You are on page 1of 129

Riemannian Geometry

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


Manchester, 8 May 2016

Contents
1 Riemannian manifolds
1.1 Manifolds. Tensors. (Recalling) . . . . . . . . . . . . . . . .
1.2 Riemannian manifold manifold equipped with Riemannian
metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.1 Pseudoriemannian manifold . . . . . . . . . . . . .
1.3 Scalar product, conformally Euclidean metric. Length of tangent vectors and angle between them. Length of curves . . .
1.3.1 Conformally Euclidean metric . . . . . . . . . . . . .
1.3.2 Length of curves . . . . . . . . . . . . . . . . . . . .
1.4 Riemannian structure on the surfaces embedded in Euclidean
space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.4.1 Internal and external coordinates of tangent vector .
1.4.2 Explicit formulae for induced Riemannian metric (First
Quadratic form) . . . . . . . . . . . . . . . . . . . . .
1.4.3 Induced Riemannian metrics. Examples. . . . . . . .
1.4.4 Induced metric on two-sheeted hyperboloid embedded
in pseudo-Euclidean space. . . . . . . . . . . . . . . .
1.5 Isometries of Riemannian manifolds. . . . . . . . . . . . . .
1.5.1 Isometries of Riemannian manifold (on itself) . . . .
1.6 Infinitesimal isometries of Riemannian manifold (Killing vector fields) . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.6.1 Locally Euclidean Riemannian manifolds . . . . . .
1.6.2 Conformally Euclidean Riemannian manifolds . . . .
1

1
1

.
.

5
9

. 10
. 11
. 12
. 14
. 15
. 16
. 19
. 26
. 28
. 29
. 30
. 32
. 34

1.7

Volume element in Riemannian manifold . . . . . . . . . . . .


1.7.1 Motivation: Gramm formula for volume of parallelepiped
1.7.2 Invariance of volume element under changing of coordinates . . . . . . . . . . . . . . . . . . . . . . . . .
1.7.3 Examples of calculating volume element . . . . . . . .

2 Covariant differentiaion. Connection. Levi Civita Connection on Riemannian manifold


2.1 Differentiation of vector field along the vector field.Affine
connection . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1 Definition of connection. Christoffel symbols of connection . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.2 Transformation of Christoffel symbols for an arbitrary
connection . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.3 Canonical flat affine connection . . . . . . . . . . . . .
2.1.4 Global aspects of existence of connection . . . . . . .
2.2 Connection induced on the surfaces . . . . . . . . . . . . . . .
2.2.1 Calculation of induced connection on surfaces in E3 . . .
2.3 Levi-Civita connection . . . . . . . . . . . . . . . . . . . . . .
2.3.1 Symmetric connection . . . . . . . . . . . . . . . . . .
2.3.2 Levi-Civita connection. Theorem and Explicit formulae
2.3.3 Levi-Civita connection of En . . . . . . . . . . . . . . .
2.3.4 Levi-Civita connection on 2-dimensional Riemannian
manifold with metric G = adu2 + bdv 2 . . . . . . . . . .
2.3.5 Example of the sphere again . . . . . . . . . . . . . . .
2.4 Levi-Civita connection = induced connection on surfaces in E3
3 Parallel transport and geodesics
3.1 Parallel transport . . . . . . . . . . . . . . . . . . . . . . . .
3.1.1 Definition . . . . . . . . . . . . . . . . . . . . . . . .
3.1.2 Parallel transport is a linear map. Parallel transport
with respect to Levi-Civita connection . . . . . . . .
3.2 Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.1 Definition. Geodesic on Riemannian manifold . . . .
3.2.2 Geodesics and Lagrangians of free particle on Riemannian manifold. . . . . . . . . . . . . . . . . . . .
3.2.3 Calculations of Christoffel symbols and geodesics using the Lagrangians of a free particle. . . . . . . . . .
2

35
35
36
38

40
40
41
43
44
47
48
48
50
50
51
52
53
53
54

55
. 55
. 55
. 57
. 58
. 58
. 60
. 62

3.2.4

Magnitudes preserved along geodesicsIntegrals of


motion . . . . . . . . . . . . . . . . . . . . . . . . . .
3.2.5 Variational principe and Euler-Lagrange equations
3.2.6 Un-parameterised geodesic . . . . . . . . . . . . . . .
3.2.7 Parallel transport of vectors along geodesics . . . . .
3.2.8 Geodesics on surfaces in E3 . . . . . . . . . . . . . .
3.2.9 Geodesics and shortest distance. . . . . . . . . . . .
3.2.10 Again geodesics for sphere and Lobachevsky plane .

