1
= dr(λ0 (t)) dt
0
= r(λ(1)) − r(λ(0))
= r(q)
= L(γ).
This finishes the proof for L. The proof for J is quite similar.
64
The only difficult part in showing that d is a metric is the implication
d(p, q) = 0 ⇒ p = q.
But if q lies in a normal coordinate system centered at p, this implication follows from
the preceding theorem, while if q does not lie in such a coordinate system, the proof
of the theorem shows that d(p, q) > ε for some ε > 0.
65
Chapter 12
We can give a small extension of the results established in Lecture 10. Namely, on
an open neighborhood U of the zero-section in T M , we can define
exp : U → M × M by exp(v) = (p, expp (v)),
for v ∈ Tp M .
It follows from this Proposition and the results of Lecture 11 that any point p has a
neighborhood U with the following property: if q1 , q2 ∈ U , q1 and q2 can be joined by
a unique length-minimizing geodesic (which depends smoothly on (q1 , q2 ) ∈ M × M ).
Recall that a Riemannian manifold (M, h , i) can be made into a metric space (M, d)
by setting
d(p, q) = inf{length of γ | γ : [a, b] → M is a smooth curve with γ(a) = p, γ(b) = q}.
The proof of the theorem given last time shows that if γ : [a, b] → M is a smooth
curve with γ(a) = p and γ(b) = q and length of γ is d(p, q), then γ is a reparametriza-
tion of a constant speed geodesic.
66
Definition. A curve γ : [a, b] → M is a minimal geodesic if it is a constant speed
geodesic and its length is d(p, q).
Not Always. If M = R2 − {(0, 0)} with the standard Euclidean metric, p = (−1, 0)
and q = (1, 0), there is no minimal geodesic from p to q.
Proof of Theorem 1 We will in fact prove the following assertion (also needed in
the proof of Theorem 2): If p is a point in a Riemannian manifold (M n , h , i) such
that expp : Tp M → M is globally defined, then p can be joined to any point q ∈ M
by a minimal geodesic.
67
A candidate for the minimal geodesic: Let B eε be a closed ball of radius ε centered
at 0 ∈ Tp M lying in an open set which is mapped diffeomorphically by expp onto an
open neighborhood of p.
d(γ(t), q) = a − t (∗)
Note first that d(γ(t), q) ≥ a−t (because if d(γ(t), q) < a−t then d(p, q) ≤ d(p, γ(t))+
d(γ(t), q) < t+a−t = a). Next note that if (∗) holds to t0 ∈ [0, a] it holds for t ∈ [0, t0 ],
because if t ∈ [0, t0 ],
Let t0 = sup{t ∈ [0, a] | (∗) holds for t}. By continuity, (∗) also holds for t0 . We will
show
(1) t0 ≥ ε,
68
Hence a = ε + d(m, q), a − ε = d(m, q) = d(γ(ε), q), and (∗) holds for t = ε.
To prove (2), construct a sphere S about γ(t0 ) as we did for p, and let m be a point in
S of minimal distance from q. Then d(γ(t0 ), q) = inf{d(γ(t0 ), r) + d(r, q) | r ∈ S} =
ε + d(m, q), so a − t0 = ε + d(m, q), so d(m, q) = a − (t0 + ε).
(c)⇒(a): This follows from the theory of ordinary differential equations, as we will
see next time.
69
Chapter 13
To complete the proof of the Hopf-Rinow theorem, we need to show that if (M, d) is
a complete metric space, then geodesics on M can be extended indefinitely.
on the tangent bundle T M . Since geodesics do not depend upon the choice of coor-
dinates on M , neither can X. X is called the geodesic spray.
70
Let {ϕt : t ∈ R} be the local one-parameter group of diffeomorphisms corresponding
to X. It follows from the theory of ordinary differential equations (see 240A, Lecture
10) that given any v0 ∈ T 1 M , there is an open neighborhood U of v0 ) in T 1 M and
an ε > 0 such that
Indeed, since X is never zero, it follows from a Lemma in Lecture 11 that there are
coordinates (y 1 , . . . , y 2n−1 ) on a neighborhood U of v0 such that
∂
X= on U.
