Você está na página 1de 107

Riemannian Geometry

it is a draft of Lecture Notes of H.M. Khudaverdian.


Manchester, 20-th May 2011
Contents
1 Riemannian manifolds 4
1.1 Manifolds. Tensors. (Recalling) . . . . . . . . . . . . . . . . . 4
1.2 Riemannian manifoldmanifold equipped with Riemannian
metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2.1

Pseudoriemannian manifold . . . . . . . . . . . . . . 11
1.3 Scalar product. Length of tangent vectors and angle between
vectors. Length of the curve . . . . . . . . . . . . . . . . . . . 11
1.3.1 Length of the curve . . . . . . . . . . . . . . . . . . . . 12
1.4 Riemannian structure on the surfaces embedded in Euclidean
space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4.1 Internal and external coordinates of tangent vector . . 15
1.4.2 Explicit formulae for induced Riemannian metric (First
Quadratic form) . . . . . . . . . . . . . . . . . . . . . . 17
1.4.3 Induced Riemannian metrics. Examples. . . . . . . . . 20
1.4.4

Induced metric on two-sheeted hyperboloid embedded
in pseudo-Euclidean space. . . . . . . . . . . . . . . . . 27
1.5 Isometries of Riemanian manifolds. . . . . . . . . . . . . . . . 28
1.5.1 Examples of local isometries . . . . . . . . . . . . . . . 29
1.6 Volume element in Riemannian manifold . . . . . . . . . . . . 31
1.6.1 Volume of parallelepiped . . . . . . . . . . . . . . . . . 31
1.6.2 Invariance of volume element under changing of coor-
dinates . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
1.6.3 Examples of calculating volume element . . . . . . . . 34
1
2 Covariant dierentiaion. Connection. Levi Civita Connec-
tion on Riemannian manifold 36
2.1 Dierentiation of vector eld along the vector eld.Ane
connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
2.1.1 Denition of connection. Christoel symbols of con-
nection . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.1.2 Transformation of Christoel symbols for an arbitrary
connection . . . . . . . . . . . . . . . . . . . . . . . . . 39
2.1.3 Canonical at ane connection . . . . . . . . . . . . . 40
2.1.4

Global aspects of existence of connection . . . . . . . 43
2.2 Connection induced on the surfaces . . . . . . . . . . . . . . . 44
2.2.1 Calculation of induced connection on surfaces in E
3
. . . 45
2.3 Levi-Civita connection . . . . . . . . . . . . . . . . . . . . . . 47
2.3.1 Symmetric connection . . . . . . . . . . . . . . . . . . 47
2.3.2 Levi-Civita connection. Theorem and Explicit formulae 47
2.3.3 Levi-Civita connection on 2-dimensional Riemannian
manifold with metric G = adu
2
+ bdv
2
. . . . . . . . . . 49
2.3.4 Example of the sphere again . . . . . . . . . . . . . . . 49
2.4 Levi-Civita connection = induced connection on surfaces in E
3
50
3 Parallel transport and geodesics 52
3.1 Parallel transport . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.1.1 Denition . . . . . . . . . . . . . . . . . . . . . . . . . 52
3.1.2

Parallel transport is a linear map . . . . . . . . . . . 53
3.1.3 Parallel transport with respect to Levi-Civita connection 54
3.2 Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
3.2.1 Denition. Geodesic on Riemannian manifold. . . . . . 54
3.2.2 Un-parameterised geodesic . . . . . . . . . . . . . . . . 55
3.2.3 Geodesics on surfaces in E
3
. . . . . . . . . . . . . . . 56
3.3 Geodesics and Lagrangians of free particle on Riemannian
manifold. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
3.3.1 Lagrangian and Euler-Lagrange equations . . . . . . . 58
3.3.2 Lagrangian of free particle . . . . . . . . . . . . . . . 58
3.3.3 Equations of geodesics and Euler-Lagrange equations . 59
3.3.4 Examples of calculations of Christoel symbols and
geodesics using Lagrangians. . . . . . . . . . . . . . . . 60
3.3.5 Variational principe and Euler-Lagrange equations . . 62
3.4 Geodesics and shortest distance. . . . . . . . . . . . . . . . . . 63
2
3.4.1 Again geodesics for sphere and Lobachevsky plane . . . 65
4 Surfaces in E
3
67
4.1 Formulation of the main result. Theorem of parallel transport
over closed curve and Theorema Egregium . . . . . . . . . . . 67
4.1.1 GauTheorema Egregium . . . . . . . . . . . . . . . . . 69
4.2 Derivation formulae . . . . . . . . . . . . . . . . . . . . . . . 70
4.2.1

Gauss condition (structure equations) . . . . . . . . . 72
4.3 Geometrical meaning of derivation formulae. Weingarten op-
erator and second quadratic form in terms of derivation for-
mulae. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
4.3.1 Gaussian and mean curvature in terms of derivation
formulae . . . . . . . . . . . . . . . . . . . . . . . . . . 75
4.4 Examples of calculations of derivation formulae and curvatures
for cylinder, cone and sphere . . . . . . . . . . . . . . . . . . 76
4.5

Proof of the Theorem of parallel transport along closed curve. 80
5 Curvtature tensor 84
5.1 Denition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
5.1.1 Properties of curvature tensor . . . . . . . . . . . . . . 85
5.2 Riemann curvature tensor of Riemannian manifolds. . . . . . . 86
5.3

Curvature of surfaces in E
3
.. Theorema Egregium again . . . 87
5.4 Relation between Gaussian curvature and Riemann curvature
tensor. Straightforward proof of Theorema Egregium . . . . . 88
5.4.1

Proof of the Proposition (5.25) . . . . . . . . . . . . . 90
5.5 Gauss Bonnet Theorem . . . . . . . . . . . . . . . . . . . . . . 94
6 Appendices 97
6.1

Integrals of motions and geodesics. . . . . . . . . . . . . . . . 97
6.1.1

Integral of motion for arbitrary Lagrangian L(x, x) . . 97
6.1.2

Basic examples of Integrals of motion: Generalised
momentum and Energy . . . . . . . . . . . . . . . . . . 97
6.1.3

Integrals of motion for geodesics . . . . . . . . . . . . 98
6.1.4

Using integral of motions to calculate geodesics . . . . 100
6.2 Induced metric on surfaces. . . . . . . . . . . . . . . . . . . . 101
6.2.1 Recalling Weingarten operator . . . . . . . . . . . . . . 101
6.2.2 Second quadratic form . . . . . . . . . . . . . . . . . . 102
6.2.3 Gaussian and mean curvatures . . . . . . . . . . . . . . 103
3
6.2.4 Examples of calculation of Weingarten operator, Sec-
ond quadratic forms, curvatures for cylinder, cone and
sphere. . . . . . . . . . . . . . . . . . . . . . . . . . . . 103
1 Riemannian manifolds
1.1 Manifolds. Tensors. (Recalling)
I recall briey basics of manifolds and tensor elds on manifolds.
An n-dimensional manifold is a space such that in a vicinity of any point
one can consider local coordinates {x
1
, . . . , x
n
} (charts). One can consider
dierent local coordinates. If both coordinates {x
1
, . . . , x
n
}, {y
1
, . . . , y
n
} are
dened in a vicinity of the given point then they are related by bijective
transition functions (functions dened on domains in R
n
and taking values
in R
n
).
_

_
x
1

= x
1

(x
1
, . . . , x
n
)
x
2

= x
2

(x
1
, . . . , x
n
)
. . .
x
n1

= x
n1

(x
1
, . . . , x
n
)
x
n

= x
n

(x
1
, . . . , x
n
)
We say that manifold is dierentiable or smooth if transition functions are
dieomorphisms, i.e. they are smooth and rank of Jacobian is equal to k, i.e.
det
_
_
_
_
_
x
1

x
1
x
1

x
2
. . .
x
1

x
n
x
2

x
1
x
2

x
2
. . .
x
2

x
n
. . .
x
n

x
1
x
n

x
2
. . .
x
n

x
n
_
_
_
_
_
= 0 (1.1)
A good example of manifold is an open domain D in n-dimensional vector
space R
n
. Cartesian coordinates on R
n
dene global coordinates on D. On
the other hand one can consider an arbitrary local coordinates in dierent
domains in R
n
.
E.g. one can consider polar coordinates {r, } in a domain D = {x, y : y >
0} of R
2
(or in other domain of R
2
) dened by standard formulae:
_
x = r cos
y = r sin
,
4
det
_
x
r
x

y
r
y

_
= det
_
cos r sin
sin r cos
_
= r (1.2)
or one can consider spherical coordinates {r, , } in a domain D = {x, y, z : x >
0, y > 0, z > 0} of R
3
(or in other domain of R
3
) dened by standard for-
mulae
_

_
x = r sin cos
y = r sin sin
z = r cos
,
,
det
_
_
_
x
r
x

y
r
y

z
r
z

_
_
_
= det
_
_
sin cos r cos cos r sin sin
sin sin r cos sin r sin cos
cos r sin 0
_
_
= r
2
sin
(1.3)
Choosing domain where polar (spherical) coordinates are well-dened we
have to be award that coordinates have to be well-dened and transition
functions (1.1) have to be dieomorphisms.
Examples of manifolds: R
n
, Circle S
1
, Sphere S
2
, in general sphere S
n
,
torus S
1
S
1
, cylinder, cone, . . . .
We also have to recall briey what are tensors on manifold.
Tensors on Manifold
For every point p on manifold M one can consider tangent vector space
T
p
Mthe space of vectors tangent to the manifold at the point M.
Tangent vector A = A
i
x
i
. Under changing of coordinates it transforms
as follows:
A = A
i

x
i
= A
i
x
m

x
i

x
m

= A
m

x
m

Hence
A
i

=
x
i

x
i
A
i
(1.4)
Consider also cotangent space T

p
M (for every point p on manifold M)
space of linear functions on tangent vectors, i.e. space of 1-forms which
sometimes are called covectors.:
5
One-form (covector) =
i
dx
i
transforms as follows
=
m
dx
m
=
m
x
m
x
m

dx
m

=
m
dx
m

.
Hence

m
=
x
m
x
m

m
. (1.5)
Tensors:
One can consider contravariant tensors of the rank p
T = T
i
1
i
2
...i
p

x
i
1


x
i
2


x
i
k
with components {T
i
1
i
2
...i
k
}.
One can consider covariant tensors of the rank q
S = S
j
1
j
2
...j
q
dx
j
1
dx
j
2
. . . dx
j
q
with components {S
j
1
j
2
...j
q
}.
One can also consider mixed tensors:
Q = Q
i
1
i
2
...i
p
j
1
j
2
...j
q

x
i
1


x
i
2


x
i
k
dx
j
1
dx
j
2
. . . dx
j
q
with components {Q
i
1
i
2
...i
p
j
1
j
2
...j
q
}. We call these tensors tensors of the type
_
p
q
_
.
Tensors of the type
_
p
0
_
are called contravariant tensors of the rank p.
Tensors of the type
_
0
q
_
are called covariant tensors of the rank q.
Having in mind (1.4) and (1.5) we come to the rule of transformation for
tensors of the type
_
p
q
_
:
Q
i

1
i

2
...i

p
j

1
j

2
...j

q
=
x
i

1
x
i
1
x
i

2
x
i
2
. . .
x
i

p
x
i
p
x
j
1
x
j

1
x
j
2
x
j

2
. . .
x
j
q
x
j

q
Q
i
1
i
2
...i
p
j
1
j
2
...j
q
(1.6)
E.g. if S
ik
is a covariant tensor of rank 2 (tensor of the type
_
p
q
_
) then
S
i

k
=
x
i
x
i

x
k
x
k

S
ik
. (1.7)
6
If A
i
k
is a tensor of rank
_
1
1
_
(linear operator on T
p
M) then
A
i

k
=
x
i

x
i
x
k
x
k

A
i
k
Remark Transformations formulae (1.4)(1.7) dene vectors, covectors
and in generally any tensor elds in components. E.g. covariant tensor (co-
variant tensor eld) of the rank 2 can be dened as matrix S
ik
(matrix valued
function S
ik
(x)) such that under changing of coordinates {x
1
, x
2
, . . . , x
n
}
{x
1

, x
2

, . . . , x
n

} S
ik
change by the rule (1.7).
Remark Einstein summation rules
In our lectures we always use so called Einstein summation convention. it
implies that when an index occurs more than once in the same expression in
upper and in.. postitions.., the expression is implicitly summed over all pos-
sible values for that index. Sometimes it is called dummy indices summation
rule.
1.2 Riemannian manifoldmanifold equipped with Rie-
mannian metric
Denition The Riemannian manifold is a manifold equipped with a Rie-
mannian metric.
The Riemannian metric on the manifold M denes the length of the
tangent vectors and the length of the curves.
Denition Riemannian metric G on n-dimensional manifold M
n
denes
for every point p M the scalar product of tangent vectors in the tangent
space T
p
M smoothly depending on the point p.
It means that in every coordinate system (x
1
, . . . , x
n
) a metric G =
g
ik
dx
i
dx
k
is dened by a matrix valued smooth function g
ik
(x) (i = 1, . . . , n; k =
1, . . . n) such that for any two vectors
A = A
i
(x)

x
i
, B = B
i
(x)

x
i
,
tangent to the manifold M at the point p with coordinates x = (x
1
, x
2
, . . . , x
n
)
(A, B T
p
M) the scalar product is equal to:
A, B
G

p
= G(A, B)

p
= A
i
(x)g
ik
(x)B
k
(x) =
7
_
A
1
. . . A
n
_
_
_
_
_
g
11
(x) . . . g
1n
(x)
. . . . . . . . .
g
n1
(x) . . . g
nn
(x)
_
_
_
_
_
_
_
_
_
_
B
1

B
n
_
_
_
_
_
_
(1.8)
where
G(A, B) = G(B, A), i.e. g
ik
(x) = g
ki
(x) (symmetricity condition)
G(A, A) > 0 if A = 0, i.e.
g
ik
(x)u
i
u
k
0, g
ik
(x)u
i
u
k
= 0 i u
1
= = u
n
= 0 (positive-
deniteness)
G(A, B)

p=x
, i.e. g
ik
(x) are smooth functions.
One can say that Riemannian metric is dened by symmetric covariant
smooth tensor eld G of the rank 2 which denes scalar product in the tangent
spaces T
p
M smoothly depending on the point p. Components of tensor eld
G in coordinate system are matrix valued functions g
ik
(x):
G = g
ik
(x)dx
i
dx
k
. (1.9)
The matrix ||g
ik
|| of components of the metric G we also sometimes denote
by G.
Rule of transformation for entries of matrix g
ik
(x)
g
ik
(x)-entries of the matrix ||g
ik
|| are components of tensor eld G in a
given coordinate system.
How do these components transform under transformation of coordinates
{x
i
} {x
i

}?
G = g
ik
dx
i
dx
k
= g
ik
_
x
i
x
i

dx
i

_
x
k
x
k

dx
k

_
=
x
i
x
i

g
ik
x
k
x
k

dx
i

dx
k

= g
i

k
dx
i

dx
k

Hence
g
i

k
=
x
i
x
i

g
ik
x
k
x
k

. (1.10)
8
One can derive transformations formulae also using general formulae (1.7)
for tensors.
Important remark
g
ik
=


x
i
,

x
k

(1.11)
Later by some abuse of notations we sometimes omit the sign of tensor
product and write a metric just as
G = g
ik
(x)dx
i
dx
k
.
Examples
R
n
with canonical coordinates {x
i
} and with metric
G = (dx
1
)
2
+ (dx
2
)
2
+ + (dx
n
)
2
G = ||g
ik
|| = diag [1, 1, . . . , 1]
Recall that this is a basis example of n-dimensional Euclidean space,
where scalar product is dened by the formula:
G(X, Y) = X, Y = g
ik
X
i
Y
k
= X
1
Y
1
+ X
2
Y
2
+ + X
n
Y
n
.
In the general case if G = ||g
ik
|| is an arbitrary symmetric positive-
denite metric then
G(X, Y) = X, Y = g
ik
X
i
Y
k
.
One can show that there exist a new basis {e
i
} such that in this basis
G(e
i
, e
k
) =
ik
.
This basis is called orthonormal basis. (See the Lecture notes in Ge-
ometry)
Scalar product in vector space denes the same scalar product at all
the points. In general case for Riemannian manifold scalar product
depends on the point.
9
R
2
with polar coordinates in the domain y > 0 (x = r cos , y =
r sin ):
dx = cos dr r sin d, dy = sin dr + r cos d. In new coordi-
nates the Riemannian metric G = dx
2
+ dy
2
will have the following
appearance:
G = (dx)
2
+(dy)
2
= (cos drr sin d)
2
+(sin dr+r cos d)
2
= dr
2
+r
2
(d)
2
We see that for matrix G = ||g
ik
||
G =
_
g
xx
g
xy
g
yx
g
yy
_
=
_
1 0
0 1
_
. .
in cartesian coordinates
, G =
_
g
rr
g
r
g
r
g

_
=
_
1 0
0 r
_
. .
in polar coordinates
Circle
Interval [0, 2) in the line 0 x < 2 with Riemannian metric
G = a
2
dx
2
(1.12)
Renaming x we come to habitual formula for metric for circle of
the radius a: x
2
+ y
2
= a
2
embedded in the Euclidean space E
2
:
G = a
2
d
2
_
x = a cos
y = a sin
, 0 < 2, (1.13)
Cylinder surface
Domain in R
2
D = {(x, y): , 0 x < 2 with Riemannian metric
G = a
2
dx
2
+ dy
2
(1.14)
We see that renaming variables x , y h we come to habitual,
familiar formulae for metric in standard polar coordinates for cylinder
surface of the radius a embedded in the Euclidean space E
3
:
G = a
2
d
2
+ dh
2
_

_
x = a cos
y = a sin
z = h
, 0 < 2, < h <
(1.15)
10
Sphere
Domain in R
2
, 0 < x < 2, 0 < y < with metric G = dy
2
+sin
2
ydx
2
We see that renaming variables x , y h we come to habitual,
familiar formulae for metric in standard spherical coordinates for sphere
x
2
+y
2
+z
2
= a
2
of the radius a embedded in the Euclidean space E
3
:
G = a
2
d
2
+a
2
sin
2
d
2
_
x = a sin cos
y = a sin sin z = a cos
, 0 < 2, < h <
(1.16)
1.2.1

Pseudoriemannian manifold
If we omit the condition of positive-deniteness for Riemannian metric we
come to so called Pseudorimannian metric. Manifodl equipped with pseu-
doriemannan metric is called pseudoriemannian manifold. Pseudoriemannian
manifolds appear in applications in the special and general relativity theory.
Example Consider n+1-dimensional linear space R
n+1
with pseudomet-
ric
(dx
0
)
2
(dx
1
)
2
(dx
2
)
2
(dx
n
)
2
in coordinates x
0
, x
1
, . . . , x
n
. In the case n = 3 it is so called Minkovski
space. The coordinate x
0
the role of the time: x
0
= ct, where c is the value
of the speed of the light.
1.3 Scalar product. Length of tangent vectors and an-
gle between vectors. Length of the curve
The Riemannian metric denes scalar product of tangent vectors attached
at the given point. Hence it denes the length of tangent vectors and angle
between them. If X = X
m
x
m
, Y = Y
m
x
m
are two tangent vectors at the
given point p of Riemannian manifold with coordinates x
1
, . . . , x
n
, then we
have that lengths of these vectors equal to
|X| =

X, X =

g
ik
(x)X
i
X
k
, |Y| =

Y, Y =

g
ik
(x)Y
i
Y
k
,
(1.17)
and he angle between these vectors is dened by the relation
cos =
X, Y
|X| |Y|
=
g
ik
X
i
Y
k

g
ik
(x)X
i
X
k

g
ik
(x)Y
i
Y
k
(1.18)
11
Example Let M be 3-dimensional Riemanian manifold. Consider the
vectors X = 2
x
+2
y

z
and Y =
x
2
y
2
z
attached at the point p of
M with local coordinates (x, y, z), where x = y = 1, z = 0. Find the lengths
of these vectors and angle between them if the expression of Riemannian
metric in these coordinates is coordinates
dx
2
+dy
2
+dz
2
(1+x
2
+y
2
)
2
.
We see that matrix of Riemannian g
ik
= (x, y, z)
ik
, where (x, y, z) =
1
(1+x
2
+y
2
+z
2
)
2

ik
= 1 if i = k and
ik
= 0 if i = k, i.e. matrix ||g
ik
|| is
proportional to unity matrix. According to formulae above
|X| =

X, X =

g
ik
(x)X
i
X
k
=

(x, y, z)

X
i
X
i
= 3

(x, y, z) =

3 .
The same answer for |Y|. The scalar product between vectors X, Y equal to
zero:
X, Y = (x, y, z)
ik
X
i
Y
k
= 0
Hence these vectors are orthogonal to each other.
1.3.1 Length of the curve
Let : x
i
= x
i
(t), (i = 1, . . . , n)) (a t b) be a curve on the Riemannian
manifold (M, G).
At the every point of the curve the velocity vector (tangent vector) is
dened:
v(t) =
_
_
_
_
_
_
x
1
(t)

x
n
(t)
_
_
_
_
_
_
(1.19)
The length of velocity vector v T
x
M (vector v is tangent to the manifold
M at the point x) equals to
|v|
x
=

v, v
G

x
=

g
ik
v
i
v
k

x
=

g
ik
dx
i
(t)
dt
dx
k
(t)
dt x

x
(1.20)
The length of the curve is dened by the integral of the length of velocity
vector:
L

b
a

v, v
G

x(t)
dt =

b
a

g
ik
(x(t)) x
i
(t) x
k
(t)dt (1.21)
12
Bearing in mind that metric (1.9) denes the length we often write metric
in the following form
ds
2
= g
ik
dx
i
dx
k
(1.22)
For example consider 2-dimensional Riemannian manifold with metric
||g
ik
(u, v)|| =
_
g
11
(u, v) g
12
(u, v)
g
21
(u, v) g
22
(u, v)
_
.
Then
G = ds
2
= g
ik
du
i
dv
k
= g
11
(u, v)du
2
+ 2g
12
(u, v)dudv + g
22
(u, v)dv
2
The length of the curve : u = u(t), v = v(t), where t
0
t t
1
according to
(1.21) is equal to
L

t
1
t
0

v, v =

t
1
t
0

g
ik
(x) x
i
x
k
= (1.23)

t
1
t
0

g
11
(u(t) , v (t)) u
2
t
+ 2g
12
(u (t) , v (t)) u
t
v
t
+ g
22
(u(t) , v (t)) v
2
t
dt
(1.24)
The length of the curve dened by the formula(1.21) obeys the following
natural conditions
It coincides with the usual length in the Euclidean space E
n
(R
n
with
standard metric G = (dx
1
)
2
+ + (dx
n
)
2
in cartesian coordinates).
E.g. for 3-dimensional Euclidean space
L

b
a

g
ik
(x(t)) x
i
(t) x
k
(t)dt =

b
a

( x
1
(t))
2
+ ( x
2
(t))
2
+ ( x
3
(t))
2
dt
(1.25)
It does not depend on parameterisation of the curve
L

b
a

g
ik
(x(t)) x
i
(t) x
k
(t)dt =

g
ik
(x()) x
i
() x
k
()d,
(1.26)
where x
i
() = x
i
(t()), a

while a t b.
13
It does not depend on coordinates on Riemannian manifold M
L

b
a

g
ik
(x(t)) x
i
(t) x
k
(t)dt =

b
a

g
i

k
(x

(t)) x
i

(t) x
k

(t)dt
(1.27)
It is additive: if a curve =
1
+, i.e. : x
i
(t), a t b,
1
: x
i
(t), a
t c and
2
: x
i
(t), c t b where a point c belongs to the interval
(a, b) then
L

= L

1 + L

2, i.e.