.
.
.
.
.
.
.

64
67
68
69
69
71
73

4 Surfaces in E3
4.1 Parallel transport of the vector. Formulation of result. . . . .
4.1.1 Theorem of parallel transport over closed curve . . . .
4.1.2 Gau Theorema Egregium . . . . . . . . . . . . . . . .
4.1.3 Formula for Gaussian curvature in isothermal (conformal) coordinates . . . . . . . . . . . . . . . . . . . . .
4.2 Derivation formula . . . . . . . . . . . . . . . . . . . . . . . .
4.2.1 Gauss condition (structure equations) . . . . . . . . . .
4.2.2 Geometrical meaning of derivation formula. Weingarten
operator (shape oeprator) in terms of derivation formula. . . . . . . . . . . . . . . . . . . . . . . . . . . .
4.2.3 Gaussian and mean curvature in terms of derivation
formula . . . . . . . . . . . . . . . . . . . . . . . . . .
4.3 Calculations with use of derivation formulae . . . . . . . . . .
4.3.1 Examples of calculations of derivation formulae and
curvatures for cylinder, cone and sphere . . . . . . . .
4.3.2 Curvatures for surface z = F (x, y) . . . . . . . . . . . .
4.3.3 Proof of the Theorem on curvature of surfaces given
in conformal coordinates using derivation formulae . .

75
75
75
78

5 Curvature tensor
5.1 Curvature tensor for connection . . . . . . . . . . . . . . . . .
5.1.1 Properties of curvature tensor . . . . . . . . . . . . . .
5.2 Riemann curvature tensor of Riemannian manifolds. . . . . . .
5.2.1 Curvature of surfaces in E3 .. Theorema Egregium again
5.2.2 Relation between Gaussian curvature and Riemann
curvature tensor and Theorema Egregium . . . . . . .
5.2.3 Straightforward proof of the Proposition (5.27) . . . .
5.3 Gauss Bonnet Theorem . . . . . . . . . . . . . . . . . . . . .

97
97
98
99
100

78
80
83

84
86
86
86
91
94

102
104
108

6 Appendices
6.1 Integrals of motions and geodesics. . . . . . . . . . . . . . .
6.1.1 Integral of motion for arbitrary Lagrangian L(x, x)
.

6.1.2
Basic examples of Integrals of motion: Generalised
momentum and Energy . . . . . . . . . . . . . . . . .
6.1.3 Integrals of motion for geodesics . . . . . . . . . . .
6.1.4 Using integral of motions to calculate geodesics . . .
6.1.5 Killing vectors of Lobachevsky plane and geodesics .
6.2 Induced metric on surfaces. . . . . . . . . . . . . . . . . . .
6.2.1 Recalling Weingarten operator . . . . . . . . . . . . .
6.2.2 Recalling Weingarten operator . . . . . . . . . . . . .
6.2.3 Second quadratic form . . . . . . . . . . . . . . . . .
6.2.4 Gaussian and mean curvatures . . . . . . . . . . . . .
6.2.5 Examples of calculation of Weingarten operator, Second quadratic forms, curvatures for cylinder, cone and
sphere. . . . . . . . . . . . . . . . . . . . . . . . . . .

6.2.6
Proof of the Theorem of parallel transport along
closed curve . . . . . . . . . . . . . . . . . . . . . . .

110
. 110
. 110
.
.
.
.
.
.
.
.
.

111
111
113
114
116
116
117
118
119

. 119
. 123

1
1.1

Riemannian manifolds
Manifolds. Tensors. (Recalling)

I recall briefly basics of manifolds and tensor fields on manifolds.


An n-dimensional manifold is a space such that in a vicinity of any point
one can consider local coordinates {x1 , . . . , xn } (charts). One can consider
0
0
different local coordinates. If both coordinates {x1 , . . . , xn }, {x1 , . . . , xn }
are defined in a vicinity of the given point then they are related by bijective
transition functions (functions defined on domains in Rn and taking values
in Rn ).
0
0

x1 = x1 (x1 , . . . , xn )

20
20
1
n

x = x (x , . . . , x )
(1.1)
...