∂y 1
From this we see that U is decomposed into a disjoint union of integral curves for X.
We can regard γ ∗ T M as the total space of a vector bundle over [a, b], with projection
π : γ ∗ T M → [a, b] defined by π(t, v) = t. Then a vector field X along γ is a section
of γ ∗ T M ; that is, a smooth map X : [a, b] → γ ∗ T M such that π ◦ X = id.
71
We say that a vector field X along γ is parallel if
∇γ 0 X ≡ 0.
P i ∂
Proof. Suppose v = a ∂xi . Then (∗) is equivalent to the linear initial value
problem
df i X i dxj k
+ Γjk f = 0, f i (t0 ) = ai ,
dt dt
which has a unique solution on [a, b] by the theory of ordinary differential equations.
τγ : Tγ(a) M → Tγ(b) M
by τγ (v) = X(b), where X is the unique parallel vector field along γ such that
X(a) = v. Similarly, we can define τγ if γ is only piecewise smooth. We call τγ
parallel translation along γ.
72
so hX, Y i is constant along γ. Thus τγ : Tγ(a) M → Tγ(b) M is an isometry of inner
product spaces.
73
Chapter 14
Linear ordinary differential equations are easier to solve than nonlinear ordinary dif-
ferential equations. Given a solution to a nonlinear ordinary differential equation,
one often studies nearby solutions by means of “linearization.” The linearization of
the nonlinear geodesic equation is the Jacobi equation.
Suppose that Map([a, b], M ) denotes the space of smooth maps γ : [a, b] → M and
that
Geod([a, b], M ) = {γ ∈ Map([a, b], M ) | γ is a geodesic}.
By completing with respect to a suitable Banach space norm, we can make Map([a, b], M )
into an infinite-dimensional smooth manifold.
The tangent space of Map([a, b], M ) is the space of sections of γ ∗ T M , and we ask:
What is the tangent space to Geod([a, b], M )? To answer this question, we study
“deformations through geodesics.”
Let α : (−ε, ε) → M be a smooth map. then for each s ∈ (−ε, ε), we have a smooth
curve ᾱ(s) : [a, b] → M defined by ᾱ(s)(t). We use the notation ∂α∂s
(s0 , t0 ) = (tangent
∂α
vector to s 7→ α(s, t0 ) at s = s0 ), ∂t (s0 , t0 ) = (tangent vector to t 7→ α(s0 , t) at
t = t0 ). Although ∂α ∂αand
∂s
∂α
∂t
are only defined along the image of α, their integral
∂α
curves commute on ∂s , ∂t along the image of α. We can let
74
which is the total space of a vector bundle over (−ε, ε) × [a, b]. If we let Γ(α∗ T M )
denote the space of sections of α∗ T M , then
∂α ∂α
, ∈ Γ(α∗ T M ).
∂s ∂t
We are interested in the case where γ(t) = α(0, t) is a geodesic. We let
∂α
X(t) = (0, t) = deformation vector field.
∂s
Note that X is a vector field along γ, so X ∈ Γ(γ ∗ T M ).
Proposition. If each ᾱ(s) is a geodesic, then X will satisfy the Jacobi equation
∇γ 0 ∇γ 0 X + R(X, γ 0 )γ 0 = 0.
2. ∇X (f σ + τ ) = f ∇X σ + (Xf )σ + ∇X τ
for X, Y ∈ X(M ), σ, τ ∈ Γ(E), f ∈ F (M ).
∇ ∂ σ = ∇ ∂α σ, ∇ ∂ σ = ∇ ∂α σ.
∂s ∂s ∂t ∂t
75
This is called the pullback connection.
a first order linear system. This system has a linear space of solutions of dimension
2n. The Jacobi fields which vanish at a given point form a linear subspace of dimen-
sion n.
or equivalently, (
d2 f 1
dt2
= 0,
d2 f i
dt2
= −kf for 2 ≤ i ≤ n.
The solutions are
f 1 (t) = a1 + b1 t,
i
√ i
√
a cos( kt) + b sin( kt)
if k > 0,
i i i
f (t) = a + b t if k = 0,
i √ i
√
a cosh( −kt) + b sinh( −kt) if k < 0.