b
a

g
ik
(x(t)) x
i
(t) x
k
(t)dt =

c
a

g
ik
(x(t)) x
i
(t) x
k
(t)dt +

b
c

g
ik
(x(t)) x
i
(t) x
k
(t)dt (1.28)
Conditions (1.25) and (1.28) evidently are obeyed.
Condition (1.26) follows from the fact that
x
i
() =
dx(t())
d
=
dx(t())
dt
dt
d
= x
i
(t)
dt
d
.
Bearing in mind the above formula we have

g
ik
(x()) x
i
() x
k
()d =

g
ik
(x(t())) x
i
(t) x
k
(t)
_
dt
d
2
_
d =

g
ik
(x(t())) x
i
(t) x
k
(t)

dt
d

d =

b
a

g
ik
(x(t)) x
i
(t) x
k
(t)dt = L

Condition (1.27) follows from the condition (1.76):

b
a

g
i

k
(x

(t)) x
i

(t) x
k

(t)dt =

b
a

g
ik
(x(t))
x
i
x
i

x
k
x
k

x
i

(t) x
k

(t)dt =

b
a

g
ik
(x(t))
_
x
i
x
i

x
i

(t)
__
x
k
x
k

x
k

(t)
_
dt =

b
a

g
ik
(x(t)) x
i
(t) x
k
(t)dt .
(1.29)
14
1.4 Riemannian structure on the surfaces embedded
in Euclidean space
Let M be a surface embedded in Euclidean space. Let G be Riemannian
structure on the manifold M.
Let X, Y be two vectors tangent to the surface M at a point p M. An
External Observer calculate this scalar product viewing these two vectors as
vectors in E
3
attached at the point p E
3
using scalar product in E
3
. An
Internal Observer will calculate the scalar product viewing these two vectors
as vectors tangent to the surface M using the Riemannian metric G (see the
formula (1.39)). Respectively
If L is a curve in M then an External Observer consider this curve as a
curve in E
3
, calculate the modulus of velocity vector (speed) and the length
of the curve using Euclidean scalar product of ambient space. An Internal
Observer (an ant) will dene the modulus of the velocity vector and the
length of the curve using Riemannian metric.
Denition Let M be a surface embedded in the Euclidean space. We
say that metric G
M
on the surface is induced by the Euclidean metric if the
scalar product of arbitrary two vectors A, B T
p
M calculated in terms of
the metric G equals to Euclidean scalar product of these two vectors:
A, B
G
M
= A, B
G
Euclidean
(1.30)
In other words we say that Riemannian metric on the embedded surface is
induced by the Euclidean structure of the ambient space if External and
Internal Observers come to the same results calculating scalar product of
vectors tangent to the surface.
In this case modulus of velocity vector (speed) and the length of the curve
is the same for External and Internal Observer.
Before going in details of this denition recall the conception of Internal
and External Observers when dealing with surfaces in Euclidean space:
1.4.1 Internal and external coordinates of tangent vector
Tangent plane
Let r = r(u, v) be parameterisation of the surface M embedded in the Eu-
clidean space:
r(u, v) =
_
_
x(u, v)
y(u, v)
z(u, v)
_
_
15
Here as always x, y, z are Cartesian coordinates in E
3
.
Let p be an arbitrary point on the surface M. Consider the plane formed
by the vectors which are adjusted to the point p and tangent to the surface
M. We call this plane plane tangent to M at the point p and denote it by
T
p
M.
For a point p M one can consider a basis in the tangent plane T
p
M
adjusted to the parameters u, v. Tangent basis vectors at any point (u, v)
are
r
u
=
r(u, v)
u
=
_
_
_
x(u,v)
u
y(u,v)
u
z(u,v)
u
_
_
_
=
x(u, v)
u

x
+
y(u, v)
u

y
+
z(u, v)
u

z
Every vector X T
p
M can be expanded over this basis:
X = X
u
r
u
+ X
v
r
v
, (1.31)
where X
u
, X
v
are coecients, components of the vector X.
Internal Observer views the basis vector r
u
T
p
M, as a velocity vector
for the curve u = u
0
+t, v = v
0
, where (u
0
, v
0
) are coordinates of the point p.
Respectively the basis vector r
v
T
p
M for an Internal Observer, is velocity
vector for the curve u = u
0
, v = v
0
+ t, where (u
0
, v
0
) are coordinates of the
point p.
Let r = r(t) be a curve belonging to the surface C, which passes through
the point p, r(t) = r(u(t), v(t)) and p = r(t
0
). Then vector
r
t
=
dr
dt
=
dr(u(t), v(t))
dt
(1.32)
belongs to the tangent plane T
p
M.
Note that for the vector (1.32) components X
u
, X
v
are equal to X
u
=
u
t
, X
v
= v
t
because
r
t
=
dr
dt
=
dr(u(t), v(t))
dt
= u
t
r
u
+ v
t
r
v
(1.33)
An External Observer describes the vector r
t
as a vector in E
3
attached at
the point p. The Internal Observer describes this vector as a vector which
has components (u
t
, v
t
) in the basis r
u
, r
v
according to the formula (1.33).
In general consider an arbitrary tangent vector X T
p
M. Denote X
u
=
a, X
v
= b
16
X = X
u
r
u
+X
v
r
v
= ar
u
+br
v
= a
_
_
x
u
(u, v)
y
u
(u, v)
z
u
(u, v)
_
_
+b
_
_
x
v
(u, v)
x
v
(u, v)
x
v
(u, v)
_
_
=
_
_
ax
u
(u, v) + bx
v
(u, v)
ay
u
(u, v) + by
v
(u, v)
az
u
(u, v) + bz
v
(u, v)
_
_
(1.34)
The last column in this formula represents the three components of the vector
X in the ambient space.
The pair (a, b) can be considered as internal coordinates of the tangent
vector X. An Internal Observer, Ant living on the surface, deals with the
vector Xin terms of coordinates (a, b). External observer which contemplates
the surface embedded in three-dimensional ambient space deals with vector
X as with vector with three external coordinates (see the last right column
in the formula (1.34).)
In condensed notation instead denoting coordinates by (u, v) we often
denote them by u

= (u
1
, u
2
). Respectively we denote by
r

=
dr
du

, r
u
= r
1
, r
v
= r
2
The formula (1.34) for tangent vector eld will have the following appear-
ance:
X = X

= X
1
r
1
+ X
2
r
2
, (X
1
= X
u
, X
2
= X
v
) (1.35)
When using condensed notations we usually omit explicit summation
symbols. E.g. we write u

instead

2
i=1
u

or u
1
r
1
+ u
2
r
2
One can consider also dierentials du

= (du
1
, du
2
):
du

(r

) =

: du
1
(r
1
) = du
2
(r
2
) = 1, du
1
(r
2
) = du
2
(r
1
) = 0 (1.36)
1.4.2 Explicit formulae for induced Riemannian metric (First Quadratic
form)
Now we are ready to write down the explicit formuale for the Riemannian
metric on the surface induced by metric (scalar product) in ambient Eu-
clidean space (see the Denition (1.30)).
Let M: r = r(u, v) be a surface embedded in E
3
.
The formula (1.30) means that scalar products of basic vectors r
u
=

u
, r
v
=
v
has to be the same calculated in the ambient space and on
17
the surface: For example scalar product
u
,
v

M
= g
uv
calculated by the
Internal Observer is the same as a scalar product r
u
, r
v

E
3 calculated by the
External Observer, scalar product
v
,
v

M
= g
uv
calculated by the Internal
Observer is the same as a scalar product r
v
, r
v

E
3 calculated by the External
Observer and so on:
G =
_
g
uu
g
uv
g
vu
g
vv
_
=
_

u
,
u

u
,
v

v
,
u

v
,
v

_
=
_
r
u
, r
u

E
3 r
u
, r
v

E
3
r
v
, r
u

E
3 r
v
, r
v

E
3
_
(1.37)
where as usual we denote by ,
E
3 the scalar product in the ambient Eu-
clidean space.
(Here see also the important remark (1.11))
Remark It is convenient sometimes to denote parameters (u, v) as (u
1
, u
2
)
or u

( = 1, 2) and to write r = r(u


1
, u
2
) or r = r(u

) ( = 1, 2) instead
r = r(u, v)
In these notations:
G
M
=
_
g
11
g
12
g
12
g
22
_
=
_
r
u
, r
u

E
3 r
u
, r
v

E
3
r
u
, r
v

E
3 r
v
, r
v

E
3
_
, g

= r

, r

,
G
M
= g

du

du

= g
11
du
2
+ 2g
12
dudv + g
22
dv
2
(1.38)
where ( , ) is a scalar product in Euclidean space.
The formula (1.38) is the formula for induced Riemannian metric on the
surfaceFirst Quadratic Form.
If X, Y are two tangent vectors in the tangent plane T
p
C then G(X, Y)
at the point p is equal to scalar product of vectors X, Y:
(X, Y) = (X
1
r
1
+ X
2
r
2
, Y
1
r
1
+ Y
2
r
2
) = (1.39)
X
1
(r
1
, r
1
)Y
1
+ X
1
(r
1
, r
2
)Y
2
+ X
2
(r
2
, r
1
)Y
1
+ X
2
(r
2
, r
2
)Y
2
=
X

(r

, r

)Y

= X

= G(X, Y)
We can come to this formula just transforming dierentials. In carteisan
coordinates X, Y = X
1
Y
1
+ X
2
Y
2
+ X
3
Y
3
, i.e. the Euclidean metric in
cartesian coordinates is given by
G
E
3 = (dx)
2
+ (dy)
2
+ (dz)
2
. (1.40)
The condition that Riemannian metric (1.38) is induced by Euclidean scalar
product means that
18
G
E
3

r=r(u,v)
=
_
(dx)
2
+ (dy)
2
+ (dz)
2
_

r=r(u,v)
= G
M
= g

du

du

(1.41)
i.e. ((dx)
2
+ (dy)
2
+ (dz)
2
)

r=r(u,v)
=
_
x(u, v)
u
du +
x(u, v)
u
du
_
2
+
_
x(u, v)
u
du +
x(u, v)
u
du
_
2
+
_
x(u, v)
u
du +
x(u, v)
u
du
_
2
=
(x
2
u
+ y
2
u
+ z
2
u
)du
2
+ 2(x
u
x
v
+ y
u
y
v
+ z
u
z
v
)dudv + (x
2
v
+ y
2
v
+ z
2
v
)dv
2
We see that
G
M
= g

du

du

= g
11
du
2
+ 2g
12
dudv + g
22
dv
2
, (1.42)
where g
11
= g
uu
= (x
2
u
+ y
2
u
+ z
2
u
) = r
u
, r
u

E
3, g
12
= g
21
= g
uv
= g
vu
=
(x
u
x
v
+y
u
y
v
+z
u
z
v
) = r
u
, r
v

E
3, g
22
= g
vv
= (x
2
v
+y
2
v
+z
2
v
) = r
v
, r
v

E
3. We
come to the same formula.
(See the examples of calculations in the next subsection.)
Check explicitly again that length of the tangent vectors and of the curves
calculating by External observer (i.e. using Euclidean metric (1.40)) is the
same as calculating by Internal Observer, ant (i.e. using the induced Rie-
mannian metric (1.38))
Consider a vector X = X

= ar
u
+ br
v
tangent to the surface M.
Calculate its length by External and Internal Observer.
The square of the length |X| of this vector calculated by External observer
(he calculates using the scalar product in E
3
) equals to
|X|
2
= X, X = r
u
+ br
v
, ar
u
+ br
v
= a
2
r
u
, r
u
+ 2abr
u
, r
v
+ b
2
r
v
, r
v

(1.43)
where , is a scalar product in E
3
.
The internal observer will calculate the length using Riemannian metric
(1.38):
G(X, X) =
_
a, b
_

_
g
11
g
12
g
21
G
22
_

_
a
b
_
= g
11
a
2
+ 2g
12
ab + g
22
b
2
(1.44)
External observer (person living in ambient space E
3
) calculate the length
of the tangent vector using formula (1.43). An ant living on the surface
calculate length of this vector in internal coordinates using formula (1.44).
19
External observer deals with external coordinates of the vector, ant on the
surface with internal coordinates. They come to the same answer.
Let r(t) = r(u(t), v(t)) a t b be a curve on the surface.
Velocity of this curve at the point r(u(t), v(t)) is equal to
v = X = r
u
+ r
v
where = u
t
, = v
t
: v =
dr(t)
dt
= u
t
r
u
+ v
t
r
v
.
The length of the curve is equal to
L =

b
a
|v(t)|dt =

b
a

v(t), v(t)
E
3dt =

b
a

u
t
r
u
+ v
t
r
v
, u
t
r
u
+ v
t
r
v

E
3dt =
(1.45)

b
a

r
u
, r
u

E
3u
2
t
+ 2r
u
, r
v

E
3u
t
v
t
+r
v
, r
v

E
3v
2
t
d =

b
a

g
11
u
2
t
+ 2g
12
u
t
v
t
+ g
22
v
2
t
dt (1.46)
An external observer will calculate the length of the curve using (1.45).
An ant living on the surface calculate length of the curve using (1.46) using
Riemannian metric on the surface:
ds
2
= g
ik
du
i
du
k
= g
11
du
2
+ 2g
12
dudv + g
22
dv
2
(1.47)
They will come to the same answer.
1.4.3 Induced Riemannian metrics. Examples.
We consider here examples of calculating induced Riemanian metric on some
quadratic surfaces in E
3
. using calculations for tangent vectors (see (1.38))
or explicitly in terms of dierentials (see (1.41) and (1.42)).
First of all consider the general case when a surface M is dened by the
equation z F(x, y) = 0. One can consider the following parameterisation
of this surface:
r(u, v):
_

_
x = u
y = v
z = F(u, v)
(1.48)
Then
20
r
u
=
_
_
1
0
F
u
_
_
r
v
=
_
_
0
1
F
v
_
_
(1.49)
,
(r
u
, r
u
) = 1 + F
2
u
, (r
u
, r
v
) = F
u
F
v
, (r
v
, r
v
) = 1 + F
2
v
and induced Riemannina metric (rst quadratic form) (1.38) is equal to
||g

|| =
_
g
11
g
12
g
12
g
22
_
=
_
(r
u
, r
u
) (r
u
, r
v
)
(r
u
, r
v
) (r
v
, r
v
)
_
=
_
1 + F
2
u
F
u
F
v
F
u
F
v
1 + F
2
v
_
(1.50)
G
M
= ds
2
= (1 + F
2
u
)du
2
+ 2F
u
F
v
dudv + (1 + F
2
v
)dv
2
(1.51)
and the length of the curve r(t) = r(u(t), v(t)) on C (a t b) can be
calculated by the formula:
L =

b
a

(1 + F
2
u
)u
2
t
+ 2F
u
F
v
u
t
v
t
+ (1 + F
v
)
2
v
2
t
dt
One can calculate (1.51) explicitly using (1.41):
G
M
=
_
dx
2
+ dy
2
+ dz
2
_

x=u,y=v,z=F(u,v)
= (du)
2
+(dv)
2
+(F
u
du +F
v
dv)
2
=
= (1 + F
2
u
)du
2
+ 2F
u
F
v
dudv + (1 + F
2
v
)dv
2
. (1.52)
Cylinder
Cylinder is given by the equation x
2
+ y
2
= a
2
. One can consider the
following parameterisation of this surface:
r(h, ):
_

_
x = a cos
y = a sin
z = h
(1.53)
We have G
cylinder
= (dx
2
+ dy
2
+ dz
2
)

x=a cos ,y=a sin ,z=h


=
= (a sin d)
2
+ (a cos d)
2
+ dh
2
= a
2
d
2
+ dh
2
(1.54)
The same formula in terms of scalar product of tangent vectors:
21
r
h
=
_
_
0
0
1
_
_
r

=
_
_
a sin
a cos
0
_
_
(1.55)
,
(r
h
, r
h
) = 1, (r
h
, r

) = 0, (r

, r

) = a
2
and
||g

|| =
_
(r
u
, r
u
) (r
u
, r
v
)
(r
u
, r
v
) (r
v
, r
v
)
_
=
_
1 0
0 a
2
_
,
G = dh
2
+ a
2
d
2
(1.56)
and the length of the curve r(t) = r(h(t), (t)) on the cylinder (a t b)
can be calculated by the formula:
L =

b
a

h
2
t
+ a
2

t
dt (1.57)
Cone
Cone is given by the equation x
2
+ y
2
k
2
z
2
= 0. One can consider the
following parameterisation of this surface:
r(h, ):
_

_
x = khcos
y = khsin
z = h
(1.58)
Calculate induced Riemannian metric:
We have
G
conus
=
_
dx
2
+ dy
2
+ dz
2
_

x=khcos ,y=khsin ,z=h


=
(k cos dh khsin d)
2
+ (k sin dh + khcos d)
2
+ dh
2
G
conus
= k
2
h
2
d
2
+ (1 + k
2
)dh
2
, ||g

|| =
_
1 + k
2
0
0 k
2
h
2
_
(1.59)
The length of the curve r(t) = r(h(t), (t)) on the cone (a t b) can be
calculated by the formula:
L =

b
a

(1 + k
2
)h
2
t
+ k
2
h
2

2
t
dt (1.60)
22
Sphere
Sphere is given by the equation x
2
+y
2
+z
2
= a
2
. Consider the following
(standard ) parameterisation of this surface:
r(, ):
_

_
x = a sin cos
y = a sin sin
z = a cos
(1.61)
Calculate induced Riemannian metric (rst quadratic form)
G
S
2 =
_
dx
2
+ dy
2
+ dz
2
_

x=a sin cos ,y=a sin sin ,z=a cos


=
(a cos cos da sin sin d)
2
+(a cos sin d+a sin cos d)
2
+(a sin d)
2
=
a
2
cos
2
d
2
+ a
2
sin
2
d
2
+ a
2
sin
2
d
2
=
, = a
2
d
2
+ a
2
sin
2
d
2
, ||g

|| =
_
a
2
0
0 a
2
sin
2

_
(1.62)
One comes to the same answer calculating scalar product of tangent vec-
tors:
r

=
_
_
a cos cos
a cos sin
a sin
_
_
r

=
_
_
a sin sin
a sin cos
0
_
_
(1.63)
,
(r

, r

) = a
2
, (r
h
, r

) = 0, (r

, r

) = a
2
sin
2

and
||g|| =
_
(r
u
, r
u
) (r
u
, r
v
)
(r
u
, r
v
) (r
v
, r
v
)
_
=
_
a
2
0
0 a
2
sin
2

_
, G
S
2 = ds
2
= a
2
d
2
+ a
2
sin
2
d
2
The length of the curve r(t) = r((t), (t)) on the sphere of the radius a
(a t b) can be calculated by the formula:
L =

b
a
a

2
t
+ sin
2

2
t
dt (1.64)
We considered all quadratic surfaces except paraboloid z = x
2
y
2
.
It can be rewritten as z = xy () and is call sometimes saddle
23
Saddle (paraboloid)
Saddle is given by the equation z xy = 0.
Why paraboloid?
Exercise Show that the equation of saddle can be rewritten as z = x
2
y
2
(This surface is a ruled surface containing lines...)
Consider the following (standard ) parameterisation of this surface:
r(u, v):
_

_
x = u
y = v
z = uv
(1.65)
Calculate induced metric:
G
saddle
=
_
dx
2
+ dy
2
+ dz
2
_

x=ucos ,y=v sin ,z=uv


=
du
2
+ dv
2
+ (udv + vdu)
2
=
G
saddle
= (1 + v
2
)du
2
+ 2uvdudv + (1 + u
2
)dv
2
. (1.66)
The length of the curve r(t) = r(u(t), v(t)) on the sphere of the radius a
(a t b) can be calculated by the formula:
L =

b
a
a

(1 + v
2
)u
2
t
+ 2uvu
t
v
t
+ (1 + u
2
)v
2
t
dt (1.67)
One-sheeted and two-sheeted hyperboloids.
Consider surface given by the equation
x
2
+ y
2
z
2
= c
If c = 0 it is a cone. We considered it already above.
If c > 0 it is one-sheeted hyperboloidconnected surface in E
3
If c < 0
it is two-sheeted hyperboloid a surface with two sheets z > 0 and z < 0 L
Consider these cases separately.
1) One-sheeted hyperboloid: x
2
+ y
2
z
2
= a
2
. It is ruled surface. It
contains two family of lines
Exercise

Find the lines on two-sheeted hyperboloid


24
One-sheeted hyperboloid is given by the equation x
2
+y
2
z
2
= a
2
. it is
convenient to choose parameterisation:
r(, ):
_

_
x = a cosh cos
y = a cosh sin
z = a sinh
(1.68)
x
2
+ y
2
z
2
= a
2
cosh
2
a
2
sinh
2
= a
2
Compare the calculations with calculations for sphere! We changed functions
cos, sin on cosh, sinh
Induced Riemannian metric (rst quadratic form).
G
HyperbolI
=
_
dx
2
+ dy
2
+ dz
2
_

x=a cosh cos ,y=a cosh sin ,z=a sinh


=
(a sinh cos da cosh sin d)
2
+(a sinh sin d+a cosh cos d)
2
+(a cosh d)
2
=
a
2
sinh
2
d
2
+ a
2
cosh
2
d
2
+ a
2
cosh
2
d
2
=
, = a
2
(1+2 sinh
2
)d
2
+a
2
cosh
2
d
2
, ||g

|| =
_
1 + 2 sinh
2
0
0 cosh
2

_
(1.69)
For two-sheeted hyperboloid calculatiosn will be very similar.
1) Two-sheeted hyperboloid: z
2
x
2
y
2
= a
2
. It is not ruled surface!
it is convenient to choose parameterisation:
r(, ):
_

_
x = a sinh cos
y = a sinh sin
z = a cosh
(1.70)
z
2
x
2
y
2
= a
2
cosh

a
2
sinh
2
= a
2
Compare the calculations with calculations for sphere! We changed functions
cos, sin on cosh, sinh
Induced Riemannian metric (rst quadratic form).
G
HyperbolI
=
_
dx
2
+ dy
2
+ dz
2
_

x=a sinh cos ,y=a sinh sin ,z=a cosh


=
25
(a cosh cos da sinh sin d)
2
+(a cosh sin d+a sinh cos d)
2
+(a sinh d)
2
=
a
2
cosh
2
d
2
+ a
2
sinh
2
d
2
+ a
2
sinh
2
d
2
=
, = a
2
(1+2 sinh
2
)d
2
+a
2
sinh
2
d
2
, ||g

|| =
_
1 + 2 sinh
2
0
0 sinh
2

_
(1.71)
We calculated examples of induced Riemannian structure embedded in
Euclidean space almost for all quadratic surfaces.
Quadratic surface is a surface dened by the equation
Ax
2
+ By
2
+ Cz
2
+ 2Dxy + 2Exz + 2Fyz + ex + fy + dz + c = 0
One can see that any quadratic surface by ane transformation can be trans-
formed to one of these surfaces
cylinder (elliptic cylinder) x
2
+ y
2
= 1
hyperbolic cylinder: x
2
y
2
= 1)
parabolic cylinder z = x
2
paraboloid x
2
+ y
2
= z
hyperbolic paraboloid x
2
y
2
= z
cone x
2
+ y
2
z
2
= 0
sphere x
2
+ y
2
+ z
2
= 1
one-sheeted hyperboloid
two-sheeted hyperboloid
(We exclude degenerate cases such as point x
2
+y
2
+z
2
= 0, planes, e.t.c.)
26
1.4.4

Induced metric on two-sheeted hyperboloid embedded in
pseudo-Euclidean space.
Consider two-sheeted hyperboloid embedded in pseudo-Euclidean space with pseudo-
scalar product dened by bilinear form
X, Y
pseudo
= X
1
Y
1
+X
2
Y
2
X
3
Y
3
(1.72)
The pseudoscalar product is bilinear, symmetric. It is dened by non-degenerate
matrix. But it is not positive-denite: The pseudo-length of vectors X =
(a cos , a sin , a) equals to zero:
X = (a cos , a sin , a) X, X
pseudo
= 0, (1.73)
This is not scalar product. The pseudo-Riemannian metric is:
G
pseudo
= dx
2
+dy
2
dz
2
(1.74)
it turns out that the following remarkable fact occurs:
Proposition The pseudo-Riemannian metric (1.74) in the ambient 3-dimensional
pseudo-Euclidean space induces Riemannian metric on two-sheeted hyperboloid
x
2
+y
2
z
2
= 1.
Show it. repeat the calculations above for two-sheeted hyperboloid changing in
the ambient space Riemannian metric G = dx
2
+dy
2
+dz
2
on pseudo-Riemannian
dx
2
+dy
2
dz
2
:
Using (1.70) and (1.74) we come now to
G =
_
dx
2
+dy
2
dz
2
_

x=a sinh cos ,y=a sinh sin ,z=a cosh


=
(a cosh cos da sinh sin d)
2
+(a cosh sin d+a sinh cos d)
2
(a sinh d)
2
=
a
2
cosh
2
d
2
+a
2
sinh
2
d
2
a
2
sinh
2
d
2
, G
L
= a
2
d
2
+a
2
sinh
2
d
2
, ||g

|| =
_
1 0
0 sinh
2

_
(1.75)
The two-sheeted hyperboloid equipped with this metric is called hyperbolic or
Lobachevsky plane.
Now express Riemannian metric in stereographic coordinates.
Calculations are very similar to the case of stereographic coordinates of 2-
sphere x
2
+ y
2
+ z
2
= 1. (See homework 1). Centre of projection (0, 0, 1): For
stereographic coordinates u, v we have
u
x
=
y
v
=
1
1+z
. We come to
_
u =
x
1+z
v =
y
1+z
,
_