0
0

xn1 = xn1 (x1 , . . . , xn )

xn0 = xn0 (x1 , . . . , xn )


We say that n-dimensional manifold is differentiable or smooth if transition
functions are diffeomorphisms, i.e. they are smooth and rank of Jacobian is
equal to k, i.e.
10
0
0
x1
x1
x
.
.
.
1
2
n
x 0
x 0
x 0
x
2
2
x2

x1 x
.
.
.
2
x
xn 6= 0 .
(1.2)
det

...
0
0
n0
xn
xn
. . . x
x1
x2
xn
A good example of manifold is an open domain D in n-dimensional vector
space Rn . Cartesian coordinates on Rn define global coordinates on D. On
the other hand one can consider an arbitrary local coordinates in different
domains in Rn . E.g. one can consider polar coordinates {r, } in a domain
D = {x, y : y > 0} of R2 defined by standard formulae:
(
x = r cos
,
(1.3)
y = r sin
det

x
r
y
r


cos r sin
= det
=r
sin r cos

(1.4)

or one can consider spherical coordinates {r, , } in a domain D = {x, y, z : x >


0, y > 0, z > 0} of R3 (or in other domain of R3 ) defined by standard formulae

x = r sin cos
,
y = r sin sin

z = r cos
,

x
x
x
sin cos r cos cos r sin sin
r

y
y
sin sin r cos sin r sin cos = r2 sin
det y
r

= det
z
z
z
cos
r sin
0
r

(1.5)
Choosing domain where polar (spherical) coordinates are well-defined we
have to be award that coordinates have to be well-defined and transition
functions (1.1) have to obey condition (1.2), i.e. they have to be diffeomorphisms. E.g. for domain D in example (1.3) Jacobian (1.4) does not vanish
if and only if r > 0 in D.
Examples of manifolds: Rn , Circle S 1 , Sphere S 2 , in general sphere S n ,
torus S 1 S 1 , cylinder, cone, . . . .
Example Consider in detail circle S 1 . Suppose it is given as x2 + y 2 = 1.
One can consider two local polar coordinates: 1) , 0 < < 2 which covers
all points except the point (1, 0) and 0 : < 0 < which covers all points
except point (1, 0). In a vicinity of any point M on the circle (except these
two exceptional points) one can consider both coordinates and 0
We come to another very useful coordinates on a circle using stereographic
projection. Take north pole of the circle: the point N = (0, 1). Assign to
every point M = (x, y) on the circle the point (t, 0) on the x-axis such that
the point (t, 0), the point M and the north pole N are on the one line. This
can be done for every point of circle except the north pole (0, 1) itself. We
come to stereographic projection of circle S 1 without North pole on the line
R. In the same way we can define stereographic projection of circle without
south pole (the point (0, 1)) on the x-axis. We come to coordinate t0 . One
can see that these coordinates are related by the following simple formula:
t0 =

1
,
t

One very important property of stereographic projetion which we do not use in


this course but it is too beautiful not to mention it: under stereographic projection

all points on the circle x2 + y 2 = 1 with rational coordinates x and y and only
these points transform to rational points on line. Thus we come to Pythagorean
triples a2 + b2 = c2 .

Tensors on Manifold
Recall briefly what are tensors on manifold. For every point p on manifold
M one can consider tangent vector space Tp M the space of vectors tangent
to the manifold at the point M .
Tangent vector A(x) = Ai (x) x i . Under changing of coordinates it transforms as follows:
0

xm (x)

m0
A = A (x) i = Ai (x)
(x0 (x)) m0 .
0 = A
i
m
x
x x
x
i

Hence
0

xi (x) i
A (x) .
(1.6)
A (x ) =
xi
Consider also cotangent space Tp 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.:
One-form (covector) = i (x)dxi transforms as follows
i0

= m (x)dxm = m

xm (x0 ) m0
0
dx = m0 (x0 )dxm .
0
m
x

Hence

xm (x0 )
m (x) .
xm0
Differential form sometimes is called covector.
Tensors:
One can consider contravariant tensors of the rank p
m0 (x0 ) =

T = T i1 i2 ...ip (x)

(1.7)

xi1
xi2
xik

with components {T i1 i2 ...ik }(x). Under changing of coordinates (x1 , . . . , xn )