76
Here (a1 , b1 , . . . , an , bn ) are constants of integration to be determined by the initial
conditions.
f 1 (t) = b1 t,
i
√
b sin( kt)
if k > 0,
i
f (t) = bi t if k = 0,
i √
b sinh( −kt) if k < 0.
Example. Antipodal points on S n (1) are conjugate along the geodesics which join
them, while H n (1) and E n do not have conjugate points. Great circles which start
at he north pole p focus at the south pole q.
77
Chapter 15
⇒: Conversely, if X is a nonzero Jacobi field along γv with X(0) = 0, then (∇γ 0 X)(0) =
w for some nonzero w ∈ Tp M , and
∂α
X(t) = (0, t),
∂s
where α : (−ε, ε) × [0, 1] → M is the deformation defined by
α(s, t) = expp (t(v + sw)).
78
∂α
Then expp∗v (w) = ∂s
(0, 1) = X(1) = 0, so expp∗ is singular at v.
Proof. Let p ∈ M , v ∈ Tp M , γ(t) = expp (tv). We need to show that p = γ(0) and
q = expp (v) = γ(1) are not conjugate along γ.
Suppose, on the contrary, that X is a nonzero Jacobi field along γ which vanishes at
γ(0) and γ(1). Then
∇γ 0 ∇γ 0 X + R(X, γ 0 )γ 0 = 0,
h∇γ 0 ∇γ 0 X, Xi = −hR(X, γ 0 )γ 0 , Xi ≥ 0.
Pn i
Write X = i=1 f Ei , where (E1 , E2 , . . . , En ) is a parallel orthonormal from field
along γ. Then the above inequality becomes
n
X d2 f i
fi ≥ 0.
i=1
dt2
79
Since X(0) = X(1) = 0, f i (0) = f i (1) = 1, and we can integrate by parts
Z 1 X d2 f i
Z 1 X df i 2
0≤ fi 2 dt = − dt le0.
0 dt 0 dt
df i
We conclude that dt
≡ 0, f i ≡ 0, and X ≡ 0.
80
Chapter 16
March 2, 2006
One of the most studied problems in Riemannian geometry is the relationship be-
tween curvature and topology. To discuss this problem requires an understanding of
some of the basic topological invariants: the cohomology groups H k (M ; R) that we
saw in 240A and the homotopy groups πn (M ).
It is not hard to check that ' is an equivalence relation: γ ' γ, γ ' λ ⇒ λ ' γ,
γ ' λ and λ ' µ ⇒ γ ' µ. Let π(X, x0 , x1 ) denote the set of equivalence classes, [γ]
the equivalence class of γ.
81
If [γ] ∈ π(X, x0 , x1 ) and [λ] ∈ π(X, x1 , x2 ), we define
One can check that this product is well-defined on equivalence classes: γ ' γ 0 and
λ ' λ0 implies γ · λ ' γ 0 · λ0 .
The fundamental groupoid of X is the “category” whose objects are the points of
X and whose morphisms from x0 to x1 are the elements of π(X, x0 , x1 ). The funda-
mental group of X at x0 is π(X, x0 ) = π(X, x0 , x0 ).
To check that the fundamental groupoid is really a category, we need to verify asso-
ciativity:
([γ] · [λ]) · [µ] = [γ] · ([λ] · [µ]) (∗)
and existence of an identity [εx0 ] ∈ π(X, x0 ), for each x0 ∈ X, such that
when the products are defined. To establish (∗), we need to show that
(γ · λ) · µ ' γ · (λ · µ),
Explicit homotopies are given in Spanier, Algebraic Topology, Chapter 1, and many
other places.
To make π(X, x0 ) into a group, we need to define an inverse to [γ] ∈ π(X, x0 ). This
is the equivalence class of γ −1 ∈ Ω(X, x0 , x0 ), where
82
One verifies that γ · γ −1 ' εx0 and γ −1 · γ ' εx0 , so [γ] · [γ −1 ] = [εx0 ] = [γ −1 ] · [γ].