_
x =
2u
1u
2
v
2
y =
2v
1u
2
v
2
z =
u
2
+v
2
+1
1u
2
v
2
(4)
27
The image of upper-sheet is an open disc u
2
+v
2
= 1 since u
2
+v
2
=
x
2
+y
2
(1+z)
2
=
z
2
1
(1+z)
2
=
z1
z+1
. Since for upper sheet z > 1 then 0
z1
z+1
< 1.
G = (dx
2
+dy
2
dz
2
)

x=x(u,v),y=y(u,v),z=z(u,v)
=
_
d
_
2u
1 u
2
v
2
__
2
+
_
d
_
2v
1 u
2
v
2
__
2

_
d
_
u
2
+v
2
+ 1
1 u
2
v
2
__
2
=
(Compare with calculations for sphere x
2
+y
2
+z
2
= 1). We have G =
_
2du
1 u
2
v
2
+
2u(2udu + 2vdv)
(1 u
2
v
2
)
2
_
2
+
_
2dv
1 u
2
v
2
+
2v(2udu + 2vdv)
(1 u
2
v
2
)
2
_
2

_
2udu + 2vdv
1 u
2
v
2
+
(u
2
+v
2
+ 1)(2udu + 2vdv)
(1 u
2
v
2
)
2
_
2
=
4(du)
2
+ 4(dv)
2
(1 +u
2
+v
2
)
2
(To perform these calculations it is convenient to denote by s = 1 u
2
v
2
).
1.5 Isometries of Riemanian manifolds.
Let (M
1
, G
(1)
), (M
2
, G
(2)
) be two Riemannian manifoldsmanifolds equipped
with Riemannian metric G
1
and G
2
respectively.
Denition
We say that these Riemannian manifolds are isometric if there exists a
dieomorphism F (one-one smooth map) which preserves the distances. This
means the following:
We say that these Riemannian manifolds are isometric if there exists a
dieomorphism F (one-one smooth map) such that
F

G
(2)
= G
(1)
,
which means the following:
Let p
1
be an arbitrary point on manifold M
1
and p
2
M
2
be its image:F(p
1
) =
p
2
. Let {x
i
} be coordinates in a vicinity of a point p
1
M
1
and {y
a
} be
coordinates in a vicinity of a point p
2
M
2
. Let Riemannian metrics G
1
on
M
1
has local expression G
(1)
= g
(1)ik
(x)dx
i
dx
k
in coordinates {x
i
} and re-
spectively Riemannian metrics G
(2)
has local expression G
2
= g
(2)ab
(y)dy
a
dy
b
in coordinates {y
i
} on M
2
. Then
g
(1)ik
(x)dx
i
dx
k
= g
(2)ab
(y)dy
a
dy
b
= g
(2)ab
(y(x))
y
a
(x)
x
i
dx
i
y
b
(x)
x
k
dx
k
(1.76)
28
i.e.
g
(1)ik
(x) =
y
a
(x)
x
i
g
(2)ab
(y(x))
y
b
(x)
x
k
. (1.77)
where y
i
= y
i
(x) is local expression for dieomorphism F.
Denition We say that two Riemannian manifolds (M
1
, G
(1)
), (M
2
, G
(2)
)
are locally isometric if the following conditions hold:
For arbitrary point p
1
on the manifold M
1
there exists a point p
2
on the
manifold M
2
such that there exist coordinates {x
i
} in a vicinity of a point
p
1
M
1
and coordinates {y
a
} in a vicinity of a point p
2
M
2
such that local
expression for metric G
(1)
on M
1
in coordinates {x
i
} and local expression for
metric G
(2)
on M
2
in coordinates {y
a
} are related via formuale (1.76), (1.29).
Locally isometric Riemannian manifolds have not to be dieomorphic.
E.g. Euclidean plane is locally isometric to cylinder, but they are not dieo-
morphic.
1.5.1 Examples of local isometries
Consider examples.
Example 1 Cylinder-ConePlane,
Riemannian metric on cylinder is G
cylinder
= a
2
d
2
+dh
2
and on the cone
G
conus
= k
2
h
2
d
2
+ (1 + k
2
)dh
2
(see formulae (1.54) and (1.59)).
To show that cylinder is isometric to Euclidean plan we have to nd
new local coordinates u, v on the cylinder such that in these coordinates the
metric on the cylinder equals to du
2
+ dv
2
. If we put u = a, v = h then
du
2
+ dv
2
= d(a)
2
+ dh
2
= a
2
d
2
+ dh
2
= G
cylinder
. (1.78)
Thus we prove the local isometry. Of course the coordinate u = is not
global coordinate on the surface of cylinder. It is evident that cylinder and
plane are not globally isometric since there are no dieomorphism of cylinder
on the plane. (They are dierent non-homeomorphic topological spaces.)
Now show that cone is locally isometric to the plane.
This means that we have to nd local coordinates u, v on the cone such
that in these coordinates induced metric G|
c
on cone would have the appear-
ance G|
c
= du
2
+ dv
2
.
29
First of all calculate the metric on cone in natural coordinates h, where
r(h, ):
_

_
x = khcos
y = khsin
z = h
.
(x
2
+ y
2
k
2
z
2
= k
2
h
2
cos
2
+ k
2
h
2
sin
2
k
2
h
2
= k
2
h
2
k
2
h
2
= 0.
Calculate metric G
c
on the cone in coordinates h, induced with the
Euclidean metric G = dx
2
+ dy
2
+ dz
2
:
G
c
=
_
dx
2
+ dy
2
+ dz
2
_

x=khcos ,y=khsin,z=h
= (k cos dh khsin d)
2
+
(k sin dh + khcos d)
2
+ dh
2
= (k
2
+ 1)dh
2
+ k
2
h
2
d
2
.
In analogy with polar coordinates try to nd new local coordinates u, v such
that
_
u = hcos
v = hsin
, where , are parameters. We come to du
2
+dv
2
=
(cos dh hsin d)
2
+(sin dh + hcos d)
2
=
2
dh
2
+
2

2
h
2
d
2
.
Comparing with the metric on the cone G
c
= (1 + k
2
)dh
2
+ k
2
h
2
d
2
we see
that if we put = k and =
k

1+k
2
then du
2
+dv
2
=
2
dh
2
+
2

2
h
2
d
2
=
(1 + k
2
)dh
2
+ k
2
h
2
d
2
.
Thus in new local coordinates
_
u =

k
2
+ 1hcos
k

k
2
+1

v =

k
2
+ 1hsin
k

k
2
+1

induced metric on the cone becomes G|


c
= du
2
+ dv
2
, i.e. cone locally is
isometric to the Euclidean plane
Of course these coordinates are local. Cone and plane are not homeo-
morphic, thus they are not globally isometric.
One can show also that
2) Plane with metric
4R
2
(dx
2
+dy
2
)
(1+x
2
+y
2
)
2
is isometric to the sphere with radius
R(a) = ...
3) Disc with metric
du
2
+dv
2
(1u
2
v
2
)
2
is isometric to half plane with metric
dx
2
+dy
2
4y
2
.
(see exercises in Homeworks and Coursework.)
30
1.6 Volume element in Riemannian manifold
The volume element in n-dimensional Riemannian manifold with metric G =
g
ik
dx
i
dx
k
is dened by the formula

det g
ik
dx
1
dx
2
. . . dx
n
(1.79)
If D is a domain in the n-dimensional Riemannian manifold with metric
G = g
ik
dx
i
then its volume is equal to to the integral of volume element over
this domain.
V (D) =

det g
ik
dx
1
dx
2
. . . dx
n
(1.80)
Remark Students who know the concept of exterior forms can read the
volume element as

det g
ik
dx
1
dx
2
dx
n
(1.81)
Note that in the case of n = 1 volume is just the length, in the case if
n = 2 it is area.
1.6.1 Volume of parallelepiped
Note that the formula (1.79) gives the volume of n-dimensional parallelepiped.
Show this. Let E
n
be Euclidean vector space with orthonormal basis {e
i
}.
Let v
i
be an arbitrary basis in this vector space (vectors v
i
in general have not
unit length and are not orthogonal to each other). Consider n-parallelepiped
spanned by vectors {v
i
}:

v
i
: r = t
i
v
i
, 0 t
i
1.
The volume of this parallelepiped equals to
V ol(
v
i
) = det ||a
m
i
|| , (1.82)
where A = ||a
m
i
|| is transition matrix, v
i
= e
m
a
m
i
. On the other hand
r = x
i
e
i
= t
m
v
m
, hence x
i
= a
i
m
t
m
, where v
m
= e
i
a
i
m
.
Let G = (dx
1
)
2
+ + (dx
n
)
2
= g
ik
dt
i
dt
k
be usual Euclidean metric in new
coordinates t
i
. The
G = (dx
1
)
2
+ + (dx
n
)
2
= dx
i

ik
dx
k
= dt
i
x
i

t
i

i

x
k

t
k
dt
k
.
31
Since
x
i

t
i
= a
i

i
then
g
ik
=

a
i

i
a
i

k
det g = (det A)
2
, det g =

det A.
and according to the formula (1.80)
V ol(
v
i
) =

0t
i
1

det gdt
1
dt
2
. . . dt
n
= det A.
We come to (1.82).
Perform these calculations in detail for 3-dimensional case.
E.g. if Euclidean space is 3-dimensional then the parallelepiped spanned
by basis vectors {a, b, c}

a,b,c
= t
1
a + t
2
b + t
3
c , 0 t
1
, t
2
, t
3
1.
Volume of parallelepiped equals to
V ol
a,b,c
= det
_
_
a
x
b
x
c
x
a
y
b
y
c
y
a
z
b
z
c
z
_
_
, (1.83)
and
_
_
x
y
z
_
_
=
_
_
a
x
b
x
c
x
a
y
b
y
c
y
a
z
b
z
c
z
_
_
_
_
t
1
t
2
t
3
_
_
.
The Riemannian metric in new coordinates (t
1
, t
2
, t
3
) equals to
dx
2
+ dy
2
+ dz
2
=
_
dt
1
dt
2
dt
3
_
_
_
a
x
a
y
a
z
b
x
b
y
b
z
c
x
c
y
c
z
_
_
_
_
a
x
b
x
c
x
a
y
b
y
c
y
a
z
b
z
c
z
_
_
_
_
dt
1
dt
2
dt
3
_
_
,
i.e. in coordinates t
i
Riemannian metric G = g
ik
dt
i
dt
k
where
_
_
g
11
g
12
g
13
g
21
g
22
g
23
g
31
g
32
g
33
_
_
=
_
_
a
x
a
y
a
z
b
x
b
y
b
z
c
x
c
y
c
z
_
_
_
_
a
x
b
x
c
x
a
y
b
y
c
y
a
z
b
z
c
z
_
_
i.e.

det g
ik
= det
_
_
a
x
b
x
c
x
a
y
b
y
c
y
a
z
b
z
c
z
_
_
= V ol
a,b,c
. (1.84)
32
1.6.2 Invariance of volume element under changing of coordinates
Prove that volume element is invariant under coordinate transformations, i.e. if
y
1
, . . . , y
n
are new coordinates: x
1
= x
1
(y
1
, . . . , y
n
), x
2
= x
2
(y
1
, . . . , y
n
)...,
x
i
= x
i
(y
p
), i = 1, . . . , n , p = 1, . . . , n
and g
pq
(y) matrix of the metric in new coordinates:
g
pq
(y) =
x
i
y
p
g
ik
(x(y))
x
k
y
q
. (1.85)
Then

det g
ik
(x) dx
1
dx
2
. . . dx
n
=

det g
pq
(y) dy
1
dy
2
. . . dy
n
(1.86)
This follows from (1.85). Namely

det g
ik
(y) dy
1
dy
2
. . . dy
n
=

det
_
x
i
y
p
g
ik
(x(y))
x
k
y
q
_
dy
1
dy
2
. . . dy
n
Using the fact that det(ABC) = det A det B det C and det
_
x
i
y
p
_
= det
_
x
k
y
q
_
1
we see that from the formula above follows:

det g
ik
(y) dy
1
dy
2
. . . dy
n
=

det
_
x
i
y
p
g
ik
(x(y))
x
k
y
q
_
dy
1
dy
2
. . . dy
n
=

_
det
_
x
i
y
p
__
2

det g
ik
(x(y))dy
1
dy
2
. . . dy
n
=

det g
ik
(x(y)) det
_
x
i
y
p
_
dy
1
dy
2
. . . dy
n
= (1.87)
Now note that
det
_
x
i
y
p
_
dy
1
dy
2
. . . dy
n
= dx
1
. . . dx
n
according to the formula for changing coordinates in n-dimensional integral
2
.
Hence

det g
ik
(x(y)) det
_
x
i
y
p
_
dy
1
dy
2
. . . dy
n
=

det g
ik
(x(y))dx
1
dx
2
. . . dx
n
(1.88)
1
determinant of matrix does not change if we change the matrix on the adjoint, i.e.
change columns on rows.
2
Determinant of the matrix
_
x
i
y
p
_
of changing of coordinates is called sometimes Ja-
cobian. Here we consider the case if Jacobian is positive. If Jacobian is negative then
formulae above remain valid just the symbol of modulus appears.
33
Thus we come to (1.86).
1.6.3 Examples of calculating volume element
Consider rst very simple example: Volume element of plane in cartesian
coordinates, metric g = dx
2
+ dy
2
. Volume element is equal to

det gdxdy =

det
_
1 0
0 1
_
dxdy = dxdy
Volume of the domain D is equal to
V (D) =

det gdxdy =

D
dxdy
If we go to polar coordinates:
x = r cos , y = r sin (1.89)
Then we have for metric:
G = dr
2
+ r
2
d
2
because
dx
2
+ dy
2
= (dr cos r sin d)
2
+ (dr sin + r cos d)
2
= dr
2
+ r
2
d
2
(1.90)
Volume element in polar coordinates is equal to

det gdrd =

det
_
1 0
0 r
2
_
drd = drd.
Lobachesvky plane.
In coordinates x, y (y > 0) metric G =
dx
2
+dy
2
y
2
, the corresponding matrix
G =
_
1/y
2
0
0
1/y
2
_
. Volume element is equal to

det gdxdy =
dxdy
y
2
.
Sphere in stereographic coordinates Consider the two dimensional plane
with Riemannian metrics
G =
4R
2
(du
2
+ dv
2
)
(1 + u
2
+ v
2
)
2
(1.91)
34
(It is isometric to the sphere of the radius R without North pole in stere-
ographic coordinates (see the Homeworks.))
Calculate its volume element and volume. It is easy to see that:
G =
_
4R
2
(1+u
2
+v
2
)
2
0
0
4R
2
(1+u
2
+v
2
)
2
_
det g =
16R
4
(1 + u
2
+ v
2
)
4
(1.92)
and volume element is equal to

det gdudv =
4R
2
dudv
(1+u
2
+v
2
)
2
One can calculate volume in coordinates u, v but it is better to consider
volume form in polar coordinates u = r cos , v = r sin . Then it is easy
to see that according to (1.90) we have for the metric G =
R
2
(du
2
+dv
2
)
(1+u
2
+v
2
)
2
=
R
2
(dr
2
+r
2
d
2
)
(1+r
2
)
2
and volume form is equal to

det gdrd =
4R
2
rdrd
(1+r
2
)
2
Now calculation of integral becomes easy:
V =

4R
2
rdrd
(1 + r
2
)
2
= 8R
2


0
rdr
(1 + r
2
)
2
= 4R
2


0
du
(1 + u)
2
= 4R
2
.
Segment of the sphere.
Consider sphere of the radius a in Euclidean space with standard Riema-
nian metric
a
2
d
2
+ a
2
sin
2
d
2
This metric is nothing but rst quadratic form on the sphere (see (1.4.3)).
The volume element is

det gdd =

det
_
a
2
0
0 a
2
sin
_
dd = a
2
sin dd
Now calculate the volume of the segment of the sphere between two parallel
planes, i.e. domain restricted by parallels
1

0
: Denote by h be the
height of this segment. One can see that
h = a cos
0
a cos
1
= a(cos
0
a cos
1
)
There is remarkable formula which express the area of segment via the height
h:
V =

0
_
a
2
sin
_
dd =

0
_
2
0
_
a
2
sin
_
d
_
d =
35

1
2a
2
sin d = 2a
2
(cos
0
cos
1
) = 2a(a cos
0
acos
1
) = 2ah
(1.93)
E.g. for all the sphere h = 2a. We come to S = 4a
2
. It is remarkable
formula: area of the segment is a polynomial function of radius of the sphere
and height (Compare with formula for length of the arc of the circle)
2 Covariant dierentiaion. Connection. Levi
Civita Connection on Riemannian mani-
fold
2.1 Dierentiation of vector eld along the vector eld.
Ane connection
How to dierentiate vector elds on a (smooth )manifold M?
Recall the dierentiation of functions on a (smooth )manifold M.
Let X = X
i
(x)e
i
(x) =

x
i
be a vector eld on M. Recall that vector
eld
3
X = X
i
e
i
denes at the every point x
0
an innitesimal curve: x
i
(t) =
x
i
0
+ tX
i
(More exactly the equivalence class [(t)]
X
of curves x
i
(t) = x
i
0
+
tX
i
+ . . . ).
Let f be an arbitrary (smooth) function on M and X = X
i
x
i
. Then
derivative of function f along vector eld X = X
i
x
i
is equal to

X
f =
X
f = X
i
f
x
i
The geometrical meaning of this denition is following: If X is a velocity
vector of the curve x
i
(t) at the point x
i
0
= x
i
(t) at the time t = 0 then
the value of the derivative
X
f at the point x
i
0
= x
i
(0) is equal just to the
derivative by t of the function f(x
i
(t)) at the time t = 0:
if X
i
(x)

x
0
=x(0)
=
dx
i
(t)
dt

t=0
, then
X
f

x
i
=x
i
(0)
=
d
dt
f
_
x
i
(t)
_

t=0
(2.1)
Remark In the course of Geometry and Dierentiable Manifolds the
operator of taking derivation of function along the vector eld was denoted
3
here like always we suppose by default the summation over repeated indices. E.g.X =
X
i
e
i
is nothing but X =

n
i=1
X
i
e
i
36
by
X
f. In this course we prefer to denote it by
X
f to have the uniform
notation for both operators of taking derivation of functions and vector elds
along the vector eld.
One can see that the operation
X
on the space C

(M) (space of smooth


functions on the manifold) satises the following conditions:

X
(f + g) = a
X
f +b
X
g where , R (linearity over numbers
)

hX+gY
(f) = h
X
(f) +g
Y
(f) (linearity over the space of functions)

X
(fg) = f
X
(g) + g
X
(f) (Leibnitz rule)
(2.2)
Remark One can prove that these properties characterize vector elds:operator
on smooth functions obeying the conditions above is a vector eld. (You will
have a detailed analysis of this statement in the course of Dierentiable Man-
ifolds.)
How to dene dierentiation of vector elds along vector elds.
The formula (2.1) cannot be generalised straightforwardly because vec-
tors at the point x
0
and x
0
+ tX are vectors from dierent vector spaces.
(We cannot substract the vector from one vector space from the vector from
the another vector space, because apriori we cannot compare vectors from
dierent vector space. One have to dene an operation of transport of vec-
tors from the space T
x
0
M to the point T
x
0
+tX
M dening the transport from
the point T
x
0
M to the point T
x
0
+tX
M).
Try to dene the operation on vector elds such that conditions (2.2)
above be satised.
2.1.1 Denition of connection. Christoel symbols of connection
Denition Ane connection on M is the operation which assigns to every
vector eld X a linear map, (but not necessarily C(M)-linear map!) (i.e. a
map which is linear over numbers not necessarily over functions)
X
on the
space O(M) of vector elds:

X
(Y + Z) =
X
Y +
X
Z, for every , R (2.3)
37
(Compare the rst condition in (2.2)).
which satises the following conditions:
for arbitrary (smooth) functions f, g on M

fX+gY
(Z) = f
X
(Z) + g
Y
(Z) (C(M)-linearity) (2.4)
(compare with second condition in (2.2))
for arbitrary function f

X
(fY) = (
X
f) Y + f
X
(Y) (Leibnitz rule) (2.5)
Recall that
X
f is just usual derivative of a function f along vector
eld:
X
f =
X
f.
(Compare with Leibnitz rule in (2.2)).
The operation
X
Y is called covariant derivative of vector eld Y along
the vector eld X.
Write down explicit formulae in a given local coordinates {x
i
} (i =
1, 2, . . . , n) on manifold M.
Let
X = X
i
e
i
= X
i

x
i
Y = Y
i
e
i
= Y
i

x
i
The basis vector elds

x
i
we denote sometimes by
i
sometimes by e
i
Using properties above one can see that

X
Y =
X
i

i
Y
k

k
= X
i
_

i
_
Y
k

k
__
, where
i
=

i
(2.6)
Then according to (2.4)

i
_
Y
k

k
_
=
i
_
Y
k
_

k
+ Y
k

k
Decompose the vector eld
i

k
over the basis
i
:

k
=
m
ik

m
(2.7)
and

i
_
Y
k

k
_
=
Y
k
(x)
x
i

k
+ Y
k

m
ik

m
, (2.8)
38

X
Y = X
i
Y
m
(x)
x
i

m
+ X
i
Y
k

m
ik

m
, (2.9)
In components
(
X
Y)
m
= X
i
_
Y
m
(x)
x
i
+ Y
k

m
ik
_
(2.10)
Coecients {
m
ik
} are called Christoel symbols in coordinates {x
i
}. These
coecients dene covariant derivativeconnection.
If operation of taking covariant derivative is given we say that the con-
nection is given on the manifold. Later it will be explained why we us the
word connection
We see from the formula above that to dene covariant derivative of vector
elds, connection, we have to dene Christoel symbols in local coordinates.
2.1.2 Transformation of Christoel symbols for an arbitrary con-
nection
Let be a connection on manifold M. Let {
i
km
} be Christoel symbols
of this connection in given local coordinates {x
i
}. Then according (2.7) and
(2.8) we have

X
Y = X
m
Y
i
x
m

x
i
+ X
m

i
mk
Y
k

x
i
,
and in particularly

i
mk

i
=

k
Use this relation to calculate Christoel symbols in new coordinates x
i

k

i
=

We have that
m
=

x
m

=
x
m
x
m

x
m
=
x
m
x
m

m
. Hence due to properties
(2.4), (2.5) we have

k

i
=

k
=

m
_
x
k
x
k

k
_
=
_
x
k
x
k

k
+

x
m

_
x
k
x
k

k
=
_
x
k
x
k

_
x
m
x
m

k
+

2
x
k
x
m

x
k

k
=
x
k
x
k

x
m
x
m

k
+

2
x
k
x
m

x
k

k
x
k
x
k

x
m
x
m

i
mk

i
+

2
x
k
x
m

x
k

k
=
x
k
x
k

x
m
x
m

i
mk
x
i

x
i

i
+

2
x
k
x
m

x
k

x
i

x
k

39
Comparing the rst and the last term in this formula we come to the trans-
formation law:
If {
i
km
} are Christoel symbols of the connection in local coordinates
{x
i
} and {
i

m
} are Christoel symbols of this connection in new local
coordinates {x
i

} then

m
=
x
k
x
k

x
m
x
m

x
i

x
i

i
km
+

2
x
r
x
k

x
m

x
i

x
r
(2.11)
Remark Christoel symbols do not transform as tensor. If the second
term is equal to zero, i.e. transformation of coordinates are linear (see the
Proposition on at connections) then the transformation rule above is the the
same as a transformation rule for tensors of the type
_
1
2
_
(see the formula
(1.6)). In general case this is not true. Christoel symbols does not trans-
form as tensor under arbitrary non-linear coordinate transformation: see the
second term in the formula above.
2.1.3 Canonical at ane connection
It follows from the properties of connection that it is suce to dene con-
nection at vector elds which form basis at the every point using (2.7), i.e.
to dene Christoel symbols of this connection.
Example Consider n-dimensional Euclidean space E
n
with cartesian co-
ordinates {x
1
, . . . , x
n
}.
Dene connection such that all Christoel symbols are equal to zero in
these cartesian coordinates {x
i
}.

e
i
e
k
=
m
ik
e
m
= 0,
m
ik
= 0 (2.12)
Does this mean that Christoel symbols are equal to zero in an arbitrary
cartesian coordinates if they equal to zero in given cartesian coordinates?
Does this mean that Christoel symbols of this connection equal to zero
in arbitrary coordinates system?
To answer these questions note that the relations (2.12) mean that

X
Y = X
m
Y
i
x
m

x
i
(2.13)
in coordinates {x
i
}
40
Consider an arbitrary new coordinates x
i

= x
i

(x
1
, . . . , x
n
). Recall the
transformation rule for an arbitrary vector eld (see subsection 1.1)
R = R
m

x
m
= R
m
x
m

x
m

x
m

, i.e.R
m

=
x
m

x
m
R
m
, and , R
m
=
x
m
x
m

R
m

.
Hence we have from (2.13) that

X
Y = X
m
Y
i
x
m

x
i
= X
m

x
m
_
Y
i
_

x
i
= X
m
x
m

x
m

x
m

_
x
i
x
i

Y
i

_

x
i
=
X
m

x
m

_
x
i
x
i

Y
i

_

x
i
= X
m

x
m

_
Y
i

_
x
i
x
i

x
i
+X
m

2
x
i
x
m

x
i

_
Y
i

_

x
i
=
X
m
Y
i

x
m

x
i

+ X
m

2
x
i
x
m

x
i

Y
i

x
i
. .
an additional term
= (2.14)
We see that an additional term equals to zero for arbitrary vector elds X, Y
if and only if the relations between new and old coordinates are linear:

2
x
i
x
m

x
i

= 0, i.e. x
i
= b
i
+ a
i
k
x
k
(2.15)
Comparing formulae (2.15) and (2.13) we come to simple but very important
Proposition Let all Christoel symbols of a given connection be equal to
zero in a given coordinate system {x
i
}. Then all Christoel symbols of this
connection are equal to zero in an arbitrary coordinate system {x
i

} such that
the relations between new and old coordinates are linear:
x
i

= b
i
+ a
i
k
x
k
(2.16)
If transformation to new coordinate system is not linear, i.e.