0
0
(x1 , . . . , xn ) (see (1.1)) they transform as follows:
0

0 0

T i1 i2 ...ip (x0 ) =

xi1 xi2
xip i1 i2 ...ip
.
.
.
T
(x) .
xi1 xi2
xip
3

(1.8)

One can consider covariant tensors of the rank q


S = Sj1 j2 ...jq dxj1 dxj2 . . . dxjq
with components {Sj1 j2 ...jq }. Under changing of coordinates (x1 , . . . , xn )
0
0
(x1 , . . . , xn ) they transform as follows:
Sj10 j20 ...jq0 (x0 ) =

xi1 xi2
xip
.
.
.
0
0
0 Sj j ...j (x) .
xi1 xi2
xip 1 2 q

(1.9)

One can also consider mixed tensors:


i i ...i

Q = Qj11 j22 ...jpq


with components

i2 i dxj1 dxj2 . . . dxjq


i
1
x
x
x k

i i ...i
{Qj11 j22 ...jpq }.

 
p
We call these tensors tensors of the type
.
q

 
p
Tensors of the type
are called contravariant tensors of the rank p. They
0
have p upper indices.  
0
Tensors of the type
are called covariant tensors of the rank q. They
q
have q lower indices.
Having in mind (1.6), (1.7),(1.8) and (1.9) we come to the rule of transformation
for tensors which have p upper and q lover indices, tensors of type


p
:
q
i0 i0 ...i0
Qj10 j20 ...jpq0 (x0 )
1 2

xi1 xi2
xjq i1 i2 ...ip
xip xj1 xj2
=
.
.
.
(x) .
.
.
.
0
0
0 Q
xi1 xi2
xip xj1 xj2
xjq j1 j2 ...jq

E.g. if Sik is a covariant tensor of rank 2 then


Si0 k0 (x0 ) =
If

Aik

xi (x0 ) xk (x0 )
Sik (x) .
xi0
xk0

 
1
is a tensor of rank
(linear operator on Tp M ) then
1
0

0
Aik0 (x0 )

xi (x) xk (x0 ) i
=
Ak (x) .
xi
xk0
4

(1.10)

If

m
Sik

 
1
is a tensor of the type
) then
2
0

0
Sim0 k0

xm xi xk m
=
S (x) .
xm xi0 xk0 ik

(1.11)

Transformations formulae (1.6)(1.11) define vectors, covectors and in


generally any tensor fields in components. E.g. covariant tensor (covariant tensor field) of the rank 2 can be defined as matrix Sik (matrix valued
function Sik (x)) such that under changing of coordinates {x1 , x2 , . . . , xn } 7
0
0
0
{x1 , x2 , . . . , xn }, (1.1) Sik change by the rule (1.10).
Remark Einstein summation rules
In our lectures we always use so called Einstein summation convention. it
implies that when an index occurs twice in the same expression in upper and
in lower postitions, then the expression is implicitly summed over all possible
values for that index. Sometimes it is called dummy indices summation rule.

1.2

Riemannian manifold manifold equipped with


Riemannian metric

Definition The Riemannian manifold is a manifold equipped with a Riemannian metric.


The Riemannian metric on the manifold M defines the length of the
tangent vectors and the length of the curves.
Definition Riemannian metric G on n-dimensional manifold M n defines
for every point p M the scalar product of tangent vectors in the tangent
space Tp M smoothly depending on the point p.
It means that in every coordinate system (x1 , . . . , xn ) a metric G =
gik dxi dxk is defined by a matrix valued smooth function gik (x) (i = 1, . . . , n; k =
1, . . . n) such that for any two vectors
A = Ai (x)

i
,
B
=
B
(x)
,
xi
xi

tangent to the manifold M at the point p with coordinates x = (x1 , x2 , . . . , xn )


(A, B Tp M ) the scalar product is equal to:


hA, BiG p = G(A, B) p = Ai (x)gik (x)B k (x) =


B1
g11 (x) . . . g1n (x)



...
...
1
n ...