We have a covariant functor from pointed topological spaces to groups such that
(X, x0 ) 7→ π(X, x0 ),
(F : (X, x0 ) → (Y, F (x0 ))) 7→ (F∗ : π(X, x0 ) → π(Y, F (x0 ))).
The fundamental group π1 (X, x0 ) is actually the first of a series of homotopy groups
πn (X, x0 ), n = 1, 2, 3 . . . To define the other homotopy groups, one makes
Ω(X, x0 ) = Ω(X, x0 , x0 )
One then defines πn (X, x0 ) to be πn−1 (Ω(X, x0 ), εx0 ). This gives a series of groups
which are usually very difficult to calculate. Thus it is still not known how to calcu-
late πn (S m , x0 ) for all values of n and m, for example.
83
Let X
e and X be metric spaces. A covering from X e to X is a surjective continuous
map π : Xe → X such that any p ∈ X lies in an open neighborhood U which is evenly
covered; even covered means that π −1 (U ) is a disjoint union of open sets each of
which is mapped homeomorphically by π onto U .
To prove this, we cover the image of γ with a finite collection of evenly covered open
sets U1 , U2 , . . . , Un . We then lift γ inductively over U1 , U2 , . . . using the defining prop-
erty of evenly covered sets.
84
connected, there is a homotopy α : [0, 1] × [0, 1] → X such that
We now complete the proof of the Hadamard-Cartan Theorem. This will follow from:
Locally, expp is an isometry from (Tp M, hh , ii) to (M, h , i), and it takes lines through
the origin in Tp M to geodesics in M . Thus lines through the origin must be geodesics
in the Riemannian manifold (Tp M, hh , ii). It follows from the Hopf-Rinow Theorem
that (Tp M, hh , ii) is complete. Thus the Proposition will be a consequence of
Proof. One easily checks that π is surjective. Let q ∈ M . We need to show that
q possesses an open neighborhood U which is evenly covered by π. There exists
ε > 0 such that expq maps the open ball of radius 2ε diffeomorphically onto {r ∈
85
M | d(r, q) < ε}. Let {e qα : α ∈ A} = π −1 (q), U = {r ∈ M | d(r, q) < ε},
eα = {e
U r∈M f | d(e
r, qeα ) < ε}. We have a commutative diagram
π∗
Tqeα M
f / Tq M
isometry
exp qeα expq
/M
M
f π
If re ∈ U
eα ∩ U
eβ , we would have geodesics γ eα , γ
eβ of length less than ε from qeα and qeβ
to re. These would project to geodesics γα and γβ from q to r = π(e r) of length less
than ε. By uniqueness of geodesics emanating from q in a normal coordinate chart,
we would have γα = γβ . Since π is a local isometry, γ eα and γeβ would satisfy the same
initial conditions at r. thus γ
eα = γ
eβ and qeα = qeβ . We have proven that U eα ∩ Ueβ 6= ∅,
so α = β.
If re ∈ π −1 (U ), r = π(e
r) ∈ U , and there exists a geodesic γ from r to q of length less
than ε. γ lifts to a geodesic from re to qeα of length less than ε for some qeα ∈ π −1 (q).
Hence re ∈ U eα for some α ∈ A, and π −1 (U ) = S{Uα : α ∈ A}.
86
Chapter 17
March 7, 2006
for x ∈ Tp M , x 6= 0.
87
The scalar curvature s : M → R is defined by
n
X
s(p) = Ric(ei , ei ),
i=1
To prove this theorem, we need to take the “second derivative” of the action function
J : Ω(M ; p, q) → R.
Here
Ω(M ; p, q) = {smooth paths γ : [0, 1] → M | γ(0) = p, γ(1) = q}
and Z 1
1
J(γ) = hγ 0 (t), γ 0 (t)i dt.