2
x
i
x
m

x
i

= 0
then Christoel symbols of this connection in general are not equal to zero in
new coordinate system {x
i

}.
Denition We call connection at if there exists coordinate system
such that all Christoel symbols of this connection are equal to zero in a
given coordinate system.
In particular connection (2.12) has zero Christoel symbols in arbitrary
cartesian coordinates.
41
Corollary Connection has zero Christoel symbols in arbitrary Cartesian
coordinates if it has zero Christoel symbols in a given Cartesian coordinates.
Hence the following denition is correct:
Denition A connection on E
n
which Christoel symbols vanish in carte-
sian coordinates is called canonical at connection.
Remark Canonical at connection in Euclidean space is uniquely dened,
sincce cartesian coordinates are dened globally. On the other hand on arbitrary
manifold one can dene at connection locally just choosing any arbitrary local
coordinates and dene locally at connection by condition that Christoel symbols
vanish in these local coordinates. This does not mean that one can dene at
connection globally. We will study this question after learning transformation law
for Christoel symbols.
Remark One can see that at connection is symmetric connection.
Example Consider a connection (2.12) in E
2
. It is a at connection.
Calculate Christoel symbols of this connection in polar coordinates
_
x = r cos
y = y sin
_
r =

x
2
+ y
2
= arctan
y
x
(2.17)
Write down Jacobians of transformationsmatrices of partial derivatives:
_
x
r
y
r
x

_
=
_
cos sin
r sin r cos
_
,
_
r
x

x
r
y

y
_
=
_
_
x

x
2
+y
2

y
x
2
+y
2
y

x
2
+y
2
x
x
2
+y
2
_
_
(2.18)
According (2.11) and since Chrsitoel symbols are equal to zero in cartesian
coordinates (x, y) we have

m
=

2
x
r
x
k

x
m

x
i

x
r
, (2.19)
where (x
1
, x
2
) = (x, y) and (x
1

, x
2

) = (r, ). Now using (2.18) we have

r
rr
=

2
x
rr
r
x
+

2
y
rr
r
y
= 0

r
r
=
r
r
=

2
x
r
r
x
+

2
y
r
r
y
= sin cos + sin cos = 0 .
42

=

2
x

r
x
+

2
y
r
r
y
= x
x
r
y
y
r
= r .

rr
=

2
x
rr

x
+

2
y
rr

y
= 0 .

r
=

r
=

2
x
r

x
+

2
y
r

y
= sin
y
r
2
+ cos
x
r
2
=
1
r

=

2
x

x
+

2
y

y
= x
x
r
2
y
y
r
2
= 0 . (2.20)
Hence we have that the covariant derivative (2.13) in polar coordinates has
the following appearance

r
=
r
rr

r
+

rr

= 0 , ,
r

=
r
r

r
+

r
=
r
r

r
+

r
,

=
r

r
+

= r
r
(2.21)
Remark Later when we study geodesics we will learn a very quick method
to calculate Christoel symbols.
2.1.4

Global aspects of existence of connection
We dened connection as an operation on vector elds obeying the special axioms
(see the subsubsection 2.1.1). Then we showed that in a given coordinates con-
nection is dened by Christoel symbols. On the other hand we know that in
general coordinates on manifold are not dened globally. (We had not this trouble
in Euclidean space where there are globally dened cartesian coordinates.)
How to dene connection globally using local coordinates?
Does there exist at least one globally dened connection?
Does there exist globally dened at connection?
These questions are not naive questions. Answer on rst and second questions
is Yes. It sounds bizzare but answer on the rst question is not Yes
4
Global denition of connection
4
Topology of the manifold can be an obstruction to existence of global at connection.
E.g. it does not exist on sphere S
n
if n > 1.
43
The formula (2.11) denes the transformation for Christoer symbols if we go
from one coordinates to another.
Let {(x
i

), U

} be an atlas of charts on the manifold M.


If connection is dened on the manifold M then it denes in any chart (local
coordinates) (x
i

) Christoer symbols which we denote by


()

i
km
. If (x
i

), (x
i

()
)
are dierent local coordinates in a vicinity of a given point then according to (2.11)
()

m
=
x
k
()
x
k

()
x
m
()
x
m

()
x
i

()
x
i
()
()

()i
mk
+

2
x
k
()
x
m

()
x
k

()
x
i

()
x
k
()
(2.22)
Denition Let {(x
i

), U

} be an atlas of charts on the manifold M


We say that the collection of Christoel symbols {
()i
km
} denes globally a
connection on the manifold M in this atlas if for every two local coordinates
(x
i
()
), (x
i
()
) from this atlas the transformation rules (2.22) are obeyed.
Using partition of unity one can prove the existence of global connection con-
structing it in explicit way. Let {(x
i

), U

} ( = 1, 2, . . . , N) be a nite atlas on
the manifold M and let {

} be a partition of unity adjusted to this atlas. Denote


by
()

i
km
local connection dened in domain U

such that its components in these


coordinates are equal to zero. Denote by
()
()

i
km
Christoel symbols of this local
connection in coordinates (x
i
()
) (
()
()

i
km
= 0). Now one can dene globally the
connection by the formula:
()

i
km
(x) =

(x)
()
()

i
km
(x) =

(x)
x
i
()
x
i

()

2
x
i

()
(x)
x
k
()
x
m
()
. (2.23)
This connection in general is not at connection
5
2.2 Connection induced on the surfaces
Let M be a manifold (surface) embedded in Euclidean space
6
. Canonical at
connection on E
N
induces the connection on surface in the following way.
Let X, Y be tangent vector elds to the surface M and
can.at
a canonical
at connection in E
N
. In general
Z =
can.at
X
Y is not tangent to manifold M (2.24)
5
See for detail the text: Global ane connection on manifold in my homepage:
www.maths.mancheser.ac.uk/khudian in subdirectory Etudes/Geometry
6
We know that every n-dimensional manifodl can be embedded in 2n + 1-dimensional
Euclidean space
44
Consider its decomposition on two vector elds:
Z = Z
tangent
+Z

,
can.at
X
, Y =
_

can.at
X
Y
_
tangent
+
_

can.at
X
Y
_

, (2.25)
where Z

is a component of vector which is orthogonal to the surface M


and Z
||
is a component which is tangent to the surface. Dene an induced
connection
M
on the surface M by the following formula

M
:
M
X
Y: =
_

can.at
X
Y
_
tangent
(2.26)
Remark One can imply this construction for an arbitrary connection in
E
N
.
2.2.1 Calculation of induced connection on surfaces in E
3
.
Let r = r(u, v) be a surface in E
3
. Let
can.at
be a at connection in E
3
.
Then

M
:
M
X
Y: =
_

can.at
X
Y
_
||
=
can.at
X
Yn(
can.at
X
Y, n), (2.27)
where n is normal unit vector eld to M. Consider a special example
Example (Induced connection on sphere) Consider a sphere of the radius
R in E
3
:
r(, ):
_

_
x = Rsin cos
y = Rsin sin
z = Rcos
then
r

=
_
_
Rcos cos
Rcos sin
Rsin
_
_
, r

=
_
_
Rsin sin
Rsin cos
0
_
_
, n =
_
_
sin cos
sin sin
cos
_
_
,
where r

=
r(,)

, r

=
r(,)

are basic tangent vectors and n is normal unit


vector.
Calculate an induced connection on the sphere.
First calculate

=
_
r

_
tangent
= (r

)
tangent
.
45
On the other hand one can see that r

=
_
_
Rsin cos
Rsin sin
Rcos
_
_
= Rn is
proportional to normal vector, i.e. (r

)
tangent
= 0. We come to

= (r

)
tangent
= 0

= 0 . (2.28)
Now calculate

and

=
_
r

_
tangent
= (r

)
tangent
,

=
_
r

_
tangent
= (r

)
tangent
We have

= (r

)
tangent
=
_
_
Rcos sin
Rcos cos
0
_
_
tangent
.
We see that the vector r

is orthogonal to n:
r

, n = Rcos sin sin cos + Rcos cos sin sin = 0.


Hence

= (r

)
tangent
= r

=
_
_
Rcos sin
Rcos cos
0
_
_
= cotan r

.
We come to

= cotan

= 0,

= cotan (2.29)
Finally calculate

= (r

)
tangent
=
_
_
_
_
Rsin cos
Rsin sin
0
_
_
_
_
tangent
Projecting on the tangent vectors to the sphere (see (2.27)) we have

= (r

)
tangent
= r

nn, r

=
46
_
_
Rsin cos
Rsin sin
0
_
_

_
_
sin cos
sin sin
cos
_
_
(Rsin cos sin cos Rsin sin sin sin ) =
sin cos
_
_
Rcos cos
Rcos sin
Rsin
_
_
= sin cos r

,
i.e.

= sin cos r

= sin cos ,

= 0 . (2.30)
2.3 Levi-Civita connection
2.3.1 Symmetric connection
Denition. We say that connection is symmetric if its Christoel symbols

i
km
are symmetric with respect to lower indices

i
km
=
i
mk
(2.31)
The canonical at connection and induced connections considered above are
symmetric connections.
Invariant denition of symmetric connection
A connection is symmetric if for an arbitrary vector elds X, Y

X
Y
Y
X[X, Y] = 0 (2.32)
If we apply this denition to basic elds
k
,
m
which commute: [
k
,
m
] = 0 we
come to the condition

k
=
i
mk

i
km

i
= 0
and this is the condition (2.31).
2.3.2 Levi-Civita connection. Theorem and Explicit formulae
Let (M, G) be a Riemannian manifold.
Denition. Theorem
A symmetric connection is called Levi-Civita connection if it is com-
patible with metric, i.e. if it preserves the scalar product:

X
Y, Z =
X
Y, Z +Y,
X
Z (2.33)
47
for arbitrary vector elds X, Y, Z.
There exists unique levi-Civita connection on the Riemannian manifold.
In local coordinates Christoel symbols of Levi-Civita connection are given
by the following formulae:

i
mk
=
1
2
g
ij
_
g
jm
x
k
+
g
jk
x
m

g
mk
x
j
_
. (2.34)
where G = g
ik
dx
i
dx
k
is Riemannian metric in local coordinates and ||g
ik
|| is
the matrix inverse to the matrix ||g
ik
||.
Proof
Suppose that this connection exists and
i
mk
are its Christoel symbols. Con-
sider vector elds X =
m
, Y =
i
and Z =
k
in (2.33). We have that

m
g
ik
=
r
mi

r
,
k
+
i
,
r
mk

r
=
r
mi
g
rk
+g
ir

r
mk
. (2.35)
for arbitrary indices m, i, k.
Denote by
mik
=
r
mi
g
rk
we come to

m
g
ik
=
mik
+
mki
, i.e.
Now using the symmetricity
mik
=
imk
since
k
mi
=
k
im
we have

mik
=
m
g
ik

mki
=
m
g
ik

kmi
=
m
g
ik
(
k
g
mi

kim
) =

m
g
ik

k
g
mi
+
kim
=
m
g
ik

k
g
mi
+
ikm
=
m
g
ik

k
g
mi
+(
i
g
km

imk
) =

m
g
ik

k
g
mi
+
i
g
km

mik
.
Hence

mik
=
1
2
(
m
g
ik
+
i
g
mk

k
g
mi
)
k
im
=
1
2
g
kr
(
m
g
ir
+
i
g
mr

r
g
mi
) (2.36)
We see that if this connection exists then it is given by the formula(2.34).
On the other hand one can see that (2.34) obeys the condition (2.35). We
prove the uniqueness and existence.
since

k
=
m
ik

m
.
Consider examples.
48
2.3.3 Levi-Civita connection on 2-dimensional Riemannian mani-
fold with metric G = adu
2
+ bdv
2
.
Example Consider 2-dimensional manifold with Riemannian metrics
G = a(u, v)du
2
+ b(u, v)dv
2
, G =
_
g
11
g
12
g
21
g
22
_
=
_
a(u, v) 0
0 b(u, v)
_
Calculate Christoel symbols of Levi Civita connection.
Using (2.36) we see that:

111
=
1
2
(
1
g
11
+
1
g
11

1
g
11
) =
1
2

1
g
11
=
1
2
a
u

211
=
121
=
1
2
(
1
g
12
+
2
g
11

1
g
12
) =
1
2

2
g
11
=
1
2
a
v

221
=
1
2
(
2
g
12
+
2
g
12

1
g
22
) =
1
2

1
g
22
=
1
2
b
u

112
=
1
2
(
1
g
12
+
1
g
12

2
g
11
) =
1
2

2
g
11
=
1
2
a
v

122
=
212
=
1
2
(
2
g
21
+
1
g
22

2
g
21
) =
1
2

1
g
22
=
1
2
b
u

222
=
1
2
(
2
g
22
+
2
g
22

2
g
22
) =
1
2

2
g
22
=
1
2
b
v
(2.37)
To calculate
i
km
= g
ir

kmr
note that for the metric a(u, v)du
2
+ b(u, v)dv
2
G
1
=
_
g
11
g
12
g
21
g
22
_
=
_
1
a(u,v)
0
0
1
b(u,v)
_
Hence

1
11
= g
11

111
=
a
u
2a
,
1
21
=
1
12
= g
11

121
=
a
v
2a
,
1
22
= g
11

221
=
b
u
2a

2
11
= g
22

112
=
a
v
2b
,
2
21
=
2
12
= g
22

122
=
b
u
2b
,
2
22
= g
22

222
=
b
v
2b
(2.38)
2.3.4 Example of the sphere again
Calculate Levi-Civita connection on the sphere.
On the sphere rst quadratic form (Riemannian metric) G = R
2
d
2
+
R
2
sin
2
d
2
. Hence we use calculations from the previous example with
49
a(, ) = R
2
, b(, ) = R
2
sin
2
(u = , v = ). Note that a

= a

= b

= 0.
Hence only non-trivial components of will be:

=
b

2a
=
sin 2
2
,
_

=
R
2
sin 2
2
_
, (2.39)

=
b

2b
=
cos
sin
_

=
R
2
sin 2
2
_
(2.40)
All other components are equal to zero:

= 0
Remark Note that Christoel symbols of Levi-Civita connection on the
sphere coincide with Christoel symbols of induced connection calculated in
the subsection Connection induced on surfaces. later we will understand
the geometrical meaning of this fact.
2.4 Levi-Civita connection = induced connection on
surfaces in E
3
Christoel symbols of canonical at connection vanish in Cartesian coordi-
nates. On the other hand standard Riemannian metric on Euclidean space
has the appearance: G = dx
i

ik
dx
k
. The matrix g
ik
has constant entries.
Hence according to Levi-Civita formula (2.34) the Christoel symbols of
Levi-Civita connection vanish also. Hence we proved very important fact:
Canonical at connection of Euclidean space is the Levi-Civita connection
of the standard metric on Euclidean space.
Now we show that Levi-Civita connection on surfaces in Euclidean space
coincides with the connection induced on the surfaces by canonical at con-
nection.
We perform our analysis for surfaces in E
3
.
Let M: r = r(u, v) be a surface in E
3
. Let G be induced Riemannian
metric on M and LeviCivita connection of this metric.
We know that the induced connection
(M)
is dened in the following
way: for arbitrary vector elds X, Y tangent to the surface M,
M
X
Y equals
to the projection on the tangent space of the vector eld
can.at
X
Y:

M
X
Y =
_

can.at
X
Y
_
tangent
,
50
where
can.at
is canonical at connection in E
3
(its Christoel symbols
vanish in cartesian coordinates). We denote by A
tangent
a projection of
the vector A attached at the point of the surface on the tangent space:
A

= An(A, n) , (n is normal unit vector eld to the surface.)


Theorem Induced connection on the surface r = r(u, v) in E
3
coincides
with Levi-Civita connection of Riemannian metric induced by the canonical
metric on Euclidean space E
3
.
Proof
Let
M
be induced connection on a surface M in E
3
given by equations
r = r(u, v). Considering this connection on the basic vectors r
h
, r
v
we see
that it is symmetric connection. Indeed

v
= (r
uv
)
tangent
= (r
vu
)
tangent
=
M

u
.
u
uv
=
u
vu
,
v
uv
=
v
vu
.
Prove that this connection preserves scalar product on M. For arbitrary
tangent vector elds X, Y, Z we have

X
Y, Z
E
3 =
can. at
X
Y, Z
E
3 +Y,
can. at
X
Z
E
3 .
since canonical at connection in E
3
preserves Euclidean metric in E
3
(it
is evident in Cartesian coordinates). Now project the equation above on
the surface M. If A is an arbitrary vector attached to the surface and
A
tangent
is its projection on the tangent space to the surface, then for ev-
ery tangent vector B scalar product A, B
E
3 equals to the scalar product
A
tangent
, B
E
3 = A
tangent
, B
M
since vector A A
tangent
is orthogonal to
the surface. Hence we deduce from (2) that
X
Y, Z
M
=

can. at
X
Y
_
tangent
, Z
E
3+Y,
_

can. at
X
Z
_
tangent

E
3 =
M
X
Y, Z
M
+Y,
M
X
Z
M
.
We see that induced connection is symmetric connection which preserves
the induced metric. Hence due to Levi-Civita Theorem it is unique and is
expressed as in the formula (2.34).
Remark One can easy to reformulate and prove more general statement: Let M be
a submanifold in Riemannian manifold (E, G). Then Levi-Civita connection of the metric
induced on this submanifold coincides with the connection induced on the manifold by
Levi-Civita connection of the metric G.
51
3 Parallel transport and geodesics
3.1 Parallel transport
3.1.1 Denition
Let M be a manifold equipped with ane connection .
Denition Let C: x(t) with coordinates x
i
= x
i
(t), t
0
t t
1
be a
curve on the manifold M. Let X = X(t
0
) be an arbitrary tangent vector
attached at the initial point x
0
(with coordinates x
i
(t
0
)) of the curve C, i.e.
X(t
0
) T
x
0
M is a vector tangent to the manifold M at the point x
0
with
coordinates x
i
(t
0
). (The vector X is not necessarily tangent to the curve C)
We say that X(t), t
0
t t
1
is a parallel transport of the vector X(t
0
)
T
x
0
M along the curve C: x
i
= x
i
(t), t
0
t t
1
if
For an arbitrary t, t
0
t t, vector X = X(t), (X(t)|
t=t
0
= X(t
0
)) is
a vector attached at the point x(t) of the curve C, i.e. X(t) is a vector
tangent to the manifold M at the point x(t) of the curve C.
The covariant derivative of X(t) along the curve C equals to zero:
X
dt
=
v
X = 0 . (3.1)
In components: if X
m
(t) are components of the vector eld X(t) and
v
m
(t) are components of the velocity vector v of the curve C ,
X(t) = X
m
(t)

x
m
|
x(t)
, v =
dx(t)
dt
=
dx
i
dt

x
m
|
x(t)
then the condition (3.1) can be rewritten as
dX
i
(t)
dt
+ v
k
(t)
i
km
(x
i
(t))X
m
(t) 0 . (3.2)
Remark We say sometimes that X(t) is covariantly constant along the
curve C If X(t) is parallel transport of the vector X along the curve C. If
we consider Euclidean space with canonical at connection then in Cartesian
coordinates Christoel symbols vanish and parallel transport is nothing but
dX
dt
=
v
X = 0, X(t) is constant vector.
52
Remark Compare this denition of parallel transport with the deni-
tion which we consider in the course of Introduction to Geometry where
we consider parallel transport of the vector along the curve on the surface
embedded in E
3
and dene parallel transport by the condition, that only
orthogonal component of the vector changes during parallel transport, i.e.
dX(t)
dt
is a vector orthogonal to the surface (see the Exercise in the Homework
7).
3.1.2

Parallel transport is a linear map
Consider two dierent points x
0
, x
1
on the manifold M with connection .
Let C be a curve x(t) joining these points. The parallel transport (3.1) X(t)
denes the map between tangent vectors at the point x
0
and tangent vectors
at the point x
1
. This map depends on the curve C. Parallel transport along
dierent curves joining the same points is in general dierent (if we are not
in Euclidean space).
On the other hand parallel transport is a linear map of tangent spaces
which does not depend on the parameterisation of the curve joining these
points.
Proposition Let X(t), t
0
t t
1
be a parallel transport of the vector
X(t
0
) T
x
0
M along the curve C: x = x(t), t
0
t t
1
, joining the points
x
0
= x(t
0
) and x
1
= x(t
1
). Then the map

C
: T
x
0
M X(t
0
) X(t
1
) T
x
1
M (3.3)
is a linear map from the vector space T
x
0
M to the vector space T
x
1
M which
does not depend on the parametersiation of the curve C.
The fact that the map (3.3) does not depend on the parameterisation
follows from the dierential equation (3.2) also.
Indeed let t = t(),
0

1
, t(
0
) = t
0
, t(
1
) =
1
be another
parameterisation of the curve C. Then multiplying the equation (3.2) on
dt
d
and using the fact that velocity v