A ...A

gn1 (x) . . . gnn (x)



Bn

(1.12)

where
G(A, B) = G(B, A), i.e. gik (x) = gki (x) (symmetricity condition)
G(A, A) > 0 if A 6= 0, i.e.
gik (x)ui uk 0, gik (x)ui uk = 0 iff u1 = = un = 0 (positivedefiniteness)

G(A, B) p=x , i.e. gik (x) are smooth functions.
The matrix ||gik || of components of the metric G we also sometimes denote
by G.
Now we establish rule of transformation for entries of matrix gik (x), of
metric G.
Notice that an arbitrary matrix entry gik is nothing but scalar product
of vectors i , k at the given point:



, k , in coordinates (x1 , . . . , xn )
(1.13)
gik (x) =
i
x x
Use this formula for establishing rule of transformations of gik (x). In the new
0
0
0
coordinates xi = (x1 , . . . , xn ) according this formula we have that



0
,
, in coordinates (x1 , . . . , xn ) .
gi0 k0 (x ) =
xi0 xk0
Now using chain rule, linearity of scalar product and formula (??) we see
that

  i


x xk
0
,
=
,
gi0 k0 (x ) =
xi0 xk0
xi0 xi xk0 xk



xi
xk
,
=
=
xi0 xi xk xk0
|
{z
}
gik (x)

xi
xk
g
(x)
(1.14)
ik
xi0
xk0
This transformation law justifies that gik entries of matrix ||gik || are components of covariant tensor field G = gik dxi dxk of rank 2(see equation (1.10)).
One can say that Riemannian metric is defined by symmetric covariant
smooth tensor field G of the rank 2 which defines scalar product in the tangent
spaces Tp M smoothly depending on the point p. Components of tensor field
G in coordinate system are matrix valued functions gik (x):
G = gik (x)dxi dxk .

(1.15)

In practice it is more convenient to perform transformation of metric G under


changing of coordinates in the following way:
  k

 i
x
x
i
k
i0
k0
dx
dx
=
G = gik dx dx = gik
xi0
xk0
xk i0
xi
0
0
k0
g
= gi0 k0 dxi dxk
0 ik
0 dx dx
i
k
x
x
Hence

xk
xi
g
.
(1.16)
ik
xi0 xk0
We come to transfomation rule (1.14).
Later by some abuse of notations we sometimes omit the sign of tensor
product and write a metric just as
gi0 k0 =

G = gik (x)dxi dxk .


Examples
Rn with canonical coordinates {xi } and with metric
G = (dx1 )2 + (dx2 )2 + + (dxn )2
G = ||gik || = diag [1, 1, . . . , 1]
Recall that this is a basis example of n-dimensional Euclidean space,
where scalar product is defined by the formula:
G(X, Y) = hX, Yi = gik X i Y k = X 1 Y 1 + X 2 Y 2 + + X n Y n .
7

In the general case if G = ||gik || is an arbitrary symmetric positivedefinite metric then G(X, Y) = hX, Yi = gik X i Y k . One can show that
there exists a new basis {ei } such that in this basis G(ei , ek ) = ik . This
basis is called orthonormal basis. (See the Lecture notes in Geometry)
Scalar product in vector space defines the same scalar product at all
the points. In general case for Riemannian manifold scalar product
depends on a point. In Riemannian manifold we consider arbitrary
transformations from local coordinates to new local coordinates.
R2 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 coordinates the Riemannian metric G = dx2 + dy 2 will have the following
appearance:
G = (dx)2 +(dy)2 = (cos drr sin d)2 +(sin dr+r cos d)2 = dr2 +r2 (d)2
We see that for matrix G = ||gik ||

 

gxx gxy
1 0
,
G=
=
gyx gyy
0 1
{z
}
|
in Cartesian coordinates

 

grr gr
1 0
G=
=
gr g
0 r
|
{z
}
in polar coordinates

Circle
Interval [0, 2) in the line 0 x < 2 with Riemannian metric
G = a2 dx2

(1.17)

Renaming x 7 we come to habitual formula for metric for circle of


the radius a: x2 + y 2 = a2 embedded in the Euclidean space E2 :
(
x = a cos
G = a2 d2
, 0 < 2,
(1.18)
y = a sin
Domain in R2 with metric G = du2 + u2 dv 2
(Compare with R2 with polar coordinates).