2 0
88
is smooth, we say that ᾱ or α is a deformation of γ with deformation field
X(t) = ∂α
∂s
(0, t). The second derivative is defined by requiring
d2
2
d E(f )(X, X) = 2 J(ᾱ(s))
ds s=0
d2
So we need to calculate ds J(ᾱ(s)) s = 0, and we find that
2
d2 d2 1 1 ∂α ∂α
Z
J(ᾱ(s)) = , dt
ds2
s=0 ds 2 2
0 ∂t ∂t
Z 1 s=0
∂ ∂α ∂α
= ∇ ∂α , dt
0 ∂s
∂s ∂t ∂t
Z 1 s=0
∂α ∂α ∂α ∂α
= ∇ ∂α ∇ ∂α , + ∇ ∂α , ∇ ∂α dt
∂s ∂s ∂t ∂t ∂s ∂t ∂s ∂t
0
Z 1 s=0
∂α ∂α ∂α ∂α
= ∇ ∂α ∇ ∂α , + ∇ ∂α , ∇ ∂α dt
∂s ∂t ∂s ∂t ∂t ∂s ∂t ∂s
0
Z 1 s=0
∂α ∂α ∂α ∂α ∂α ∂α
= R , , + ∇ ∂α ∇ ∂α ,
0 ∂s ∂t ∂s ∂t ∂t ∂s ∂s ∂t
∂α ∂α
+ ∇ ∂α , ∇ ∂α dt
∂t ∂s ∂t ∂s
Z 1 s=0
∂α ∂α ∂α ∂α ∂ ∂α ∂α
= R , , + ∇ ∂α ,
0 ∂s ∂t ∂s ∂t ∂t ∂s ∂s ∂t
∂α ∂α ∂α ∂α
− ∇ ∂α , ∇ ∂α + ∇ ∂α , ∇ ∂α dt .
∂s ∂s ∂t ∂t ∂t ∂s ∂t ∂s
s=0
89
We can integrate by parts to obtain
Z 1
2
(d J)(γ)(X, Y ) = − h∇γ 0 ∇γ 0 X + R(X, γ 0 )γ 0 , Y i dt. (∗∗)
0
Proof of Myers’s Theorem. Suppose that p, q ∈ M and d(p, q) > πa. Let γ :
[0, 1] → M be a geodesic which minimizes J. Since |γ 0 | is constant, |γ 0 | = d(p, q).
1
Choose parallel orthonormal vector fields (E1 , . . . , En ) along γ such that E1 = d(p,q) γ0.
Let Xi (t) = (sin πt)Ei (t) for i = 2, . . . , n. Then Xi (0) = 0 = Xi (1), so Xi ∈
Tγ Ω(M ; p, q). Then
∇γ 0 Xi = −π cos πt Ei ,
∇γ 0 ∇γ 0 Xi = −π 2 sin πt Ei ,
h∇γ 0 ∇γ 0 Xi , Xi i = −π 2 sin2 πt.
Hence
n
X Z 1
2
d J(γ)(Xi , Xi ) = sin2 πt[(n − 1)π 2 − d(p, q)2 Ric(E1 , E1 )] dt.
i=2 0
But
d(p, q) > πa ⇒ d(p, q)2 > π 2 a2
⇒ d(p, q)2 Ric(E1 , E1 ) > (n − 1)π 2 .
Ric(E1 , E1 ) ≥ n−1
a2
Hence n
X
d2 J(γ)(Xi , Xi ) < 0 and d2 J(γ)(Xi , Xi ) < 0
i=2
90
for some i, 2 ≤ i ≤ n. This contradicts minimality of the geodesic. Thus it must be
the case that d(p, q) ≤ πa.
It follows that M is compact because it lies in the image under the exponential map
of the ball of radius πa.
To prove that π1 (M ) is finite, we use the fact that any connected smooth manifold
M has a universal cover M f: a simply connected manifold such that π : M f → M is
a smooth covering. Give M f the metric π ∗ h , i, where h , i is the Riemannian metric
on M . Then Mf is also complete and has Ricci curvature satisfying (∗). Hence M f is
−1 −1
also compact, and if p ∈ M , π (p) must be a finite set. But the points of π (p) are
in one-to-one correspondence with the elements of π1 (M, p). Thus π1 (M ) is finite.
91
Chapter 18
March 9, 2006
In the proof of Myers’s Theorem, we used the following theorem from topology:
We call M
f the universal cover of M .
92
This example shows that universal covers can sometimes be used to calculate funda-
mental groups.