() = t

v(t) we come to dierential


equation:
dX
i
(t())
d
+ v
k
(t())
i
km
(x
i
(t()))X
m
(t()) 0 . (3.4)
The functions X
(
t()) with the same initial conditions are the solutions
of this equation.
53
The fact that it is a linear map follows immediately from the fact that
dierential equations (3.2) are linear. E.g. let vector elds X(t), Y(t) be
covariantly constant along the curve C. Then since linearity X(t) + Y(t) is
a solution too.
3.1.3 Parallel transport with respect to Levi-Civita connection
We mostly consider parallel transport on Riemannian manifold. If (M, G) is
Riemannian manifold we mostly consider parallel transport with respect to
connection which is Levi-Civita connection of the Riemannian metric G
Proposition The length of the vector preserves during parallel transport
with respect to Levi-Civita connection.
Proof follows immediately from the denition (3.1) of aprallel traspotr
and the denition (2.33) of Levi-Civita connection
d
dt
X(t), X(t) =
v
X(t), X(t) = 2
v
X(t), X(t) = 2 0, X(t) = 0 .
(3.5)
Hence |X(t)| is constant.
Exercise Show that scalar product preserves during parallel transport.
3.2 Geodesics
3.2.1 Denition. Geodesic on Riemannian manifold.
Let M be manifold equipped with connection .
Denition A parameterised curve C: x
i
= x
i
(t) is called geodesic if
velocity vector v(t): v
i
(t) =
dx
i
(t)
dt
is covariantly constant along this curve,
i.e. it remains parallel along the curve:

v
v =
v
dt
=
dv
i
(t)
dt
+ v
k
(t)
i
km
(x(t))v
m
(t) = 0, i.e. (3.6)
d
2
x
i
(t)
dt
2
+
dx
k
(t)
dt

i
km
(x(t))
dx
m
(t)
dt
= 0 .
These are linear second order dierential equations. One can prove that
this equations have solution and it is unique
7
for an arbitrary initial data
(x
i
(t
0
) = x
i
0
, x
i
(t
0
) = x
i
0
. )
7
this is true under additional technical conditions which we do not discuss here
54
In other words the curve C: x(t) is a geodesic if parallel transport of
velocity vector to along the curve is a velocity vector at any point of the
curve.
Geodesics dened with Levi-Civita connection on the Riemannian mani-
fold is called geodesic on Riemannian manifold. We mostly consider geodesics
on Riemannian manifolds.
Since velocity vector of the geodesics on Riemannian manifold at any
point is a parallel transport with the Levi-Civita connection, hence due to
Proposition (3.5) the length of the velocity vector remains constant:
Proposition If C: x(t) is a geodesics on Riemannian manifold then the
length of velocity vector is preserved along the geodesic.
Proof Since the connection is Levi-Civita connection then it preserves
scalar product of tangent vectors, (see (2.33)) in particularly the length of
the velocity vector v:
3.2.2 Un-parameterised geodesic
We dened a geodesic as a parameterised curve such that the velocity vector
is covariantly constant along the curve.
What happens if we change the parameterisation of the curve?
Another question: Suppose a tangent vector to the curve remains tangent
to the curve during parallel transport. Is it true that this curve (in a suitable
parameterisation) becomes geodesic?
Denition We call un-parameterised curve geodesic if under suitable
parameterisation it obeys the equation (3.6) for geodesics.
Let Cbe un-parameterised geodesic. Then the following statement is
valid.
Proposition A curve C (un-parameterised) is geodesic if an only if a
non-zero vector tangent to the curve remains tangent to the curve during
parallel transport.
Proof. Let A be tangent vector at the point p C of the curve. Paral-
lel transport does not depend on parametersiation of the curve (see 3.1.2).
Choose a suitable parameterisation x
i
= x
i
(t) such that x
i
(t) obeys the equa-
tions (3.6) for geodesics, i.e. the velocity vector v(t) is covariantly constant
along the curve:
v
v = 0. If A(t
0
) = cv(t
0
) at the given point p (c is a scalar
55
coecient) then due to linearity A(t) = cv(t) is a parallel transport of the
vector A. The vector A(t) is tangent to the curve since it is proportional to
velocity vector. We proved that any tangent vector remains tangent during
parallel transport.
Now prove the converse: Let A(t) be a parallel transport of non-zero
vector and it is proportional to velocity. If in a given parameterisation A(t) =
c(t)v(t) choose a reparameterisation t = t() such that
dt()
d
= c(t). In the
new parameterisation the velocity vector v

() =
dt()
d
v(t()) = c(t)v(t) =
A(t()). We come to parameterisation such that velocity vector remains
covariantly constant. Thus we come to parameterised geodesic. Hence C is
a geodesic.
Remark In particularly it follows from the Proposition above the follow-
ing important observation:
Let C is un-parameterised geodesic, x
i
(t) be its arbitrary parameterisa-
tion and v(t) be velocity vector in this parameterisation. Then the velocity
vector remains parallel to the curve since it is a tangent vector.
In spite of the fact that velocity vector is not covariantly constant along
the curve, i.e. it will not remain velocity vector during parallel transport,
since it will be remain tangent to the curve during parallel transport.
Remark One can see that if x
i
= x
i
(t) is geodesic in an arbitrary pa-
rameterisation and s = s(t) is a natural parameter (which denes the length
of the curve) then x
i
(t(s)) is parameterised geodesic.
3.2.3 Geodesics on surfaces in E
3
Let M: r = r(u, v) be a surface in E
3
. Let G
M
be induced Riemannian
metric and a Levi-Civita connection on M. We consider on M Levi-Civita
connection of the metric G
M
.
Let C be an arbitrary geodesic and v(t) =
dr(t)
dt
the velocity vector. Ac-
cording to the denition of geodesic
v
v = 0. On the other hand we know
that Levi-Civita connection coincides with the connection induced on the
surface by canonical at connection in E
3
(see the Theorem in subsection
2.4). Hence

v
v = 0 =
M
v
v =
_

can.at
v
v
_
tangent
(3.7)
In cartesian coordinates
can.at
v
v =
v
v =
d
dt
v(u(t), v(t)) =
d
2
r(t)
dt
2
= a.
Hence according to (3.7) the tangent component of acceleration equals to
zero.
56
Converse if for the curve r(t) = r(u(t), v(t)) the acceleration vector a(t)
is orthogonal to the surface then due to (3.7)
v
v = 0.
We come to very beautiful observation:
Proposition The acceleration vector of an curve r = r(u(t), v(t)) on M
is orthogonal to the surface M if and only if this curve is geodesic.
In other words due to Newton second law particle moves along along
geodesic on the surface if and only if the force is orthogonal to the surface.
One can very easy using this Proposition to calculate geodesics of cylinder
and sphere.
Geodesic on the cylinder
Let r(h(t), (t)) be a geodesic on the cylinder
_

_
x = a cos
y = a sin
z = h
. We have
v =
dr
dt
=
_
_
a sin
a cos

h
_
_
and for acceleration:
a =
dv
dt
=
_
_
_
a

sin
a

cos

h
_
_
_
. .
tangent acceleration
+
_
_
a
2
cos
a
2
sin
0
_
_
. .
normal acceleration
Since tangential acceleration equals to zero hence
d
2
h
dt
2
= 0 and h(t) = h
0
+ct
Normal acceleration is centripetal acceleration of the rotation over circle with
constant speed (projection on the plane OXY ). The geodesic is helix.
Geodesics on sphere
Let r = r((t), (t)) be a geodesic on the sphere of the radius a: r(, ):
_

_
x = a sin cos
y = a sin sin
z = a cos
Consider the vector product of the vectors r(t) and velocity vector v(t)
M(t) = r(t) v(t). Acceleration vector a(t) is proportional to the r(t) since
due to Proposition it is orthogonal to the surface of the sphere. This implies
that M(t) is constant vector:
d
dt
M(t) =
d
dt
(r(t) v(t)) = (v(t) v(t)) + (r(t) a(t)) = 0 (3.8)
57
We have M(t) = M
0
. r(t) is orthogonal to M = r(t) v(t). We see
that r(t) belongs to the sphere and to the plane orthogonal to the vector
M
0
= r(t) v(t). The intersection of this plane with sphere is a great
circle. We proved that if r(t) is geodesic hence it belongs to great circle (as
un-parameterised curve).
The converse is evident since if particle moves along the great circle with
constant velocity then obviously acceleration vector is orthogonal to the sur-
face.
Remark The vector M = r(t) v(t) is the torque. The torque is inte-
gral of motion in isotropic space.This is the core of the considerations for
geodesics on the sphere.
3.3 Geodesics and Lagrangians of free particle on
Riemannian manifold.
3.3.1 Lagrangian and Euler-Lagrange equations
A function L = L(x, x) on points and velocity vectors on manifold M is a
Lagrangian on manifold M.
We assign to Lagrangian L = L(x, x) the following second order dieren-
tial equations
d
dt
_
L
x
i
_
=
L
x
i
(3.9)
In detail
d
dt
_
L
x
i
_
=

2
L
x
m
x
i
x
m
+

2
L
x
m
x
i

x
m
=
L
x
i
. (3.10)
These equations are called Euler-Lagrange equations of the Lagrangian
L. We will explain later the variational origin of these equations
8
.
3.3.2 Lagrangian of free particle
Let (M, G), G = g
ik
dx
i
dx
k
be a Riemannian manifold.
8
To every mechanical system one can put in correspondence a Lagrangian on cong-
uration space. The dynamics of the system is described by Euler-Lagrange equations.
The advantage of Lagrangian approach is that it works in an arbitrary coordinate system:
Euler-Lagrange equations are invariant with respect to changing of coordinates since they
arise from variational principe.
58
Denition We say that Lagrangian L = L(x, x) is the Lagrangian of a
free particle on the Riemannian manifold M if
L =
g
ik
x
i
x
k
2
(3.11)
Example Free particle in Euclidean space. Consider E
3
with standard
metric G = dx
2
+ dy
2
+ dz
2
L =
g
ik
x
i
x
k
2
=
x
2
+ y
2
+ z
2
2
(3.12)
Note that this is the Lagrangian that describes the dynamics od free
particle.
Example Free particle on sphere.
The metric on the sphere of radius R is G = R
2
d
2
+ R
2
sin
2
d
2
. Re-
spectively for the Lagrnagian of free particle we have
L =
g
ik
x
i
x
k
2
=
R
2

2
+ R
2
sin
2

2
2
(3.13)
3.3.3 Equations of geodesics and Euler-Lagrange equations
Theorem. Euler-Lagrange equations of Lagrangian of free particle are equiv-
alent to the second order dierential equations for geodesics.
This Theorem makes very easy calculations for Christoel indices.
This Theorem can be proved by direct calculations.
Calculate Euler-Lagrange equations (3.9) for the Lagrangian (3.11):
d
dt
_
L
x
i
_
=
d
dt
_
_

_
g
mk
x
m
x
k
2
_
x
i
_
_
=
d
dt
_
g
ik
x
k
_
= g
ik

x
k
+
g
ik
x
m
x
m
x
k
and
L
x
i
=

_
g
mk
x
m
x
k
2
_
x
i
=
1
2
g
mk
x
i
x
m
x
k
.
Hence we have
d
dt
_
L
x
i
_
= g
ik

x
k
+
g
ik
x
m
x
m
x
k
=
L
x
i
=
1
2
g
mk
x
i
x
m
x
k
,
59
i.e.
g
ik

x
k
+
m
g
ik
x
m
x
k
=
1
2

i
g
mk
x
m
x
k
.
Note that
m
g
ik
x
m
x
k
=
1
2
_

m
g
ik
x
m
x
k
+
k
g
im
x
m
x
k
_
. Hence we come to
equation:
g
ik
d
2
x
k
dt
2
+
1
2
(
m
g
ik
+
k
g
im

i
g
mk
) x
m
x
k
Multiplying on the inverse matrix g
ik
we come
d
2
x
i
dt
2
+
1
2
g
ij
_
g
jm
x
k
+
g
jk
x
m

g
mk
x
j
_
dx
m
dt
dx
k
dt
= 0 . (3.14)
We recognize here Christoel symbols of Levi-Civita connection (see (2.34))
and we rewrite this equation as
d
2
x
i
dt
2
+
dx
m
dt

i
mk
dx
k
dt
= 0 . (3.15)
This is nothing but the equation (3.6).
Applications of this Theorem: calculation of Christoel symbols of Levi-
Civita connection.
3.3.4 Examples of calculations of Christoel symbols and geodesics
using Lagrangians.
Consider two examples: We calculate Levi-Civita connection on sphere in E
3
and on Lobachevsky plane using Lagrangians and nd geodesics.
1) Sphere of the radius R in E
3
:
Lagrangian of free particle on the sphere is given by (3.13):
L =
R
2

2
+ R
2
sin
2

2
2
Euler-Lagrange equations dening geodesics are
d
dt
_
L

=
d
dt
_
R
2

_
R
2
sin cos
2

sin cos
2
= 0 ,
(3.16)
d
dt
_
L

_

=
d
dt
_
R
2
sin
2

_
= 0

+ cotan

= 0 .
60
Comparing Euler-Lagrange equations with equations for geodesic in terms of
Christoel symbols:

2
+ 2


2
= 0,

2
+ 2


2
= 0
we come to

= 0 ,

= sin cos , (3.17)

= 0,

= cotan . (3.18)
(Compare with previous calculations for connection in subsections 2.2.1 and
2.3.4)
We proved that geodesics on the sphere are great circles. (see subsection 3.2.3 above).
Consider another more straightforward proof of this fact. To nd geodesics one have to
solve second order dierential equations (3.16):
One can see that the great circles: =
0
, =
0
+ t are solutions of second order
dierential equations (3.16) with initial conditions
(t)

t=0
=
0
,

(t)

t=0
= 1, (t)

t=0
=
0
,

(t)

t=0
= 0 . (3.19)
The rotation of the sphere is isometry, which does not change Levi-Civta connection.
Hence an arbitrary great circle is geodesic.
Prove that an arbitrary geodesic is an arc of great circle. Let the curve = (t), =
(t), 0 t t
1
be geodesic. Rotating the sphere we can come to the curve =

(t), =

(t), 0 t t
1
such that velocity vector at the initial time is direccted along meridian,
i.e. initial conditions are

(t)

t=0
=
0
,

(t)

t=0
= a,

(t)

t=0
=
0
,

(t)

t=0
= 0 . (3.20)
(Compare with initial conditions (3.19)) Second order dierential equations with boundary
conditions for coordinates and velocities at t = 0 have unique solution. The solutions of
second order dierential equations (3.16) with initial conditions (3.20) is a curve

(t) =

0
+ at,

(t) =
0
. It is great circle. Hence initial curve the geodesic = (t), = (t),
0 t t
1
is an arc of great circle too.
This is another proof that geodesics are great circles.
2) Lobachevsky plane.
Lagrangian of free particle on the Lobachevsky plane with metric G =
dx
2
+dy
2
y
2
is
L =
1
2
x
2
+ y
2
y
2
.
Euler-Lagrange equations are
61
L
x
= 0 =
d
dt
L
x
=
d
dt
_
x
y
2
_
=

x
y
2

2 x y
y
3
, i.e.

x
2 x y
y
= 0 ,
L
y
=
x
2
+ y
2
y
3
=
d
dt
L
y
=
d
dt
_
y
y
2
_
=

y
y
2

2 y
2
y
3
, i.e.

y +
x
2
y

y
2
y
= 0 .
Comparing these equations with equations for geodesics:

x
i
x
k

i
km
x
m
= 0
(i = 1, 2, x = x
1
, y = x
2
) we come to

x
xx
= 0,
x
xy
=
x
yx
=
1
y
,
x
yy
= 0,
y
xx
=
1
y
,
y
xy
=
y
yx
= 0,
y
yy
=
1
y
.
In a similar way as for a sphere one can nd geodesics on Lobachevsky plane. First we
note that vertical rays are geodesics. Then using the inversions with centre on the absolute
one can see that arcs of the circles with centre at the absolute (y = 0) are geodesics too.
3.3.5 Variational principe and Euler-Lagrange equations
Here very briey we will explain how Euler-Lagrange equations follow from
variational principe.
Let M be a manifold (not necessarily Riemannian) and L = L(x
i
, x
i
) be
a Lagrangian on it.
Denote my M
x
2
,t
2
x
1
,t
1
the space of curves (paths) such that they start at the
point x
1
at the time t = t
1
and end at the point x
2
at the time t = t
2
:
M
x
2
,t
2
x
1
,t
1
= {C: x(t), t
1
t t
2
, x(t
1
) = x
1
, x(t
2
) = x
2
} . (3.21)
Consider the following functional S on the space M
x
2
,t
2
x
1
,t
1
:
S [x(t)] =

t
2
t
1
L
_
x
i
(t), x
i
(t)
_
dt . (3.22)
for every curve x(t) M
x
2
,t
2
x
1
,t
1
.
This functional is called action functional.
Theorem Let functional S attaints the minimal value on the path x
0
(t)
M
x
2
,t
2
x
1
,t
1
, i.e.
x(t) M
x
2
,t
2
x
1
,t
1
S[x
0
(t)] S[x(t)] . (3.23)
62
Then the path x
0
(t) is a solution of Euler-Lagrange equations of the La-
grangian L:
d
dt
_
L
x
i
_
=
L
x
i
if x(t) = x
0
(t) . (3.24)
Remark The path x(t) sometimes is called extremal of the action func-
tional (3.22).
We will use this Theorem to show that the geodesics are in some sense
shortest curves
9
.
3.4 Geodesics and shortest distance.
Many of you know that geodesics are in some sense shortest curves. We
will give an exact meaning to this statement and prove it using variational
principe:
Let M be a Riemannian manifold.
Theorem Let x
1
and x
2
be two points on M. The shortest curve which
joins these points is an arc of geodesic.
Let C be a geodesic on M and x
1
C. Then for an arbitrary point x
2
C
which is close to the point x
1
the arc of geodesic joining the points x
1
, x
2
is
a shortest curve between these points
10
.
This Theorem makes a bridge between two dierent approach to geodesic:
the shortest disntance and parallel transport of velocity vector.
Sketch a proof:
Consider the following two Lagrangians: Lagrangian of a free particle
L
free
=
g
ik
(x) x
i
x
k
2
and the length Lagrangian
L
length
(x, x) =

g
ik
(x) x
i
x
k
=

2L
free
.
9
The statement of this Theorem is enough for our purposes. In fact in classical mechan-
ics another more useful statement is used: the path x
0
(t) is a solution of Euler-Lagrange
equations of the Lagrangian L if and only if it is the stationary point of the action
functional (3.22), i.e.
S[x
0
(t) +x(t)] S[x
0
(t) +x(t)] = 0(x(t)) (3.25)
for an arbitrary innitesimal variation of the path x
0
(t): x
(
t
1
) = x
(
t
2
) = 0.
10
More precisely: for every point x
1
C there exists a ball B

(x
1
) such that for an
arbitrary point x
2
C B

(x
1
) the arc of geodesic joining the points x
1
, x
2
is a shortest
curve between these points.
63
If C: x
i
(t), t
1
t t
2
is a curve on M then
Length of the curve C =

t
2
t
1
L
length
(x
i
(t), x
i
(t))dt =

t
2
t
1

g
ik
(x(t)) x
i
(t) x
k
(t)dt . (3.26)
The proof of the Theorem follows from the following observation:
Observation Euler-Lagrange equations for the length functional (3.26) are
equivalent to the Euler-Lagrange equations for action functional (3.22). This
means that extremals of the length functional and action functionals coincide.
Indeed it follows from this observation and the variational principe that
the shortest curves obey the Euler-Lagrange equations for the action func-
tional. We showed before that Euler-Lagrange equations for action functional
(3.22) dene geodesics. Hence the shortest curves are geodesics.
One can check the observation by direct calculation: Calculate Euler-
Lagrange equations for the Lagrangian L
length
=

g
ik
(x) x
i
x
k
=

2L
free
:
d
dt
_
L
length
x
i
_

L
length
x
i
=
d
dt
_
1

g
ik
x
i
x
k
g
ik
x
k
_

1
2

g
ik
x
i
x
k
g
km
x
k
x
m
x
i
=
d
dt
_
1
L
length
L
free

x
i
_

1
L
length
L
free
x
i
= 0 . (3.27)
To facilitate calculations note that the length functional (3.26) is repa-
rameterisation invariant. Choose the natural parameter s(t) or a parame-
ter proportional to the natural parameter on the curve x
i
(t). We come to
L
length
= const and it follows from (3.27) that
d
dt
_
L
length
x
i
_

L
length
x
i
=
1
L
length
_
d
dt
_
L
free

x
i
_

L
free
x
i
_
= 0 .
We prove that Euler-Lagrange equations for length and action Lagrangians
coincide.
In the Euclidean space straight lines are the shortest distances between
two points. On the other hand their velocity vectors are constant. We realise
now that in general Riemannian manifold the role of geodesic is twofold also:
they are locally shortest and have covariantly constant velocity vectors.
64
3.4.1 Again geodesics for sphere and Lobachevsky plane
The fact that geodesics are shortest gives us another tool to calculate geodesics.
Consider again examples of sphere and Lobachevsky plane and nd geodesics
using the fact that they are shortest. The fact that geodesics are locally the
shortest curves
Consider again sphere in E
3
with the radius R: Coordinates , , induced
Riemannian metrics (rst quadratic form):
G = R
2
(d
2
+ sin
2
d
2
) . (3.28)
Consider two arbitrary points A and B on the sphere. Let (
0
,
0
) be coor-
dinates of the point A and (
1
,
1
) be coordinates of the point B
Let C
AB
be a curve which connects these points: C
AB
: (t), (t) such
that (t
0
) =
0
, (t
1
) =
1
, (t
0
) =
0
, (t
1
) =
1
then:
L
C
AB
=

2
t
+ sin
2
(t)
2
t
dt (3.29)
Suppose that points A and B have the same latitude, i.e. if (
0
,
0
) are
coordinates of the point A and (
1
,
1
) are coordinates of the point B then

0
=
1
(if it is not the fact then we can come to this condition rotating the
sphere)
Now it is easy to see that an arc of meridian, the curve =
0
is geodesics:
Indeed consider an arbitrary curve (t), (t) which connects the points A, B:
(t
0
) = (t
1
) =
0
, (t
0
) = (t
1
) =
0
. Compare its length with the length
of the meridian which connects the points A, B:

t
1
t
0
R

2
t
+ sin
2

2
t
dt R

t
1
t
0

2
t
dt = R

t
1
t
0

t
dt = R(
1

0
) (3.30)
Thus we see that the great circle joining points A, B is the shortest. The
great circles on sphere are geodesics. It corresponds to geometrical intuition:
The geodesics on the sphere are the circles of intersection of the sphere with
the plane which crosses the centre.
Geodesics on Lobachevsky plane
Riemannian metric on Lobachevsky plane:
G =
dx
2
+ dy
2
y
2
(3.31)
65
The length of the curve : x = x(t, y = y(t)) is equal to
L =


x
2
t
+ y
2
t
y
2
(t)
dt
In particularly the length of the vertical interval [1, ] tends to innity if
0:
L =


x
2
t
+ y
2
t
y
2
(t)
dt =

1
t
2
dt = log
1

One can see that the distance from every point to the line y = 0 is equal to
innity. This motivates the fact that the line y = 0 is called absolute.
Consider two points A = (x
0
, y
0
), B = (x
1
, y
1
) on Lobachevsky plane.
It is easy to see that vertical lines are geodesics of Lobachevsky plane.
Namely let points A, B are on the ray x = x
0
. Let C
AB
be an arc of
the ray x = x
0
which joins these points: C
AB
: x = x
0
, y = y
0
+ t Then
it is easy to see that the length of the curve C
AB
is less or equal than the
length of the arbitrary curve x = x(t), y = y(t) which joins these points:
x(t)

t=0
= x
0
, y(t)

t=0
= y
0
, x(t)

t=t
1
= x
0
, y(t)

t=t
1
= y
1
:

t
0

x
2
t
+ y
2
t
y
2
(t)
dt

t
0

y
2
t
y
2
(t)
dt =

y
1
y
0
dt
t
dt = log
y
1
y
0
= length of C
AB
Hence C
AB
is shortest. We prove that vertical rays are geodesics.
Consider now the case if x
0
= x
1
. Find geodesics which connects two
points A, B which are not on the same vertical ray. Consider semicircle
which passes these two points such that its centre is on the absolute. We
prove that it is a geodesic.
Proof Let coordinates of the centre of the circle are (a, 0). Then consider polar coor-
dinates (r, ):
x = a +r cos , y = r sin (3.32)
In these polar coordinates r-coordinate of the semicircle is constant.
Find Lobachevsky metric in these coordinates: dx = r sin d + cos dr, dy =
r cos d + sin dr, dx
2
+dy
2
= dr
2
+r
2
d
2
. Hence:
G =
dx
2
+dy
2
y
2
=
dr
2
+r
2
d
2
r
2
sin
2