Cylinder surface
Domain in R2 D = {(x, y) : , 0 x < 2 with Riemannian metric
G = a2 dx2 + dy 2

(1.19)

We see that renaming variables x 7 , y 7 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 E3 :

x = a cos
2
2
2
G = a d + dh
y = a sin , 0 < 2, < h <

z=h
(1.20)
Sphere
Domain in R2 , 0 < x < 2, 0 < y < with metric G = dy 2 + sin2 ydx2
We see that renaming variables x 7 , y 7 h we come to habitual,
familiar formulae for metric in standard spherical coordinates for sphere
x2 + y 2 + z 2 = a2 of the radius a embedded in the Euclidean space E3 :
(
x = a sin cos
G = a2 d2 +a2 sin2 d2
, 0 < 2, < h <
y = a sin sin z = a cos
(1.21)
(See examples also in the Homeworks.)
1.2.1

Pseudoriemannian manifold

If we omit the condition of positive-definiteness for Riemannian metric we


come to so called Pseudoriemannian metric. Manifold equipped with pseudoriemannian metric is called pseudoriemannian manifold. Pseudoriemannian manifolds appear in applications in the special and general relativity
theory.
In pseudoriemanninan space scalar product (X, X) may take an arbitrary
real values: it can be positive, negative, it can be equal to zero. Vectors X
such that (X, X) = 0 are called null-vectors. (See the problem 6 in Homework
1).

Example Consider n + 1-dimensional linear space Rn+1 with pseudometric

G = (dx0 )2 (dx1 )2 (dx2 )2 (dxn )2 .


For an arbitrary vector X = (a0 , a1 , a2 , . . . , an ) scalar product (X, X) is positive if (a0 )2 > (a1 )2 + (a2 )2 + + (an )2 , it is negative if (a0 )2 < (a1 )2 +
(a2 )2 + + (an )2 , and X is null-vector if (a0 )2 = (a1 )2 + (a2 )2 + + (an )2 .

1.3

Scalar product, conformally Euclidean metric. Length


of tangent vectors and angle between them. Length
of curves

The Riemannian metric defines scalar product of tangent vectors attached


at the given point. Hence it defines the length of tangent vectors and angle
between them. If X = X m xm , Y = Y m xm are two tangent vectors at the
given point p of Riemannian manifold with coordinates x1 , . . . , xn , then we
have that lengths of these vectors equal to
p
p
p
p
|X| = hX, Xi = gik (x)X i X k , |Y| = hY, Yi = gik (x)Y i Y k ,
(1.22)
and an angle between these vectors is defined by the relation
cos =

gik X i Y k
hX, Yi
p
=p
|X| |Y|
gik (x)X i X k gik (x)Y i Y k

(1.23)

Remark We say angle but we calculate just cosinus of angle.


Example Let M be 3-dimensional Riemannian manifold. Consider the
vectors X = 2x + 2y z and Y = x 2y 2z 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
G=

dx2 + dy 2 + dz 2
.
(1 + x2 + y 2 )2

(1.24)

We see that matrix of Riemannian metric gik = (x, y, z)ik , where (x, y, z) =
1
is a scalar function, i.e. matrix G = ||gik || is proportional to
(1+x2 +y 2 +z 2 )2
1

In the case n = 3 it is so called Minkovski space. The coordinate x0 plays a role of


the time: x0 = ct, where c is the value of the speed of the light. Vectors X such that
(X, X) > 0 are called time-like vectors and they called space-like vectors if (X, X) < 0

10

unity matrix. According to formulae above

p
p
p
p
|X| = hX, Xi = gik (x)X i X k = (x, y, z) X i X i = 3 (x, y, z) = 1 .
The same answer for |Y|. The scalar product between vectors X, Y equal to
zero:
hX, Yi = (x, y, z)ik X i Y k = 0
Hence these vectors are unit vectors which are orthogonal to each other.
(ik is Kronecker symbol: ik = 1 if i = k and it vanishes otherwise.)
This example induces the notion of so called conformally euclidean metric
1.3.1

Conformally Euclidean metric

??
Definition We say that metric G is locally conformally Euclidean in a
vicinity of the point p if in a vicinity of this point there exist local coordinates
{xi } such that in these coordinates metric has an appearance

G = (x)ik dxi dxk = (x) (dx1 )2 + + (dxn )2 ,
(1.25)
i.e. it is proportional to Euclidean metric. We say that metric is conformally
Euclidean if it is locally conformally Euclidean in the vicinity of every point.
In the example above we considered just conformally Euclidean metric. One
can see that for conformally Euclidean metric the angle between vectors
(more precisely the cosinus of the angle) is the same as for Eucldean vectors.
It is evident that coefficient (x) in (1.25) has to be positive. It is convenient sometimes to denote it as (x) = e(x) ,
G = (x)ik dxi dxk = e(x) ik dxi dxk ,