π1 (S 1 × S n ) ∼
= Z for n ≥ 2. Hence S 1 × S 2 does not have a metric of positive Ricci
curvature by Myers’s Theorem. It therefore does not have a metric of positive sec-
tional curvature.
Can we find further relationships between curvature and the fundamental group? The
answer is “yes,” and a famous example is
93
The proof of Synge’s theorem is based upon the notion of “free homotopy class.”
Let Map(S 1 , M ) = {continuous maps γ : S 1 → M }, where S 1 = [0, 1] with the
points 0 and 1 identified. We say that two elements γ, λ ∈ Map(S 1 , M ) are freely
homotopic, and write γ ' λ, if there is a continuous map
(
α(0, t) = γ(t),
α : [0, 1] × S 1 → M such that
α(1, t) = λ(t).
Free homotopy is just homotopy without regard to base points. It is easily verified
that ' is an equivalence relation. We let [S 1 , M ] be the set of equivalence classes and
write [γ] for the equivalence class of γ ∈ Map(S 1 , M ).
It is not too difficult to show that [S 1 , M ] is in one-to-one correspondence with the set
of conjugacy classes in π1 (M ), when M is a connected manifold. Thus a connected
manifold is simply connected iff [S 1 , M ] consists of a single point.
94
Let us assume the Lemma for now and prove Synge’s Theorem with its help. The
proof uses the “second derivative” of J in a manner quite similar to that employed
in Myers’s Theorem.
then α is smooth.
d2
J(ᾱ(s)) = d2 J(γ)(X, X).
ds2
s=0
he steps are virtually the same, except that we do not have to worry about endpoints.
The result is
Z 1
2
d J(γ)(X, X) = [h∇γ 0 X, ∇γ 0 Xi − hR(X, γ 0 )γ 0 , Xi] dt. (∗)
0
95
If γ : S 1 → M minimizes in its free homotopy class, we must have d2 J(γ)(X, X) ≥ 0
for all X ∈ Tγ Map(S 1 , M ).
Let us now turn to part 1 of Synge’s Theorem. If M is not simply connected, the
Lemma says there is a smooth closed geodesic γ : S 1 → M which minimizes in its free
homotopy class. Let p = γ(0). Parallel translation around γ gives an isomorphism
τγ : Tp M → Tp M .
τγ is defined by τγ (x) = X(1), where X(t) is the unique section of γ ∗ T M → [0, 1] such
that ∇γ 0 X ≡ 0 and X(0) = x. Note that |τγ (x)| = |x|; that is, τγ is an orthogonal
transformation of Tp M . Since M is oriented, τγ preserves orientation.
A canonical form theorem for elements of SO(n) states that after conjugation, we can
arrange that τγ is represented by a matrix of the form
±1
±1
..
.
±1
A= cos θ1 − sin θ1 .
sin θ1 cos θ1
..
.
cos θk − sin θk
sin θk cos θk
Since the determinant of A is one, the number of (−1)’s is even. Since dim M is
even the number of (+1)’s is also even. There must be a +1 that corresponds to
1
e1 = L(γ) γ 0 , and hence there must be at least one more +1 corresponding to e2 . Thus
there is a unit-length vector field E2 along γ such that ∇γ 0 E2 ≡ 0, E2 (1) = E2 (0),
96
and E2 ⊥ E1 . Then
Z 1
2
d J(γ)(E2 , E2 ) = − hR(E2 , γ 0 )γ 0 , E2 i dt,
0
which is less than 0 since sectional curvatures are assumed positive. This yields a
contradiction. Hence M must be simply connected.
The proof of part 2 is similar; one uses a smooth closed geodesic γ : S 1 → M such
that parallel translation around γ is orientation-reversing.
97
Chapter 19
1 1 0
Z
J(γ) = hγ (t), γ 0 (t)i dt.