==
d
2
sin
2

+
dr
2
r
2
sin
2

(3.33)
66
We see that the length of the arbitrary curve which connects points A, B is greater or
equal to the length of the arc of the circle:
L
AB
=

t
1
t
0


2
t
sin
2

+
r
2
t
r
2
sin
2

dt

t
1
t
0


2
t
sin
2

dt = (3.34)

t
1
t
0

t
sin
dt =

0
d
sin
= log
tan
1
tan
1
The proof is nished.
4 Surfaces in E
3
Now equipped by the knowledge of Riemannian geometry we consider sur-
faces in E
3
. We reconsider again conceptions of Shape (Weingarten) opera-
tor, Gaussian and mean curvatures, focusing attention on the the fact what
properties are internal and what properties are external.
4.1 Formulation of the main result. Theorem of paral-
lel transport over closed curve and Theorema Egregium
Let M be a surface in Euclidean space E
3
. Consider a closed curve C on M,
M: r = r(u, v), C: r = r(u(t, v(t)), 0 t t
1
, x(0) = x(t
1
). (u(t), v(t) are
internal coordinates of the curve C.)
Consider the parallel transport of an arbitrary tangent X vector along
the closed curve C:
X(t) = X

(t)

u

(t)
. .
Internal observer
= X

(t)r

r(u(t),v(t))
. .
External observer
,
_
r

=
x
i
u

x
i
_
.
X(t):
X(t)
dt
= 0, 0 t t
1
,
i.e.
dX

(t)
dt
+ X

(t)

(u(t))
du

(t)
dt
= 0, 0 t t
1
, (4.1)
where is the connection induced on the surface M by canonical at con-
nection (see (??)), i.e. the Levi-Civita connection of the induced Riemannian
67
metric on the surface M and

its Christoel symbols:

=
1
2
g

_
g

+
g

_
, where g

= r

, r

=
x
i
u

x
i
u

(4.2)
are components of induced Riemannian metric G
M
= g

du

du

/
Let r(0) = p be a starting (and ending) point of the curve C: r(0) =
r(t
1
) = p. The dierential equation (4.1) denes the linear operator
R
C
: T
p
M T
p
M (4.3)
For any vector X T
p
M, its image the vector R
C
X as the solution of the
dierential equation (4.1) with initial condition X(t)

t=0
= X.
On the other hand we know that parallel transport of the vector does not
change its length (see (3.5) in the subsection 3.2.1):
X, X = R
C
X, R
C
X (4.4)
We see that R
C
is an orthogonal operator in the 2-dimensional vector space
T
p
M. We know that orthogonal operator preserving orientation is the oper-
ator of rotation on some angle .
We see that if R
C
preserves orientation
11
then the action of operator R
C
on vectors is rotation on the angle, i.e. the result of parallel transport along
closed curve is rotation on the angle
= (C) (4.5)
which depends on the curve.
The very beautiful question arises: How to calculate this angle (C)
Theorem Let M be a surface in Euclidean space E
3
. Let C be a closed
curve C on M such that C is a boundary of a compact oriented domain
D M. Consider the parallel transport of an arbitrary tangent vector along
the closed curve C. As a result of parallel transport along this closed curve
any tangent vector rotates through the angle
= (X, R
C
X) =

D
Kd , (4.6)
11
We consider the case if the operator R
C
preserves orientation. In our considerations
we consider only the case if the closed curve C is a boundary of a compact oriented domain
D M. In this case one can see that operator R
C
preserves an orientation.
68
where K is the Gaussian curvature and d =

det gdudv is the area element


induced by the Riemannian metric on the surface M, i.e. d =

det gdudv.
Remark One can show that the angle of rotation does not depend on
initial point of the curve.
Example Consider the closed curve, latitude C

0
: =
0
on the sphere
of the radius R. Calculations show that
(C

0
) = 2(1 cos
0
) (4.7)
(see the Homework 6). On the other hand the latitude C

0
is the boundary
of the segment D with area 2RH where H = R(1 cos
0
). Hence
(X, R
C
X) =
2RH
R
2
=
1
R
2
area of the segment =

D
Kd
since Gaussian curvature is equal to
1
R
2
The proof of this Theorem is the content of the subsections above.
4.1.1 GauTheorema Egregium
Show that this Theorem implies the remarkable corollary.
Let D be a small domain around a given point p, let C its boundary
and (D) the angle of rotation. Denote by S(D) an area of this domain.
Applying the Theorem for the case when area of the domain D tends to zero
we we come to the statement that
if S(D) 0 then (D) =

D
Kd K(p)S(D), i.e.
K(p) = lim
S(D)0
(D)
S(D)
, i.e. (4.8)
Now notice that left hand side od this equation dening Gaussian cur-
vature K(p) depends only on Riemannian metric on the surface C. Indeed
numerator of LHS is dened by the solution of dierential equation (4.1)
which depends on Levi-Civita connection depending on the induced Rieman-
nian metric, and denominator is an area depending on Riemannian metric
too.
Thus we come to very important
69
Corollary Gau Egregium Theorema
Gaussian curvature of the surface is invariant of isometries.
In next subsections we develop the technique which itself is very interest-
ing. One of the applications of this technique is the proof of the Theorem
(4.6). Thus we will prove Theorema Egregium too.
Later in the fth section we will give another proof of the Theorema
Egregium.
4.2 Derivation formulae
Let M be a surface embedded in Euclidean space E
3
, M: r = r(u, v).
Let e, f, n be three vector elds dened on the points of this surface such
that they form an orthonormal basis at any point, so that the vectors e, f
are tangent to the surface and the vector n is orthogonal to the surface
12
.
Vector elds e, f, n are functions on the surface M:
e = e(u, v), f = f(u, v) , n = n(u, v) .
Consider 1-forms de, df , dn:
de =
e
u
du +
e
v
dv, df =
f
u
du +
f
v
dv , dn =
n
u
du +
n
v
dv
These 1-forms take values in the vectors in E
3
, i.e. they are vector valued
1-forms. Any vector in E
3
attached at an arbitrary point of the surface can
be expanded over the basis {e, f, n}. Thus vector valued 1-forms de, df , dn
can be expanded in a sum of 1-forms with values in basic vectors e, f , n. E.g.
for de =
e
u
du+
e
v
dv expanding vectors
e
u
and
e
v
over basis vectors we come
to
e
u
= A
1
(u, v)e+B
1
(u, v)f+C
1
(u, v)n,
e
v
= A
2
(u, v)e+B
2
(u, v)f +C
2
(u, v)n
thus
de =
e
u
du +
e
v
dv = (A
1
e + B
1
f + C
1
n) du + (A
2
e + B
2
f + C
2
n) dv =
12
One can say that {e, f , n} is an orthonormal basis in T
p
E
3
at every point of surface
p M such that {e, f } is an orthonormal basis in T
p
E
3
at every point of surface p M.
70
= (A
1
du + A
2
dv)
. .
M
11
e + (B
1
du + B
2
dv)
. .
M
12
f + (C
1
du + C
2
dv)
. .
M
11
e, (4.9)
i.e.
de = M
11
e + M
12
f + M
13
n,
where M
11
, M
12
and M
13
are 1-forms on the surface M dened by the relation
(4.9).
In the same way we do the expansions of vector-valued 1-forms df and
dn we come to
de = M
11
e + M
12
f + M
13
n
df = M
21
e + M
22
f + M
23
n
dn = M
31
e + M
32
f + M
33
n
This equation can be rewritten in the following way:
d
_
_
e
f
n
_
_
=
_
_
M
11
M
12
M
13
M
21
M
22
M
23
M
31
M
32
M
33
_
_
_
_
e
f
n
_
_
(4.10)
Proposition The matrix M in the equation (4.10) is antisymmetrical ma-
trix, i.e.
M
11
= M
22
= M
33
= 0
M
12
= M
21
= a
M
13
= M
31
= b
M
23
= M
32
= b
(4.11)
i.e.
d
_
_
e
f
n
_
_
=
_
_
0 a b
a 0 c
b c 0
_
_
_
_
e
f
n
_
_
, (4.12)
where a, b, c are 1-forms on the surface M.
Formulae (5.13) are called derivation formulae.
Prove this Proposition. Recall that {e, f, n} is orthonormal basis, i.e. at
every point of the surface
e, e = f , f = n, n = 1, and e, f = e, n = f , n = 0
Now using (4.10) we have
e, e = 1 de, e = 0 = 2e, de = e, M
11
e + M
12
f + M
13
n =
71
M
11
e, e + M
12
e, f + M
13
e, n = M
11
M
11
= 0 .
Analogously
f , f = 1 df , f = 0 = 2f , df = f , M
21
e+M
22
f+M
23
n = M
22
M
22
= 0 ,
n, n = 1 dn, n = 0 = 2n, dn = n, M
31
e+M
32
f+M
33
n = M
33
. M
33
= 0 .
We proved already that M
11
= M
22
= M
33
= 0. Now prove that M
12
=
M
21
, M
13
= M
31
and M
13
= M
31
.
e, f = 0 de, f = 0 = e, df +de, f =
e, M
21
e + M
22
f + M
23
n +M
11
e + M
12
f + M
13
n, f = M
21
+ M
12
= 0.
Analogously
e, n = 0 de, n = 0 = e, dn +de, n =
e, M
31
e + M
32
f + M
33
n +M
11
e + M
12
f + M
13
n, n = M
31
+ M
13
= 0
and
f, n = 0 df, n = 0 = f , dn +df, n =
f, M
31
e + M
32
f + M
33
n +M
21
e + M
22
f + M
23
n, n = M
32
+ M
23
= 0.
Remark This proof could be performed much more shortly in condensed
notations. Derivation formulae (5.13) in condensed notations are
de
i
= M
ik
e
k
(4.13)
Orthonormality condition means that e
i
, e
k
=
ik
. Hence
de
i
, e
k
= 0 = de
i
, e
k
+e
i
, de
k
= M
im
e
m
, e
k
+e
i
, M
kn
e
n
= M
ik
+M
ki
= 0
(4.14)
Much shorter, is not it?
4.2.1

Gauss condition (structure equations)
Derive the relations between 1-forms a, b and c in derivation formulae.
Recall that a, b, c are 1-forms, e, f, n are vector valued functions (0-forms)
and de, df, dn are vector valued 1-forms. (We use the simple identity that
72
ddf = 0 and the fact that for 1-form = 0.) We have from derivation
formulae (5.13) that
d
2
e = 0 = d(af + bn) = daf a df + dbn b dn =
da f a (ae + cn) + db n b (be cf) =
(da+bc)f +(aa+bb)e+(dbac)n = (da+bc)f +(db+ca)n = 0 .
We see that
(da + b c)f + (db + c a)n = 0 (4.15)
Hence components of the left hand side equal to zero:
(da + b c) = 0 (db + c a) = 0 . (4.16)
Analogously
d
2
f = 0 = d(ae + cn) = dae + a de + dcn c dn =
dae + a (af + bn) + dc n c (be cf) =
(da + c b)e + (dc + a b)n = 0 .
Hence we come to structure equations:
da + b c = 0
db + c a = 0
dc + a a = 0
(4.17)
4.3 Geometrical meaning of derivation formulae. Wein-
garten operator and second quadratic form in terms
of derivation formulae.
Let M be a surface in E
3
.
Let e, f, n be three vector elds dened on the points of this surface such
that they form an orthonormal basis at any point, so that the vectors e, f are
tangent to the surface and the vector n is orthogonal to the surface. Note
that in generally these vectors are not coordinate vectors.
Describe Riemannian geometry on the surface M in terms of this basis
and derivation formulae (5.13).
Induced Riemannian metric
73
If G is the Riemannian metric induced on the surface M then since e, f
is orthonormal basis at every tangent space T
p
M then
G(e, e) = G(f , f) = 1, G(e, f ) = G(f , e) = 0 (4.18)
The matrix of the Riemannian metric in the basis {e, f} is
G =
_
1 0
0 1
_
(4.19)
Induced connection Let be the connection induced by the canonical at
connection on the surface M.
Then according equations (??) and derivation formulae (5.13) for every
tangent vector X

X
e = (
X
e)
tangent
= (de(X))
tangent
= (a(X)f + b(X)n)
tangent
= a(X)f .
(4.20)
and

X
f = (
X
f )
tangent
= (df(X))
tangent
= (a(X)e + c(X)n)
tangent
= a(X)e .
(4.21)
In particular

e
e = a(e)f
f
e = a(f)f

e
f = a(e)e
f
f = a(f)e
(4.22)
We know that the connection is Levi-Civita connection of the induced
Riemannian metric (4.20) (see the subsection 4.2.1)
13
.
Second Quadratic form Second quadratic form is a bilinear symmetric function A(X, Y)
on tangent vectors which is well-dened by the condition A(X, Y)n = (
X
Y)
orthogonal
(see
e.g. subsection 6.4 in Appendices.)
Let A(X, Y) be second quadratic form. Then according to derivation formulae (5.13)
we have
A(e, e) =
e
e, n = de(e), n = a(e)f +b(e)n, n = b(e) ,
A(f , e) =
f
e, n = de(f ), n = a(f )f +b(f )n, n = b(f ) ,
A(e, f ) =
e
f , n = df (e), n = a(e)f +c(e)n, n = c(e) ,
13
In particular this implies that this is symmetric connection, i.e.
fe
e
f [f , e] = a(f )f +a(e)e [f , e] = 0 . (4.23)
74
A(f , f ) =
f
f , n = df (f ), n = a(f )f +c(f )n, n = c(f ) ,
The matrix of the second quadratic form in the basis {e, f } is
A =
_
A(e, e) A(f , e)
A(e, f ) A(f , f )
_
=
_
b(e) b(f )
c(e) c(f )
_
(4.24)
This is symmetrical matrix (see the subsection 4.3.2):
A(f , e) = b(f ) = A(e, f ) = c(e) . (4.25)
Weingarten operator
Let S be Weingarten operator: SX =
X
n (see the subsection 6.4 in
Appendix, or last year Geometry letures). Then it follows from derivation
formulae that
SX =
X
n = dn(X) = (b(X)e c(X)f) = b(X)e + c(X)f
In particular
S(e) = b(e)e + c(e)f , S(f) = b(f)e + c(f)f
and the matrix of the Weingarten operator in the basis {e, f} is
S =
_
b(e) c(e)
b(f) c(f)
_
(4.26)
Remark According to the condition (4.25) the matrix S is symmetrical. The relations
A = GS, S = G
1
A for Weingarten operator, Riemannian metric and second quadratic
form are evidently obeyed for matrices of these operators in the basis e, f where G = 1,
A = S.
4.3.1 Gaussian and mean curvature in terms of derivation formu-
lae
Now we are equipped to express Gaussian and mean curvatures in terms of
derivation formulae. Using (4.26) we have for Gaussian curvature
K = det S = b(e)c(f) c(e)b(f ) = (b c)(e, f ) (4.27)
and for mean curvature
H = Tr S = b(e) + c(f ) (4.28)
What next? We will study in more detail formula (4.27) later.
Now consider some examples of calculation derivation formulae, Wein-
garten operator, e.t..c. for some examples using derivation formulae.
75
4.4 Examples of calculations of derivation formulae and
curvatures for cylinder, cone and sphere
Last year we calculated Weingarten oeprator, second quadratic form and
curvatures for cylinder, cone and sphere (see also the subsection 6.4 in Ap-
pendices.). Now we do the same but in terms of derivation formulae.
Cylinder
We have to dene three vector elds e, f , n on the points of the cylinder
r(h, ):
_

_
x = Rcos
y = Rsin
z = h
such that they form an orthonormal basis at any
point, so that the vectors e, f are tangent to the surface and the vector n is
orthogonal to the surface. We calculated many times coordinate vector elds
r
h
, r

and normal unit vector eld:


r
h
=
_
_
0
0
1
_
_
, r

=
_
_
Rsin
Rcos
0
_
_
, n =
_
_
cos
sin
0
_
_
. (4.29)
Vectors r
h
, r

and n are orthogonal to each other but not all of them have
unit length. One can choose
e = r
h
=
_
_
0
0
1
_
_
, f =
r

R
=
_
_
sin
cos
0
_
_
, n =
_
_
cos
sin
0
_
_
(4.30)
These vectors form an orthonormal basis and e, f form an orthonormal basis
in tangent space.
Derive for this basis derivation formulae (5.13). For vector elds e, f, n
in (4.30) we have
de = 0, df = d
_
_
sin
cos
0
_
_
=
_
_
cos
sin
0
_
_
d = nd,
dn = d
_
_
cos
sin
0
_
_
=
_
_
sin
cos
0
_
_
= f d,
76
i.e.
d
_
_
e
f
n
_
_
=
_
_
0 a b
a 0 c
b c 0
_
_
_
_
e
f
n
_
_
=
_
_
0 0 0
0 0 d
0 d 0
_
_
_
_
e
f
n
_
_
, (4.31)
i.e. in derivation formulae a = b = 0, c = d.
The matrix of Weingarten operator in the basis {e, f} is
S =
_
b(e) c(e)
b(f) c(f )
_
=
_
0 d(e)
0 d(f)
_
=
_
0 0
0
1
R
_
According to (4.27) and(4.28) Gaussian curvature K = b(e)c(f )b(e)c(f ) = 0
and mean curvature
H = b(e) + c(f) = d(f ) = d
_
r

R
_
=
1
R
(Compare with calculations in the subsection 4.3.4)
Cone
For cone:
r(h, ):
_

_
x = khcos
y = khsin
z = h
,
r
h
=
_
_
k cos
k sin
1
_
_
, r

=
_
_
khsin
khcos
0
_
_
, n =
1

1 + k
2
_
_
cos
sin
k
_
_
Tangent vectors r
h
, r

are orthogonal to each other. The length of the vector


r
h
equals to

1 + k
2
and the length of the vector r

equals to kh. Hence we


can choose orthonormal basis {e, f, n} such that vectors e, f are unit vectors
in the directions of the vectors r
h
, r

:
e =
r
h

1 + k
2
=
1

1 + k
2
_
_
k cos
k sin
1
_
_
, f =
r

hk
=
_
_
sin
cos
0
_
_
, n =
1

1 + k
2
_
_
cos
sin
k
_
_
Calculate de, df and dn:
de = d
_
_
k cos
k sin
1
_
_
=
kd

1 + k
2
_
_
sin
cos
0
_
_
=
kd

1 + k
2
f ,
77
df = d
_
_
sin
cos
0
_
_
=
_
_
cos
sin
0
_
_
d =
k
1 + k
2
_
_
k cos
k sin
1
_
_
d
d
1 + k
2
_
_
cos
sin
k
_
_
=
kd

1 + k
2
e
d

1 + k
2
n,
and
dn =
1

1 + k
2
d
_
_
cos
sin
k
_
_
=
d

1 + k
2
_
_
sin
cos
0
_
_
.
We come to
d
_
_
e
f
n
_
_
=
_
_
0 a b
a 0 c
b c 0
_
_
_
_
e
f
n
_
_
=
_
_
_
0
kd

1+k
2
0

kd

1+k
2
0
d

1+k
2
0
d

1+k
2
0
_
_
_
_
_
e
f
n
_
_
,
(4.32)
i.e. in derivation formulae for 1-forms a =
kd

1+k
2
, b = 0 and and c =
d

1+k
2
.
The matrix of Weingarten operator in the basis {e, f} is
S =
_
b(e) c(e)
b(f ) c(f)
_
= S =
_
0
d(e)

1+k
2
0
d(f )

1+k
2
_
=
_
0 0
0
1
kh

1+k
2
_
.
since d(f) = d
_
r

kh
_
=
1
kh
d(

) =
1
kh
.
According to (4.27), (4.28) Gaussian curvature
K = b(e)c(f) b(e)c(f ) = 0
and mean curvature
H = b(e) + c(f) = d(f ) = d
_
r

R
_
=
1
R
Remark Equipped by the properties of derivation formulae we do not need
to calculate df. The calculation of de and dn and the property that the
matrix in derivation formulae is antisymmetric gives us the answer for df.
Sphere
78
For sphere
r(, ):
_

_
x = Rsin cos
y = Rsin sin
z = Rcos
(4.33)
r

(, ) =
r

=
_
_
Rcos cos
Rcos sin
Rsin
_
_
, r

(, ) =
r

=
_
_
Rsin sin
Rsin cos
0
_
_
,
n(, ) =
r
R
=
_
_
sin cos
sin sin
cos
_
_
.
Tangent vectors r

, r

are orthogonal to each other. The length of the vector


r

equals to R and the length of the vector r

equals to Rsin . Hence we


can choose orthonormal basis {e, f, n} such that vectors e, f are unit vectors
in the directions of the vectors r

, r

:
e(, ) =
r

R
=
_
_
cos cos
cos sin
sin
_
_
, f (, ) =
r

Rsin
=
_
_
sin
cos
0
_
_
, n(, ) =
r
R
=
_
_
sin cos
sin sin
cos
_
_
.
Calculate de, df and dn:
de = d
_
_
cos cos
cos sin
sin
_
_
=
_
_
cos sin
cos cos
0
_
_
d
_
_
cos cos
cos sin
sin
_
_
= cos df dn,
df = d
_
_
sin
cos
0
_
_
=
_
_
cos
sin
0
_
_
d =
cos d
_
_
cos cos
cos sin
sin
_
_
sin d
_
_
sin cos
sin sin
cos
_
_
= cos de sin dn,
dn = d
_
_
sin cos
sin sin
cos
_
_
=
_
_
cos cos
cos sin
sin
_
_
d +
_
_
sin sin
sin cos
0
_
_
d
79
= de + sin df .
i.e.
d
_
_
e
f
n
_
_
=
_
_
0 a b
a 0 c
b c 0
_
_
_
_
e
f
n
_
_
=
_
_
0 cos d d
cos d 0 sin d
d sin d 0
_
_
_
_
e
f
n
_
_
,
(4.34)
i.e. in derivation formulae a = cos d, b = d, c = sin d.
The matrix of Weingarten operator in the basis {e, f} is
S =
_
b(e) c(e)
b(f) c(f)
_
=
_
d(e) sin d(e)
d(f) sin d(f)
_
=
_
1
R
0
0
1
R
_
since d(e) = d
_

R
_
=
1
R
d(

) =
1
R
, d(e) = d
_

R
_
=
1
R
d(

) = 0.
According to (4.27) and(4.28) Gaussian curvature
K = b(e)c(f) b(e)c(f ) =
1
R
2
and mean curvature
H = b(e) + c(f) =
2
R
Notice that for calculation of Weingarten operator and curvatures we used
only 1-forms b and c, i.e. the derivation equation for dn, (dn = de +
sin df). (Compare with calculations in the subsection 4.3.4)
Mean curvature is dene up to a sign. If we change n n mean
curvature H
1
R
and Gaussian curvature will not change.
We see that for the sphere Gaussian curvature is not equal to zero, whilst
for cylinder and cone Gaussian curvature equals to zero.
Remark The same remark as for cone: equipped by the properties of
derivation formulae we do not need to calculate df. The calculation of de and
dn and the property that the matrix in derivation formulae is antisymmetric
gives us the answer for df .
4.5

Proof of the Theorem of parallel transport along
closed curve.
We are ready now to prove the Theorem. Recall that the Theorem states
following:
80
If C is a closed curve on a surface M such that C is a boundary of a
compact oriented domain D M, then during the parallel transport of an
arbitrary tangent vector along the closed curve C the vector rotates through
the angle
= (X, R
C
X) =

D
Kd , (4.35)
where K is the Gaussian curvature and d =

det gdudv is the area element


induced by the Riemannian metric on the surface M, i.e. d =

det gdudv.
(see (4.6).
Recall that for derivation formulae (5.13) we obtained structure equations
da + b c = 0
db + c a = 0
dc + a a = 0
(4.36)
We need to use only one of these equations, the equation
da + b c = 0 . (4.37)
This condition sometimes is called Gau condition.
Let as always {e, f, n} be an orthonormal basis in T
p
E
3
at every point
of surface p M such that {e, f} is an orthonormal basis in T
p
M at every
point of surface p M. Then the Gau condition (4.37) and equation (4.27)
mean that for Gaussian curvature on the surface M can be expressed through
the 2-form da and base vectors {e, f }:
K = b c(e, f ) = da(e, f) (4.38)
We use this formula to prove the Theorem.
Now calculate the parallel transport of an arbitrary tangent vector over
the closed curve C on the surface M.
Let r = r(u, v) = r(u) ( = 1, 2, (u, v) = (u
1
, v
1
)) be an equation of the
surface M.
Let u

= u

(t) ( = 1, 2) be the equation of the curve C. Let X(t) be


the parallel transport of vector eld along the closed curve C, i.e. X(t) is
tangent to the surface M at the point u(t) of the curve C and vector eld
X(t) is covariantly constant along the curve:
X(t)
dt
= 0
81
To write this equation in components we usually expanded the vector eld
in the coordinate basis {r
u
=
u
, r
v
=
v
} and used Christoel symbols of
the connection

.
Now we will do it in dierent way: instead coordinate basis {r
u
=
u
, r
v
=

v
} we will use the basis {e, f }. In the subsection 3.4.4 we obtained that the
connection has the following appearance in this basis

v
e = a(v)f , ,
v
f = a(v)e (4.39)
(see the equations (4.20) and (4.21))
Let
X = X(u(t)) = X
1
(t)e(u(t)) + X
2
(t)f(u(t))
Lbe an expansion of tangent vector eld X(t) over basis {e, f }. Let v be ve-
locity vector of the curve C. Then the equation of parallel transport
X(t)
dt
= 0
will have the following appearance:
X(t)
dt
= 0 =
v
_
X
1
(t)e(u(t)) + X
2
(t)f(u(t))
_
=
dX
1
(t)
dt
e(u(t)) + X
1
(t)
v
e(u(t)) +
dX
2
(t)
dt
f(u(t)) + X
2
(t)
v
f(u(t)) =
dX
1
(t)
dt
e(u(t)) + X
1
(t)a(v)f (u(t)) +
dX
2
(t)
dt
f(u(t)) X
2
(t)a(v)e(u(t)) =
_
dX
1
(t)
dt
X
2
(t)a(v)
_
e(u(t)) +
_
dX
2
(t)
dt
+ X
1
(t)a(v)
_
f(u(t)) = 0.
Thus we come to equation:
_

X
1
(t) a(v(t))X
2
= 0

X
2
(t) + a(v(t))X
1
= 0
There are many ways to solve this equation. It is very convenient to consider
complex variable
Z(t) = X
1
(t) + iX
2
(t)
We see that

Z(t) =

X
1
(t) + i

X
2
(t) = a(v(t))X
2
ia(v(t)X
1
= ia(v)Z(t),
82
i.e.
dZ(t)
dt
= ia(v(t))Z(t) (4.40)
The solution of this equation is:
Z(t) = Z(0)e
i

t
0
a(v())d
(4.41)
Calculate

t
1
0
a(v())d for closed curve u(0) = u(t
1
). Due to Stokes Theo-
rem:

t
1
0
a(v(t))dt =

C
a =

D
da
Hence using Gauss condition (4.37) we see that

t
1
0
a(v(t))dt =

C
a =

D
da =

D
b c
Claim

D
b c =

D
da =

Kd . (4.42)
Theorem follows from this claim:
Z(t
1
) = Z(0)e
i

C
a
= Z(0)e
i

D
bC
(4.43)
Denote the integral i

D
b C by : = i

D
b C. We have
Z(t
1
) = X
1
(t
1
) + iX
2
(t
1
) =
_
X
1
(0) + iX
2
(0)
_
e
i
= (4.44)
It remains to prove the claim. The induced volume form d is 2-form.
Its value on two orthogonal unit vector e, f equals to 1:
d(e, f ) = 1 (4.45)
(In coordinates u, v volume form d =

det gdu dv).