(1.26)

Let G be conformally Euclidean metric and let xi be local coordinates such


that the metric has an appearnace (1.25) in these coordinates. Then it is easy
to see that the angle between two vectors (1.23) (more exactly the cosinus
of the angle) has the same appearance as for Euclidean case. Namely let
we have two non-vanishing vectors X = X m (x) xm , Y = Y m (x) xm (|X| 6=
0, |Y| =
6 0). Then
hX, Yi
gik X i Y k
p
p
cos =
=
=
|X| |Y|
gik (x)X i X k gik (x)Y i Y k
11

P k k
(x)ik X i Y k
kX Y
p
p
pP
= pP
.
kX k
kY k
(x)X
Y
(x)ik (x)X i X k (x)ik (x)Y i Y k
k
k
(1.27)
(see also the example above)
We say that Riemannian manifold (M.G) is conformally Euclidean if the
metric G on it is conformally Euclidean.
Later we cosnider many interesting examples of confromally Euclidean
Riemannian manifolds.
1.3.2

Length of curves

Let : xi = xi (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
defined:
1
x (t)

v(t) =


x n (t)
The length of velocity vector v Tx M (vector v is tangent to the manifold
M at the point x) equals to
r
q
p


dxi (t) dxk (t)
.
|v|x = hv, viG x = gik v i v k x = gik
dt
dt x x
For an arbitrary curve its length is equal to the integral of the length of
velocity vector:
Z bq
Z bp


L =
hv, viG x(t) dt =
gik (x(t))x i (t)x k (t)dt .
(1.28)
a

Bearing in mind that metric (1.15) defines the length we often write metric
in the following form
ds2 = gik dxi dxk
For example consider 2-dimensional Riemannian manifold with metric


g11 (u, v) g12 (u, v)
||gik (u, v)|| =
.
g21 (u, v) g22 (u, v)
12

Then
G = ds2 = gik dui dv k = g11 (u, v)du2 + 2g12 (u, v)dudv + g22 (u, v)dv 2 .
The length of the curveR : p
u = u(t), v R= v(t),
where t0 t t1 according to
t
t p
(1.28) is equal to L = t01 hv, vi = t01 gik (x)x i x k =
Z t1 q
g11 (u (t) , v (t)) u2t + 2g12 (u (t) , v (t)) ut vt + g22 (u (t) , v (t)) vt2 dt .
t0

(1.29)
The length of curves defined by the formula(1.28) obeys the following
natural conditions
It coincides with the usual length in the Euclidean space En (Rn with
standard metric G = (dx1 )2 + + (dxn )2 in Cartesian coordinates).
E.g. for 3-dimensional Euclidean space
Z bp
Z bp
i
k
gik (x(t))x (t)x (t)dt =
(x 1 (t))2 + (x 2 (t))2 + (x 3 (t))2 dt
L =
a

It does not depend on parameterisation of the curve


Z bp
Z
i
k
L =
gik (x(t))x (t)x (t)dt =

b0

p
gik (x( ))x i ( )x k ( )d ,

a0

(xi ( ) = xi (t( )), a0 b0 while a t b) since under changing of


parameterisation
x i ( ) =

dx(t( )) dt
dt
dx(t( ))
=
= x i (t) .
d
dt
d
d

It does not depend on coordinates on Riemannian manifold M


Z bp
Z bp
L =
gik (x(t))x i (t)x k (t)dt =
gi0 k0 (x0 (t))x i0 (t)x k0 (t)dt .
a

This immediately follows from transformation rule (1.67) for Riemannian metric:
 k


x k0
xi i0
i0
k0
(t) gik x i (t)x k (t) .
gi0 k0 x (t)x (t) = gik
x (t)
0
0x
i
k
x (t)
x
13