2 0
98
Let N be a large integer, and for integers i such that 0 ≤ i ≤ N , let ti = Ni . Let
BGN (S 1 , M ) be the set of continuous maps γ : [0, 1] → M with γ(0) = γ(1) and
γ |[ti−1 ,ti ] is a smooth constant speed geodesic for 1 ≤ i ≤ N } and
We have a map
h : BGN (S 1 , M )c → M
| × ·{z
· · × M}
N
p
We set f (t) = hγ 0 (t), γ 0 (t)i and g(t) = 1. Then
Z b 2 Z b Z b
p 0 0
hγ 0 (t), γ 0 (t)i dt ≤ hγ (t), γ (t)i dt a dt,
a a a
Thus if J(γ) ≤ c, √
L(γ) ≤ b − a · c. (∗)
By taking b − a sufficiently small, we can arrange that L(γ) < δ where δ > 0 is any
prescribed number. Since M is compact, there exists δ > 0 such that if p, q ∈ M ,
d(p, q) < δ, p and q are connected by a unique geodesic γp,q of length d(p, q) which
depends smoothly on p and q. Indeed, we can choose δ > 0 so that if Ω(M ; p, q)2δ =
99
{piecewise smooth γ : [0, 1] → M such that γ(0) = p, γ(1) = q, and L(γ) ≤ 2δ} with
the topology defined by the metric
q
1
We can choose N so that N
· c < δ, so that it follows from (∗) that
L γ |[ i−1 , i ] < δ.
N N
γ i−1 i
1
N
to γ N
. This defines a map g : Map(S , M ))c → BGN (S 1 , M )c . In fact,
with the metric space topology defined above, the homotopy H shows that g is a
“homotopy equivalence.”
We call γ
e a broken geodesic and regard it as an approximation to γ.
100
b : S 1 → M with J(b
γ γ ) < µ by smoothing off the corner. Hence γ is a smooth closed
geodesic with J(γ) = µ.
A similar argument could be used to give an alternate proof of the following theorem
from Lecture 12:
101
Chapter 20
so 2
∂xi
det(e
gαβ ) = det(gij ) det ,
∂y α
so i
∂x
q q
det(e
gαβ ) = det(gij ) det .
∂y α
It follows that if M is oriented and the coordinate systems (x1 , . . . , xn ) and (y 1 , . . . , y n )
are positively oriented,
q p
det(gij ) dx1 ∧ · · · ∧ dxn = det geαβ dy 1 ∧ · · · ∧ dy n .
On a positively oriented Riemannian manifold (M, h , i), we can thus define a volume
form Θ by q
Θ= det(gij ) dx1 ∧ · · · ∧ dxn ,
102
whenever (x1 , . . . , xn ) are positively oriented coordinates on M and h , i = gij dxi ⊗
P
dxj .
h , i : Te G × Te G → R
The conclusion is that any compact Lie group has a biinvariant Riemannian metric.
For example, if G = O(n),
o(n) = TI O(n) ∼
= {n × n matrices A | A + AT = 0},
103
and hhA, Bii =trace(AT B) defines a biinvariant metric on O(n).
It follows from Lecture 9 that if G is a compact Lie group with a biinvariant Rieman-
nian metric h , i, then the curvature of G is given by
1
hR(X, Y )W, Zi = h[X, Y ], [Z, W ]i
4
for X, Y, Z, W ∈ g, the Lie algebra of G. Thus curvature is given by a simple explicit
formula in this case.
(i∗ T M )p = T pM0 ⊕ Np M0 ,
where (·)T and (·)⊥ are orthogonal projections to tangent and normal space, respec-
tively.
104
The proof is exactly the same as before (when M was Euclidean space).
The symmetric bilinear form α : X(M0 ) × X(M0 ) → X(M0 )⊥ is called the second
fundamental form of M0 in M . Note that α is defined pointwise and restricts to a
bilinear map
α : Tp M0 × Tp M0 → Np M0
for each p ∈ M0 .
hR0 (x, y)w, zi = hR(x, y)w, zi + hα(z, z), α(y, w)i − hα(x, w), α(y, z)i, (∗)
for x, y, z, w ∈ Tp M .
Indeed, we easily see that M0 ⊆ M is totally geodesic iff geodesics in M0 are also
geodesics in M .