The value of the form b c on vectors {e, f } equals to Gaussian curvature
according to (4.38). We see that
b c(e, f) = da(e, f) = Kd(e, f)
Hence 2-forms bc, da and volume form d coincide. Thus we prove (4.42).
83
5 Curvtature tensor
5.1 Denition
Let X, Y, Z be an arbitrary vector elds on the manifold equipped with ane
connection .
Consider the following operation which assings to the vector elds X, Y
and Z the new vector eld.
R(X, Y)Z =
_

[X,Y]
_
Z (5.1)
This operation is obviously linear over the scalar coecients and it is C

(M)
with respect to vector elds X, YZ, i.e. for an arbitrary functions f, g, h
R(fX, gY)(hZ) = fghR(X, Y)Z, (5.2)
i.e. it denes the tensor eld of the type
_
1
3
_
: If X = X
i

i
, X = X
i

i
, X =
X
i

i
then according to (5.2)
R(X, Y)Z = R(X
m

m
, Y
n

n
)(Z
r

r
) = Z
r
R
i
rmn
X
m
Y
n
where we denote by R
i
rmn
the components of the tensor R in the coordinate
basis
i
R
i
rmn

i
= R(
m
,
n
)
r
(5.3)
Express components of the curvature tensor in terms of Christoel symbols
of the connection. If
m

n
=
r
mn

r
then according to the (5.1) we have:
R
i
rmn

i
= R(
m
,
n
)
r
=

r
,
R
i
rmn
=

m
(
p
nr

p
)

n
(
p
mr

p
) =

i
nr
+
i
mp

p
nr
(m n) (5.4)
The proof of the property (5.2) can be given just by straightforward
calculations: Consider e.g. the case f = g = 1, then
R(X, Y)(hZ) =
X

Y
(hZ)
Y

X
(hZ)
[X,Y]
(hZ) =

X
(
Y
hZ + h
Y
Z)
Y
(
X
hZ + h
X
Z)
[X,Y]
hZ h
[X,Y]
Z =

Y
hZ +
Y
h
X
Z +
X
h
Y
Z + h
X

Y
Z
84

X
hZ
X
h
Y
Z
Y
h
X
Z + h
Y

X
Z

[X,Y]
hZ h
[X,Y]
Z =
h
_

Y
Z
Y

X
Z)
[X,Y]
Z

+
_

Y
h
Y

X
h
[X,Y]h

Z =
h
X

Y
Z
Y

X
Z)
[X,Y]
Z = hR(X, Y)Z,
since
X

Y
h
Y

X
h
[X,Y]
h = 0.
5.1.1 Properties of curvature tensor
Tensor R
i
kmn
is expressed trough derivatives of Christoel symbols. In spite
this fact it is is much more pleasant object than Christoel symbols, since
the latter is not the tensor.
It follows from the denition that the tensor R
i
kmn
is antisymmetrical
with respect to indices m, n:
R
i
kmn
= R
i
knm
. (5.5)
One can prove that for symmetric connection this tensor obeys the following identities:
R
i
kmn
+R
i
mnk
+R
i
nkm
= 0 , (5.6)
The curvature tensor corresponding to Levi-Civita connection obeys also
another identities too (see the next subsection.)
We know well that If Christoel symbols vanish in a vicinity of a given
point p in some chosen coordinate system then in general Christoel symbols
do not vanish in arbitrary coordinate systems. E.g. Christoel symbols of
canonical at connection in E
2
vanish in Cartesian coordinates but do not
vanish in polar coordinates. This unpleasant property of Christoel symbols
is due to the fact that Christoel symbols do not form a tensor.
In particular if a tensor vanishes in some coordinate system, then it van-
ishes in arbitrary coordinate system too. This implies very simple but im-
portant
Proposition If curvature tensor R
i
kmn
vanishes in some coordinate sys-
tem, then it vanishes in arbitrary coordinate systems.
We see that if Christoel symbols vanish in a vicinity of a given point
p in some chosen coordinate system then it Riemannian curvature tensor
vanishes at the point p (see the formula (5.4)) and hence it vanishes at the
point p in arbitrary coordinate system.
85
5.2 Riemann curvature tensor of Riemannian mani-
folds.
Let M be Riemannian manifold equipped with Riemannian metric G
In this section we will consider curvature tensor of Levi-Civita connection
of Riemannian metric G.
The curvature tensor for Levi-Civita connection will be called later Rie-
mann curvature tensor, or Riemann tensor.
Using Riemannian metric one can consider Riemann tensor with all low
indices
R
ikmn
= g
ij
R
j
kmn
(5.7)
Due to identities (5.5) and (5.6) for curvature tensor Riemann tensor
obeys the following identities:
R
ikmn
= R
iknm
, R
ikmn
+ R
imnk
+ R
inkm
= 0 (5.8)
Riemann curvature tensor which is curvature tensor for Levi-Civita con-
nection obeys also the following identities:
R
ikmn
= R
kimn
, R
ikmn
= R
mnki
. (5.9)
These condition lead to the fact that for 2-dimensional Riemannian man-
ifold the Riemann curvature tensor of Levi-Civita connection has essentially
only one non-vanishing component: all components vanish or equal to compo-
nent R
1212
up to a sign. Indeed consider for 2-dimensional Riemannian man-
ifold Riemann tensor R
ikmn
, where i, k, m, n = 1, 2. Since antisymmetricity
with respect to third and fourth indices (R
ikmn
= R
iknm
), R
ik11
= R
ik22
= 0
and R
ik12
= R
ik21
. The same for rst and second indices: since antisym-
metricity with respect to the the rst and second indices (R
12mn
= R
21mn
),
R
11mn
= R
22mn
= 0 and R
12mn
= R
ik21
. If we denote R
1212
= a then
R
1212
= R
2121
= a, R
1221
= R
2112
= a
and all other components vanish.
For Riemann tensor one can consider Ricci tensor,
R
mn
= R
i
min
(5.10)
which is symmetrical tensor: R
mn
= R
nm
.
86
One can consider scalar curvature:
R = R
i
kin
g
kn
= g
kn
R
kn
(5.11)
where g
kn
is Riemannian metric with indices above (the matrix ||g
ik
|| is
inverse to the matrix ||g
il
||).
Ricci tensor and scalar curvature play essential role for formulation of Ein-
stein gravity equations. In particular the space without matter the Einsten
equations have the following form:
R
ik

1
R
g
ik
= 0 . (5.12)
5.3

Curvature of surfaces in E
3
.. Theorema Egregium
again
Express Riemannian curvature of surfaces in E
3
in terms of derivation formulae (5.13).
Consider derivation formulae for the orthonormal basis {e, f , n, } adjusted to the sur-
face M:
d
_
_
e
f
n
_
_
=
_
_
0 a b
a 0 c
b c 0
_
_
_
_
e
f
n
_
_
, (5.13)
where as usual e, f , n vector elds of unit length which are orthogonal to each other
and n is orthogonal to the surface M. As we know the induced connection on the surface
M is dened by the formulae (4.20) and (4.21):

Y
e = (de(Y))
tangent
= a(X)f ,
Y
f = (df (Y))
tangent
= a(X)e , (5.14)
According to the denition of curvature calculate
R(e, f )e =
e

f
e
f

e
e
[e,f ]
e .
Note that since the induced connection is symmetrical connection then:

e
f
f
e [e, f ] = 0 . (5.15)
hence due to (5.14)
[e, f ] =
e
f
f
e = a(e)e +a(f )f (5.16)
Thus we see that R(e, f )e =

f
e
f

e
e
[e,f ]
e =
e
(a(f )f )
f
(a(e)f )
a(e)e+a(f )f
e =

e
a(f )f +a(f )
e
f
f
a(e)f a(e)
f
f a(e)
e
e a(f )
f
e =
87

e
a(f )f a(f )a(e)e
f
a(e)f +a(e)a(f )e a(e)a(e)f a(f )a(f )f =
[
e
a(f )f
f
a(e)f a ((e)e a(f )f )] f = da(e, f )f .
Recall that we established in 4.38 that for Gaussian curvature K
K = b c(e, f ) = da(e, f )
Hence we come to the relation:
R(e, f )e = da(e, f ) = Kf . (5.17)
This means that
R
2
112
= K
(in the basis e, f ), i.e. the scalar curvature
R = 2R
1212
= 2K
We come to fundamental relation which claims that the Gaussian curvature (the magni-
tude dened in terms of External observer) equals to the scalar curvature, the magnitude
dened in terms of Internal observer. This gives us the straightforward proof of Theorema
Egregium.
Below we will give the proof of Theorema Egregium by straightforward calcultaions.
5.4 Relation between Gaussian curvature and Riemann
curvature tensor. Straightforward proof of Theo-
rema Egregium
Let M be a surface in E
3
and R
i
kmp
be Riemann tensor, Riemann curva-
ture tensor of Levi-Civita connection. Recall that this means that R
i
kmp
is curvature tensor of the connection , which is Levi-Civita connection of
the Riemannian metric g

induced on the surface M by standard Euclidean


metric dx
2
+ dy
2
+ dz
2
. Recall that Riemann curvature tensor is expressed
via Christoel symbols of connection by the formula
R
i
kmn
=
m

i
nk
+
i
mp

p
nk

i
mk

i
np

p
mk
(5.18)
(see he formula (5.4)) where Christoel symbols of Levi-Civita connection
are dened by the formula

i
mk
=
1
2
g
ij
_
g
jm
x
k
+
g
jk
x
m

g
mk
x
j
_
(5.19)
(see Levi-Civita Theorem)
88
Recall that scalar curvature R of Riemann tensor equals to R = R
i
kim
g
km
,
where g
km
is Riemannina metric tensor with upper indices (matrix ||g
ik
|| is
inverse to the matrix ||g
ik
||). Note that as it was mentioned before the
formula for scalar curvature becomes very simple in two-dimensional case (see
formulae (5.8) and (5.8) above). Using these formulae calculate Ricci tensor
and scalar curvature R. Let R
1212
= a. We know that all other components
of Riemann tensor equal to zero or equal to a (see (5.8) and (5.8)). One can
show that scalar curvature R can be expressed via the component R
1212
= a
by the formula
R =
2R
1212
det g
(5.20)
where det g = det g
ik
= g
11
g
22
= g
2
12
.
Show it. Using identities (5.8) and (5.8) we see that
R
11
= R
i
1i1
= R
2
121
= g
22
R
2121
+g
21
R
1121
= g
22
R
1212
= g
22
a (5.21)
R
22
= R
i
2i2
= R
1
212
= g
11
R
1212
+g
12
R
2221
= g
11
R
1212
= g
11
a (5.22)
R
12
= R
21
= R
i
1i2
= R
1
112
= g
12
R
2112
= g
12
R
1212
= g
12
a (5.23)
Thus
R
ik
=
_
g
2
2a g
12
a
g
21
a g
11
a
_
Now for scalar curvature we have
R = R
i
kim
g
km
= R
km
g
km
= g
11
R
11
+g
12
R
12
+g
21
R
21
+g
22
R
22
= (5.24)
2R
1212
(g
11
g
22
g
12
g
12
) = 2a det g
1
=
2a
det g
Now we formulate very important
Proposition For an arbitrary point of the surface M
R = 2K (5.25)
where R = R
i
kim
g
km
is scalar curvature and K is Gaussian curvature.
We know also that for surface M the scalar curvature R is expressed
via Riemann curvature tensor by the formula (5.20). Hence if we know
the Gaussian curvature then we know all components of Riemann curvature
tensor (since all components vanish or equal to a.): using (5.20) one can
rewrite the formula (5.25)
R
1212
det g
= K (5.26)
89
This is nothing but Theorema Egregium! Indeed Theorema Egregium
(see beginning of the section 4) immediately follows from this Proposition.
Indeed according to the formulae (5.18) and (5.19) the left hand side of
the relation R = 2K depends only on induced Riemannian metric. Hence
Gaussian curvature K depends only on induced Riemannian metric.
It remains to prove the Proposition. We do it by direct calculations in
the next subsection.
Example Let M = S
2
be sphere of radius R in E
3
. Show that one cannot
nd local coordinates u, v on the sphere such that induced Riemannian metric
equals to du
2
+ dv
2
in these coordinates.
This immediately follows from the Proposition. Indeed suppose there ex-
ist local coordinates u, v on the sphere such that induced Riemannian metric
equals to du
2
+ dv
2
, i.e. Riemannian metric is given by unity matrix. Then
according to the formulae for Levi-Civita connection, the Christoel symbols
equal to zero in these coordinates. Hence Riemann curvature tensor equals
to zero, and scalar curvature too. Due to Proposition this is in contradiction
with the fact that Gaussian curvature of the sphere equals to
1
R
2
.
5.4.1

Proof of the Proposition (5.25)
We prove this fact by direct calculations. The plan of calculations is following:
Let M be a surface in E
3
For an arbitrary point p of the surface M we
consider cartesian coordinates x, y, z such that orgin coincides with the point
p and coordinate plane OXY is the palne attached at the surface at the
point p and the axis OZ is orthogonal to the surface. In these coordinates
calculations become easy. The surface M in these Cartesian coordinates can
be expressed by the equation
_

_
x = u
y = v
z = F(u, v)
(5.27)
where F(u, v) has local extremum at the point u = v = 0
14
.
14
more in detail this is stationary point. It is local extremum if quadratic form corre-
sponding to second dierential is positive or negative denite, i.e. F
uu
F
vv
F
2
uv
keeps the
sign. This means that the Gaussian curvature is not negative
90
Calculate explicitly Gaussian curvature at the point p. In a vicinity of
the point p basis vector r
u
=
_
_
1
0
F
u
_
_
, basis vector r
v
=
_
_
0
1
F
v
_
_
. One can see
that
n(u, v) =
1
1 + F
2
u
+ F
2
v
_
_
F
u
F
v
1
_
_
(5.28)
is unit normal vector. In particular in the origin at the point p, we have
r
u
=
_
_
1
0
0
_
_
, r
v
=
_
_
0
1
0
_
_
, and n =
_
_
0
0
1
_
_
(5.29)
Now calculate shape operator. We need this operator only at the origin,
hence during calculations of all derivatives we have to note that we need
the nal result only at the point u = v = 0. This will essentially simplify
calculations since at the point p derivatives F
u
, F
v
equal to zero.
Sr
u
=
u
n(u, v)|
p
=

u
_
_
1
1 + F
2
u
+ F
2
v
_
_
F
u
F
v
1
_
_
_
_
|
u=v=0
=
_
_
F
uu
F
vu
1
_
_
|
u=v=0
= F
uu
r
u
+ F
uv
r
v
and
Sr
v
=
v
n(u, v)|
p
=

v
_
_
1
1 + F
2
u
+ F
2
v
_
_
F
u
F
v
1
_
_
_
_
|
u=v=0
=
_
_
F
uv
F
vv
1
_
_
|
u=v=0
= F
uv
r
u
+ F
vv
r
v
since F
u
= F
v
= 0. We see that at the origin the Shape (Weingarten)
operator equals to
S =
_
F
uu
F
uv
F
vu
F
vv
_
(5.30)
91
(all the derivatives at the origin).
Gaussian curvature at the point p equals to
K = det S = F
uu
F
vv
F
2
uv
. (5.31)
(all the derivatives at the origin).
Now it is time to calculate the Riemann curvature tensor at the origin.
First of all recall the expression for Riemannian metric for the surface M
in a vicinity of origin is
G =
_
r
u
, r
u
r
u
, r
v

r
v
, r
u
r
v
, r
v

_
=
_
1 + F
2
u
F
u
F
v
F
v
F
u
1 + F
2
v
_
. (5.32)
This immediately follows from the expression for basic vectors r
u
, r
v
Note that Riemannian metric g
ik
at the point u = v = 0 is dened by
unity matrix g
uu
= g
vv
= 1, g
uv
= g
vu
= 0 since p is extremum point:
G =
_
1 + F
2
u
F
u
F
v
F
v
F
u
1 + F
2
v
_

p
=
_
1 0
0 1
_
since p is stationary point, extremum
(F
u
= F
v
= 0). Hence the components of the tensor R
i
kmn
and R
ikmn
=
g
ij
R
j
kmn
at the point p are the same.
Recall that the components of R
i
kmn
are dened by the formula
R
i
kmn
=
m

i
nk
+
i
mp

p
nk

i
mk

i
np

p
mk
.
Notice that at the point p not only the matrix of the metric g
ik
equals to unity
matrix, but more: Christoel symbols vanish at this point in coordinates u, v
since the derivatives of metric at this point vanish. (Why they vanish: this
immediately follows from Levi-Civita formula applied to the metric (5.32),
see also in detail the le The solution of the problem 5 in the coursework,
revisited). Hence to calculate R
i
kmn
at the point p one can consider more
simple formula:
R
i
kmn
|
p
=
m

i
nk
|
p

i
mk
|
p
Try to calculate in a more economical way. Due to Levi-Civita formula

i
mk
=
1
2
g
ij
_
g
jm
x
k
+
g
jk
x
m

g
mk
x
j
_
Since metric g
ik
equals to unity matrix g =
_
1 0
0 1
_
at the point p hence g
ij
is unity matrix also:
g
ik
|
p
=
_
1 0
0 1
_
=
ik
.
92
(We denote
ik
the unity matrix: all diagonal components equal to 1, all other
components equal to zero. (It is so called Kronecker symbols)) Moreover we
know also that all the rst derivatives of the metric vanish at the point p:
g
ik
x
m
|
p
= 0 .
Hence it follows from the formulae above that for an arbitrary indices i, j, k, m, n

x
i
_
g
km
g
pr
x
j
_

p
=
g
km
x
i

p
g
pr
x
j

p
+ g
km

2
g
pr
x
i
x
j

p
=
km

2
g
pr
x
i
x
j

p
.
Now using the Levi-Civita formula for the Christoel symbols of connection:

i
mk
=
1
2
g
ij
_
g
jm
x
k
+
g
jk
x
m

g
mk
x
j
_
we come to
n

i
mk
|
p
=

x
n
_
1
2
g
ij
_
g
jm
x
k
+
g
jk
x
m

g
mk
x
j
__

p
=
1
2

ij
_

2
g
jm
x
n
x
k
+

2
g
jk
x
n
x
m


2
g
mk
x
n
x
j
_

p
. (5.33)
Now using this formula we are ready to calculate Riemann curvature tensor
R
i
kmn
. Remember that it is enough to calculate R
1
212
and R
1
212
= R
1212
at
the point p since g
ik
=
ik
at the point p. We have that at p R
1
212
|
p
=

1
22
|
p

1
12
|
p
and according to (5.33)

1
22
=
1
2

1j
_

2
g
j2
x
1
x
2
+

2
g
j2
x
1
x
2


2
g
22
x
1
x
j
_

p
=
1
2
_

2
g
12
x
1
x
2
+

2
g
12
x
1
x
2


2
g
22
x
1
x
1
_

p
=

2
g
12
x
1
x
2

1
2

2
g
22
x
1
x
1

p
,

1
12
=
1
2

1j
_

2
g
j1
x
2
x
2
+

2
g
j2
x
2
x
1


2
g
12
x
2
x
j
_

p
=
1
2
_

2
g
11
x
2
x
2
+

2
g
12
x
2
x
1


2
g
12
x
2
x
1
_

p
=
1
2

2
g
11
x
2
x
2

p
,
Hence
R
1212
|
p
= R
1
212
|
p
=
1

1
22
|
p

1
12
|
p
=
_

2
g
12
x
1
x
2

1
2

2
g
22
x
1
x
1

1
2

2
g
11
x
2
x
2
_

p
93
Now using expression (5.32) for metric calculate second derivatives

2
g
12
x
1
x
2
,

2
g
11
x
2
x
2
and

2
g
22
x
1
x
1
:

2
g
12
x
1
x
2
|
p
=

2
(F
u
F
v
)
uv

p
=

u
(F
u
F
vv
+ F
uv
F
v
) |
p
=
_
F
uu
F
vv
+ F
2
uv
_
|
p
.
since F
u
= F
v
= 0 at the extremum point p (x
1
= u, x
2
= v). Analogoulsy

2
g
11
x
2
x
2
|
p
=

2
(1 + F
2
u
)
vv

p
=

v
(2F
u
F
uv
) |
p
= 2F
2
uv
|
p
and

2
g
22
x
1
x
1
|
p
=

2
(1 + F
2
v
)
uu

p
=

u
(2F
v
F
uv
) |
p
= 2F
2
uv
|
p
,
We have nally that
R
1212
|
p
= R
1
212
|
p
=
1

1
22
|
p

1
12
|
p
=
_

2
g
12
x
1
x
2

1
2

2
g
22
x
1
x
1

1
2

2
g
11
x
2
x
2
_

p
= F
uu
F
vv
F
2
uv
.
The scalar curvature according to (5.20) equals to R = 2R
1212
= 2(F
uu
F
vv

F
2
uv
). Comparing with the expression K = F
uu
F
vv
F
2
uv
in (5.31) for Gaussian
curvature we come to R = 2K .
5.5 Gauss Bonnet Theorem
Consider the integral of curvature over whole closed surface M. According
to the Theorem above the answer has to be equal to 0 (modulo 2), i.e. 2N
where N is an integer, because this integral is a limit when we consider very
small curve. We come to the formula:

D
Kd = 2N
(Compare this formula with formula (4.6)).
What is the value of integer N?
We present now one remarkable Theorem which answers this question
and prove this Theorem using the formula (4.6).
94
Let M be a closed orientable surface.
15
All these surfaces can be clas-
sied up to a dieomorphism. Namely arbitrary closed oriented surface M
is dieomorphic either to sphere (zero holes), or torus (one hole), or pretzel
(two holes),... Number k of holes is intuitively evident characteristic of the
surface. It is related with very important characteristicEuler characteristic
(M) by the following formula:
(M) = 2(1 g(M)), where g is number of holes (5.34)
Remark What we have called here holes in a surface is often referred
to as handles attached o the sphere, so that the sphere itself does not have
any handles, the torus has one handle, the pretzel has two handles and so
on. The number of handles is also called genus.
Euler characteristic appears in many dierent way. The simplest appear-
ance is the following:
Consider on the surface M an arbitrary set of points (vertices) connected
with edges (graph on the surface) such that surface is divided on polygons
with (curvilinear sides)plaquets. (Map of world)
Denote by P number of plaquets (countries of the map)
Denote by E number of edges (boundaries between countries)
Denote by V number of vertices.
Then it turns out that
P E + V = (M) (5.35)
It does not depend on the graph, it depends only on how much holes has
surface.
E.g. for every graph on M, P E + V = 2 if M is dieomorphic to
sphere. For every graph on M P E+V = 0 if M is dieomorphic to torus.
Now we formulate Gau -Bonnet Theorem.
Let M be closed oriented surface in E
3
.
Let K(p) be Gaussian curvature at any point p of this surface.
15
Closed means compact surface without boundaries. Intuitively orientability means
that one can dene out and inner side of the surface. In terms of normal vectors ori-
entability means that one can dene the continuous eld of normal vectors at all the
points of M. The direction of normal vectors at any point denes outward direction.
Orientable surface is called oriented if the direction of normal vector is chosen.
95
Theorem (Gau -Bonnet) The integral of Gaussian curvature over the
closed compact oriented surface M is equal to 2 multiplied by the Euler
characteristic of the surface M
1
2

M
Kd = (M) = 2(1 number of holes) (5.36)
In particular for the surface M dieomorphic to the sphere (M) = 2,
for the surface dieomorphic to the torus it is equal to 0.
The value of the integral does not change under continuous deformations
of surface! It is integer number (up to the factor ) which characterises
topology of the surface.
E.g. consider surface M which is dieomorphic to the sphere. If it is
sphere of the radius R then curvature is equal to
1
R
2
, area of the sphere is
equal to 4R
2
and left hand side is equal to
4
2
= 2.
If surface M is an arbitrary surface dieomorphic to M then metrics and
curvature depend from point to the point, Gau -Bonnet states that integral
nevertheless remains unchanged.
Very simple but impressive corollary:
Let M be surface dieomorphic to sphere in E
3
. Then there exists at least
one point where Gaussian curvature is positive.
Proof: Suppose it is not right. Then

M
K

det gdudv 0. On the other


hand according to the Theorem it is equal to 4. Contradiction.
Proof of Gau-Bonet Theorem
Consider triangulation of the surface M. Suppose M is covered by N triangles. Then
number of edges will be 3N/over2. If V number of vertices then according to Euler
Theorem
N
3N
2
+V = V
N
2
= (M).
Calculate the sum of the angles of all triangles. On the one hand it is equal to 2V . On
the other hand according the formula (4.6) it is equal to
N

i=1
_
+

i
Kd
_
= N +
N

i=1
_

i
Kd
_
= N +

M
Kd
We see that 2V = N +

M
Kd, i.e.

M
Kd =
_
2V
N
2
_
= 2(M)
96
6 Appendices
6.1

Integrals of motions and geodesics.
We see how useful in Riemannian geometry to use the Lagrangian approach.
To solve and study solutions of Lagrangian equations (in particular geodesics which
are solutions of Euler-Lagrange equations for Lagrangian of free particle) it is very useful
to use integrals of motion
6.1.1

Integral of motion for arbitrary Lagrangian L(x, x)
Let L = L(x, x) be a Lagrangian, the function of point and velocity vectors on manifold
M (the function on tangent bundle TM).
Denition We say that the function F = F(q, q) on TM is integral of motion for
Lagrangian L = L(x, x) if for any curve q = q(t) which is the solution of Euler-Lagrange
equations of motions the magnitude I(t) = F(x(t), x(t)) is preserved along this curve:
F(x(t), x(t)) = const if x(t) is a solution of Euler-Lagrange equations(3.9). (6.1)
In other words
d
dt
(F(x(t), x(t))) = 0 if x
i
(t):
d
dt
_
L

x
i
_

L
x
i
= 0 . (6.2)
6.1.2

Basic examples of Integrals of motion: Generalised momen-
tum and Energy
Let L(x
i
, x
i
) does not depend on the coordinate x
1
. L = L(x
2
, . . . , x
n
, x
1
, x
2
, . . . , x
n
).
Then the function
F
1
(x, x) =
L
x
1
is integral of motion. (In the case if L(x
i
, x
i
) does not depend on the coordinate x
i
. the
function F
i
(x, x) =
L

x
i
will be integral of motion.)
Proof is simple. Check the condition (6.2): Euler-Lagrange equations of motion are:
d
dt
_
L

x
i
_

L
x
i
= 0 (i = 1, 2, . . . , n)
We see that exactly rst equation of motion is
d
dt
_
L

x
1
_
=
d
dt
F
1
(q, q) = 0 since
L
x
1
= 0, .
(if L(x
i
, x
i
) does not depend on the coordinate x
i
then the function F
i
(x, x) =
L
x
1
is
integral of motion since i th equation is exactly the condition

F
i
= 0.)
The integral of motion F
i
=
L

x
i
is called sometimes generalised momentum.
97
Another very important example of integral of motion is: energy.
E(x, x) = x
i
L
x
i
L. (6.3)
One can check by direct calculation that it is indeed integral of motion. Using Euler
Lagrange equations
d
dt
_
L

x
i
_

L
x
i
we have:
d
dt
E(x(t), x(t)) =
d
dt
_
x
i
L
x
i
L
_
=
L
x
i
d x
i
dt
+
d
dt
_
L
x
i
_
x
i

dL
dt
=
L
x
i
d x
i
dt
+
L
x
i
dx
i
dt

dL(x, x)
dt
=
dL(x, x)
dt

dL(x, x)
dt
= 0 .
6.1.3

Integrals of motion for geodesics
Apply the integral of motions for studying geodesics.
The Lagrangian of free particle L
free
=
g
ik
(x) x
i
x
k
2
. For Lagrangian of free particle
solution of Euler-Lagrange equations of motions are geodesics.
If F = F(x, x) is the integral of motion of the free Lagrangian L
free
=
g
ik
(x) x
i
x
k
2
then
the condition (6.1) means that the magnitude I(t) = F(x
i
(t), x
i
(t)) is preserved along the
geodesics:
I(t) = F(x
i
(t), x
i
(t)) = const, i.e.
d
dt
I(t) = 0 if x
i
(t) is geodesic. (6.4)
Consider examples of integrals of motion for free Lagrangian, i.e. magnitudes which
preserve along the geodesics:
Example 1 Note that for an arbitrary free Lagrangian Energy integral (6.3) is an
integral of motion:
E = x
i
L
x
i
L = x
i

_
g
pq
(x) x
p
x
q
2
_
x
i

g
ik
(x) x
i
x
k
2
=
x
i
g
iq
(x) x
q

g
ik
(x) x
i
x
k
2
=
g
ik
(x) x
i
x
k
2
. (6.5)
This is an integral of motion for an arbitrary Riemannian metric. It is preserved on an
arbitrary geodesic
dE(t)
dt
=
d
dt
_
1
2
g
ik
(x(t)) x
i
(t) x
k
(t)
_
= 0 .
In fact we already know this integral of motion: Energy (6.5) is proportional to the square
of the length of velocity vector:
|v| =

g
ik
(x) x
i
x
k
=

2E . (6.6)
We already proved that velocity vector is preserved along the geodesic (see the Proposition
in the subsection 3.2.1 and its proof (??).)
98
Example 2 Consider Riemannian metric G = adu
2
+ bdv
2
(see also calculations in
subsection 2.3.3) in the case if a = a(u), b = b(u), i.e. coecients do not depend on the
second coordinate v:
G = a(u)du
2
+b(u)dv
2
, L
free
=
1
2
_
a(u) u
2
+b(u) v
2
_
(6.7)
We see that Lagrangian does not depend on the second coordinate v hence the magnitude
F =
L
free
v
= b(u) v (6.8)
is preserved along geodesic. It is integral of motion because Euler-Lagrange equation for
coordinate v is
d
dt
L
free
v

L
free
v
=
d
dt
L
free
v
=
d
dt
F = 0 since
L
free
v
= 0. .
In fact all revolution surfaces which we consider here (cylinder, cone, sphere,...) have
Riemannian metric of this type. Indeed consider further examples.
Example (sphere)
Sphere of the radius R in E
3
. Riemannian metric: G = Rd
2
+ R
2
sin
2
d
2
and
L
free
=
1
2
_
R
2

2
+R
2
sin
2

2
_
It is the case (6.7) for u = , v = , b(u) = R
2
sin
2
The
integral of motion is
F =
L
free

= R
2
sin
2

Example (cone)
Consider cone
_

_
x = ahcos
y = ahsin
z = bh
. Riemannian metric:
G = d(ahcos )
2
+d(ahsin )
2
+ (dbh)
2
= (a
2
+b
2
)dh
2
+a
2
h
2
d
2
.
and free Lagrangian
L
free
=
(a
2
+b
2
)

h
2
+a
2
h
2

2
2
.
The integral of motion is
F =
L
free

= a
2
h
2
.
Example (general surface of revolution)
Consider a surface of revolution in E
3
:
r(h, ):
_

_
x = f(h) cos
y = f(h) sin
z = h
(f(h) > 0) (6.9)
99
(In the case f(h) = R it is cylinder, in the case f(h) = kh it is a cone, in the case
f(h) =

R
2
h
2
it is a sphere, in the case f(h) =

R
2
+h
2
it is one-sheeted hyperboloid,
in the case z = cos h it is catenoid,...)
For the surface of revolution (6.9)
G = d(f(h) cos )
2
+d(f(h) sin )
2
+ (dh)
2
= (f

(h) cos dh f(h) sin d)


2
+
(f

(h) sin dh +f(h) cos d)


2
+dh
2
= (1 +f
2
(h))dh
2
+f
2
(h)d
2
.
The free Lagrangian of the surface of revolution is
L
free
=
_
1 +f
2
(h)
_

h
2
+f
2
(h)
2
2
.
and the integral of motion is
F =
L
free

= f
2
(h) .
6.1.4

Using integral of motions to calculate geodesics
Integrals of motions may be very useful to calculate geodesics. The equations for geodesics
are second order dierential equations. If we know integrals of motions they help us to
solve these equations. Consider just an example.
For Lobachevsky plane the free Lagrangian L =
x
2
+ y
2
2y
2
. We already calculated geodesics
in the subsection 3.3.4. Geodesics are solutions of second order Euler-Lagrange equations
for the Lagrangian L =
x
2
+ y
2
2y
2
(see the subsection 3.3.4)
_

x
2 x y
y
= 0

y +
x
2
y

y
2
y
= 0
It is not so easy to solve these dierential equations.
For Lobachevsky plane we know two integrals of motions:
E = L =
x
2
+ y
2
2y
2
, and F =
L
x
=
x
y
2
. (6.10)
These both integrals preserve in time: if x(t), y(t) is geodesics then
_
F =
x(t)
y(t)
2
E =
x(t)
2
+ y(t)
2
2y(t)
2
= C
2

_
x = C
1
y
2
y =

2C
2
y
2
C
2
1
y
4
These are rst order dierential equations. It is much easier to solve these equations in
general case than initial second order dierential equations.
100
6.2 Induced metric on surfaces.
Recall here again induced metric (see for detail subsection 1.4)
If surface M: r = r(u, v)is embedded in E
3
then induced Riemannian metric G
M
is
dened by the formulae
X, Y = G
M
(X, Y) = G(X, Y) , (6.11)
where G is Euclidean metric in E
3
:
G
M
= dx
2
+dy
2
+dz
2

r=r(u,v)
=
3

i=1
(dx
i
)
2

r=r(u,v)
=
3

i=1
_
x
i
u

du

_
2
=
x
i
u

x
i
u

du

du

,
i.e.
G
M
= g

du

, where g

=
x
i
u

x
i
u

du

du

.
We use notations x, y, z or x
i
(i = 1, 2, 3) for Cartesian coordinates in E
3
, u, v or u

( = 1, 2) for coordinates on the surface. We usually omit summation symbol over dummy
indices. For coordinate tangent vectors

..
Internal observer
= r

=
x
i
u

x
i
. .
External observer
We have already plenty examples in the subsection 1.4. In particular for scalar product
g

,

u

= x
i
x
i
.r

, r

. (6.12)
6.2.1 Recalling Weingarten operator
Continue to play with formulae
16
.
Recall the Weingarten (shape) operator which acts on tangent vectors:
SX =
X
n, (6.13)
where we denote by n-unit normal vector eld at the points of the surface M: n, r

= 0,
n, r

= 1.
Now we realise that the derivative
X
R of vector eld with respect to another vector
eld is not a well-dened object: we need a connection. The formula
X
R in Cartesian
coordinates, is nothing but the derivative with respect to at canonical connection: If
16
In some sense dierential geometry it is when we write down the formulae expressing
the geometrical facts, dierentiate these formulae then reveal the geometrical meaning of
the new obtained formulae e.t.c.
101
we work only in Cartesian coordinates we do not need to distinguish between
X
R and

can.at
X
R. Sometimes with some abuse of notations we will use
X
R instead
can.at
X
R,
but never forget: this can be done only in Cartesian coordinates where Christoel symbols
of at canonical connection vanish:

X
R =
can.at
X
R in Cartesian coordinates .
So the rigorous denition of Weingarten operator is
SX =
can.at
X
n, (6.14)
but we often use the former one (equation (6.13)) just remembering that this can be done
only in Cartesian coordinates.
Recall that the fact that Weingarten operator S maps tangent vectors to tangent
vectors follows from the property: n, X = 0 X is tangent to the surface.
Indeed:
0 =
X
n, n = 2
X
n, n = 2SX, n = 0 SX is tangent to the surface
Recall also that normal unit vector is dened up to a sign, n n. On the other hand
if n is chosen then S is dened uniquely.
6.2.2 Second quadratic form
We dene now the new object: second quadratic form
Denition. For two tangent vectors X, Y A(X, Y) is dened such that
_

can.at
X
Y
_

= A(X, Y)n (6.15)


i.e. we take orthogonal component of the derivative of Y with respect to X.
This denition seems to be very vague: to evaluate covariant derivative we have to
consider not a vector Y at a given point but the vector eld. In fact one can see that
A(X, Y) does depend only on the value of Y at the given point.
Indeed it follows from the denition of second quadratic form and from the properties
of Weingarten operator that
A(X, Y) =
_

can.at
X
Y
_

, n

can.at
X
Y, n

X
Y, n

Y,
can.at
X
n

= S(X), Y (6.16)
We proved that second quadratic form depends only on value of vector eld Y at the
given poit and we established the relation between second quadratic form and Weingarten
operator.
Proposition The second quadratic form A(X, Y) is symmetric bilinear form on tan-
gent vectors X, Y in a given point.
A: T
p
M T
p
M R, A(X, Y) = A(Y, X) = SX, Y . (6.17)
102
In components
A = A

du

du

= r

, n =

2
x
i
u

n
i
. (6.18)
and
S

= g

= g

x
i

n
i
, (6.19)
i.e.
A = GS, S = G
1
A.
Remark The normal unit vector eld is dened up to a sign.
6.2.3 Gaussian and mean curvatures
Recall that Gaussian curvature
K = det S
and mean curvature
H = Tr S
It is easy to see that for Gaussian curvature
K = det S = det(G
1
A) =
det A
det G
.
We know already the geometrical meaning of Gaussian and mean curvatures from the
point of view of the External Observer:
Gaussian curvature K equals to the product of principal curvatures, and mean cur-
vartures equals to the sum of principal curvatures.
Now we ask a principal question: what bout internal observer, aunt living on the
surface?
We will show that Gaussian curvature can be expressed in terms of induced Rieman-
nian metric, i.e. it is an internal characteristic of the surface, invariant of isometries.
It is not the case with mean curvature: cylinder is isometric to the plane but it have
non-zero mean curvature.
6.2.4 Examples of calculation of Weingarten operator, Second
quadratic forms, curvatures for cylinder, cone and sphere.
Cylinder
We already calculated induced Riemannian metric on the cylinder (see (1.56)).
Cylinder is given by the equation x
2
+ y
2
= R
2
. One can consider the following
parameterisation of this surface:
r(h, ):
_

_
x = Rcos
y = Rsin
z = h
, r
h
=
_
_
0
0
1
_
_
, r

=
_
_
Rsin
Rcos
0
_
_
, (6.20)
103
G
cylinder
=
_
dx
2
+dy
2
+dz
2
_

x=a cos ,y=a sin ,z=h


=
= (a sin d)
2
+ (a cos d)
2
+dh
2
= a
2
d
2
+dh
2
, ||g

|| =
_
1 0
0 R
2
_
.
Normal unit vector n =
_
_
cos
sin
0
_
_
. Choose n =
_
_
cos
sin
0
_
_
. Weingarten operator
S
h
=
can.at
r
h
n =
r
h
n =
h
_
_
cos
sin
0
_
_
= 0 ,
S

=
can.at
r

n =
r

n =

_
_
cos
sin
0
_
_
=
_
_
sin
cos
0
_
_
=

R
.
S
_

_
=
_
0

R
_
, S =
_
0 0
0
1
R
_
. (6.21)
Calculate second quadratic form: r
hh
=
h
r
h
=
_
_
0
0
0
_
_
, r
h
= r
h
=

h
_
_
Rsin
Rcos
0
_
_
= 0, r

_
_
Rsin
Rcos
0
_
_
=
_
_
Rcos
Rsin
0
_
_
= Rn.
We have
A

= r

, n, A =
_
r
hh
, n r
h
, n
r
h
, n r

, n
_
=
_
0 0
0 R
_
, (6.22)
A = GS =
_
1 0
0 R
2
__
0 0
0
1
R
_
=
_
0 0
0 R
_
,
For Gaussian and mean curvatures we have
K = det S =
det A
det G
= det
_
0 0
0
1
R
_
= 0 , (6.23)
and mean curvature
H = Tr S = Tr
_
0 0
0
1
R
_
=
1
R
. (6.24)
Mean curvature is dene up to a sign. If we change n n mean curvature H
1
R
and
Gaussian curvature will not change.
Cone
We already calculated induced Riemannian metric on the cone (see (1.59).
104
Cone is given by the equation x
2
+ y
2
k
2
z
2
= 0. One can consider the following
parameterisation of this surface:
r(h, ):
_

_
x = khcos
y = khsin
z = h
, r
h
=
_
_
k cos
k sin
1
_
_
, r

=
_
_
khsin
khcos
0
_
_
, (6.25)
G
cone
=
_
dx
2
+dy
2
+dz
2
_

x=khcos ,y=khsin ,z=h


=
= (khsin d +k cos dh)
2
+ (khcos d +k sin dh)
2
+dh
2
=
k
2
h
2
d
2
+ (k
2
+ 1)dh
2
, ||g

|| =
_
k
2
+ 1 0
0 k
2
h
2
_
.
One can see that N =
_
_
cos
sin
k
_
_
is orthogonal to the surface: Nr
h
, r

. Hence normal
unit vector n =
1

1+k
2
_
_
cos
sin
k
_
_
. Choose n =
1

1+k
2
_
_
cos
sin
k
_
_
. Weingarten operator
S
h
=
can.at
r
h
n =
r
h
n =
h
_
_
1

1 +k
2
_
_
cos
sin
k
_
_
_
_
= 0 ,
S

=
can.at
r

n =
r

n =

_
_
1

1 +k
2
_
_
cos
sin
k
_
_
_
_
=
1

1 +k
2
_
_
sin
cos
0
_
_
=

kh

k
2
+ 1
.
S
_

_
=
_
0

kh

k
2
+1
_
, S =
_
0 0
0
1
kh

k
2
+1
_
. (6.26)
Calculate second quadratic form: r
hh
=
h
r
h
=
_
_
0
0
0
_
_
, r
h
= r
h
=

h
_
_
khsin
khcos
0
_
_
=
_
_
k sin
k cos
0
_
_
, r

_
_
khsin
khcos
0
_
_
=
_
_
khcos
khsin
0
_
_
.
We have
A

= r

, n, A =
_
r
hh
, n r
h
, n
r
h
, n r

, n
_
=
_
0 0
0
kh

1+k
2
_
, (6.27)
105
A = GS =
_
k
2
+ 1 0
0 k
2
h
2
__
0 0
0
1
kh

k
2
+1
_
=
_
0 0
0
kh

k
2
+1
_
,
For Gaussian and mean curvatures we have
K = det S =
det A
det G
= det
_
0 0
0
1
kh

k
2
+1
_
= 0 , (6.28)
and mean curvature
H = Tr S = Tr
_
0 0
0
1
kh

k
2
+1
_
=
1
kh

k
2
+ 1
. (6.29)
Mean curvature is dene up to a sign. If we change n n mean curvature H
1
R
and
Gaussian curvature will not change.
Sphere
Sphere is given by the equation x
2
+ y
2
+ z
2
= a
2
. Consider the parameterisation of
sphere in spherical coordinates
r(, ):
_

_
x = Rsin cos
y = Rsin sin
z = Rcos
(6.30)
We already calculated induced Riemannian metric on the sphere (see (6.2.4)). Recall
that
r

=
_
_
Rcos cos
Rcos sin
Rsin
_
_
, r

=
_
_
Rsin sin
Rsin cos
0
_
_
and
G
S
2 =
_
dx
2
+dy
2
+dz
2
_

x=Rsin cos ,y=Rsin sin ,z=Rcos


=
(Rcos cos d Rsin sin d)
2
+ (Rcos sin d +Rsin cos d)
2
+
(Rsin d)
2
= R
2
cos
2
d
2
+R
2
sin
2
d
2
+R
2
sin
2
d
2
=
= R
2
d
2
+R
2
sin
2
d
2
, ||g

|| =
_
R
2
0
0 R
2
sin
2

_
.
For the sphere r(, ) is orthogonal to the surface. Hence normal unit vector n(, ) =

r(,)
R
=
_
_
sin cos
sin sin
cos
_
_
. Choose n =
r
R
=
_
_
sin cos
sin sin
cos
_
_
. Weingarten operator
S

=
can.at
r

n =

n =

_
r
R
_
=
r

R
,
S

=
can.at
r

n =

n =

_
r
R
_
=
r

R
.
106
S
_

_
=
_

R
_
, S =
_
1
R
0
0
1
R
_
. (6.31)
For second quadratic form: r

=
_
_
Rsin cos
Rsin sin
Rcos
_
_
, r

= r

_
_
Rsin sin
Rsin cos
0
_
_
=
_
_
Rcos sin
Rcos cos
0
_
_
, r

_
_
Rsin sin
Rsin cos
0
_
_
=
_
_
Rsin cos
Rsin sin
0
_
_
.
We have
A

= r

, n, A =
_
r

, n r

, n
r

, n r

, n
_
=
_
R 0
0 Rsin
2

_
, (6.32)
A = GS =
_
R
2
0
0 R
2
sin
2

__
1
R
0
0
1
R
_
= R
_
1 0
0 sin
2

_
,
For Gaussian and mean curvatures we have
K = det S =
det A
det G
= det
_

1
R
0
0
1
R
_
=
1
R
2
, (6.33)
and mean curvature
H = Tr S = Tr
_

1
R
0
0
1
R
_
=
2
R
, (6.34)
Mean curvature is dene up to a sign. If we change n n mean curvature H
1
R
and
Gaussian curvature will not change.
We see that for the sphere Gaussian curvature is not equal to zero, whilst for cylinder
and cone Gaussian curvature equals to zero.
107

Você também pode gostar