105
If M0 ⊆ M is totally geodesic, the Gauß equation (∗) becomes
Choose ε > 0 sufficiently small that expp maps Bε diffeomorphically onto an open
subset U of p in M . Since F is an isometry, it takes geodesics to geodesics. Hence if
v ∈ Tp M , γ(t) = F (expp (tv)) is a geodesic in M and γ 0 (0) = F∗p (v). Thus
and hence
F ◦ expp = expp ◦F∗p : Bε → U.
Let W = {v ∈ Tp M | F∗p (v) = v}, a linear subspace of Tp M . Moreover,
ϕ : exp−1
p : U → Tp M
We now return to the example of a compact Lie group with a biinvariant Riemannian
metric h , i. Can we construct isometries of G which have interesting fixed point
106
sets?
(ii) s2 = id.
Let K = {σ ∈ G : s(σ) = σ}, a compact subgroup of G. Of course, K is a totally
geodesic submanifold of G, but we can construct an even more interesting one as
follows:
M0 = {σ ∈ G | F (σ) = σ} = {σ ∈ G | s(σ) = σ −1 },
β : G × M0 → M0 by β(σ, τ ) = στ s(σ −1 ).
(Note that s(στ s(σ −1 )) = s(σ)s(τ ) · σ −1 = s(σ)τ −1 σ −1 = (στ s(σ −1 ))−1 , so β(σ, τ ) ∈
M0 .) The map τ 7→ β(σ, τ ) is an isometry for each σ ∈ G. Note that β(σ, e) = e iff
σs(σ −1 ) = e iff e ∈ K. This implies that β induces a map h : G/K → M0 by
h(σK) = σs(σ −1 ).
One can check that h is a bijection. Hence we havegiven a manifold structure to the
quotient G/K and have a totally geodesic imbedding of G/K in G.
107
Then K = {A ∈ P (p) × O(q) : det A = 1} and G/K is the Grassmann manifold of
p-planes in Rp+q . Thus the Grassmann manifold is a totally geodesic submanifold of
SO(p + q).
This construction gives a large class of manifolds for which we can actually calculate
the curvature quite explicitly.
108
Chapter 21
Problems
√
1. Let U = {(r, θ) : r > 0, 0 < θ < 2π}. Define F : U → R3 by
√
r cos(√ 2θ)
1
F (r, θ) = √ r sin( 2θ) .
2 r
h , i = dr ⊗ dr + r2 dθ ⊗ dθ.
(b) Show that if we let x = r cos θ, y = r sin θ, the metric takes the form
h , i = dx ⊗ dx + dy ⊗ dy.
(a) Let (e1 , e2 ) be an orthonormal basis for Tp M . Use the Gauß equations to
show that
h(e1 , e1 ) h(e1 , e2 )
K(p) =
.
h(e2 , e1 ) h(e2 , e2 )
109
(b) By citing an appropriate theorem from linear algebra, show that there is
an orthonormal basis for Tp M such that h(e1 , e2 ) = 0.
(c) If (e1 , e2 ) is such an orthonormal basis, h(e1 , e1 ) and h(e2 , e2 ) are called
principal curvatures of M 2 at p, and are denoted by κ1 (p) and κ2 (p).
The Gaussian curvature is then given by the formula K(p) = κ1 (p)κ2 (p).
Suppose that F : R × S 1 → R3 by
3. Suppose that (M, h i) is a connected Riemannian manifold which has the fol-
lowing property: Given p, q ∈ M n , there is an isometry F : M → M such that
F (p) = q. Show that (M n , h i) is geodesically complete.
5. Let F : (0, ∞) × S 1 → R3 by
u √
Z
iv −u −u
F (u, e ) = e cos v, e sin v, 1 − e−2w dw .
1
(a) Sketch M 2 .
(b) Calculate the coefficients of the metric induced by F on (0, ∞) × S 1 .
(c) Calculate the second fundamental form.
(d) Determine the Gaussian curvature κ.
110
6. (a) Let (M, h , i) be a Riemannian manifold, C a one-dimensional submani-
fold which is the fixed-point set of an isometry. Show that C consists of
geodesics.
(b) Suppose that M = {(x, y) ∈ R2 | x2 + y 2 < 1} and
4
h, i= (dx ⊗ dx + dy ⊗ dy).
(1 − x2 − y 2 )2
111