It is additive: length of the sum of two curves is equal to the sum


of their lengths. If a curve = 1 + , i.e. : xi (t), a t b,
1 : xi (t), a t c and 2 : xi (t), c t b where a point c belongs to
the interval (a, b) then L = L 1 + L 2 .
One can show that formula (1.28) for length is defined uniquely by these conditions.
More precisely one can show under some technical conditions one may show that any local
additive functional on curves which does not depend on coordinates and parameterisation,
and depends on derivatives of curves of order 1 is equal to (1.28) up to a constant
multiplier. To feel the tasteof this statement youmay do the following exercise:
dy(t)
be a function such that an integral
Exercise Let A = A x(t), y(t), dx(t)
dt , dt

R 
dx(t) dy(t)
L = A x(t), y(t), dt , dt dt over an arbitrary curve in E2 does not change under
reparameterisation of this curve and under an arbitrary isometry, i.e. translation and
rotation of the curve. Then one can easy show (show it!) that
s

2 
2

dy(t)
dx(t) dy(t)
dx(t)
,
=c
+
,
A x(t), y(t),
dt
dt
dt
dt
where c is a constant, i.e. it is a usual length up to a multiplier

1.4

Riemannian structure on the surfaces embedded


in Euclidean space

In first
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 E3 attached at the point p E3 using scalar product in E3 . 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 (1.33)). Respectively
If L is a curve in M then an External Observer consider this curve as a
curve in E3 , 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 define the modulus of the velocity vector and the
length of the curve using Riemannian metric.
Definition Let M be a surface embedded in the Euclidean space. We
say that metric GM on the surface is induced by the Euclidean metric if the
14

scalar product of arbitrary two vectors A, B Tp M calculated in terms of


the metric G equals to Euclidean scalar product of these two vectors:
hA, BiGM = hA, BiGEuclidean

(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 definition 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
Here we recall basic notions from the course of Geometry which we will need
here.
Let r = r(u, v) be parameterisation of the surface M embedded in the
Euclidean space:

x(u, v)
r(u, v) = y(u, v)
z(u, v)
Here as always x, y, z are Cartesian coordinates in E3 .
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
Tp M .
For a point p M one can consider a basis in the tangent plane Tp M
adjusted to the parameters u, v. Tangent basis vectors at any point (u, v)
are

x(u,v)

u
y(u, v)
z(u, v)
r(u, v) y(u,v)
x(u, v)
= u =
+
+
ru =
u
u x
u y
u z
z(u,v)
u

Every vector X Tp M can be expanded over this basis:


X = Xu ru + Xv rv ,
15

where Xu , Xv are coefficients, components of the vector X.


Internal Observer views the basis vector ru Tp M , as a velocity vector
for the curve u = u0 + t, v = v0 , where (u0 , v0 ) are coordinates of the point p.
Respectively the basis vector rv Tp M for an Internal Observer, is velocity
vector for the curve u = u0 , v = v0 + t, where (u0 , v0 ) are coordinates of the
point p.
1.4.2

Explicit formulae for induced Riemannian metric (First Quadratic


form)

Now we are ready to write down the explicit formulae for the Riemannian
metric on the surface induced by metric (scalar product) in ambient Euclidean space (see the Definition (1.30)).
Let M : r = r(u, v) be a surface embedded in E3 .
The formula (1.30) means that scalar products of basic vectors ru =
u , rv = v has to be the same calculated in the ambient space and on
the surface: For example scalar product hu , v iM = guv calculated by the
Internal Observer is the same as a scalar product hru , rv iE3 calculated by the
External Observer, scalar product hv , v iM = guv calculated by the Internal
Observer is the same as a scalar product hrv , rv iE3 calculated by the External
Observer and so on:
 
 


hu , u i hu , v i
hru , ru iE3 hru , rv iE3
guu guv
=
=
G=
(1.31)
gvu gvv
hv , u i hv , v i
hrv , ru iE3 hrv , rv iE3
where as usual we denote by h , iE3 the scalar product in the ambient Euclidean space. (Here see also the important remark (1.13))
Remark It is convenient sometimes to denote parameters (u, v) as (u1 , u2 )
or u ( = 1, 2) and to write r = r(u1 , u2 ) or r = r(u ) ( = 1, 2) instead
r = r(u, v)
In these notations:


 
g11 g12
hru , ru iE3 hru , rv iE3
, g = hr , r i ,
GM =
=
g12 g22
hru , rv iE3 hrv , rv iE3
GM = g du du = g11 du2 + 2g12 dudv + g22 dv 2

(1.32)

where ( , ) is a scalar product in Euclidean space.


The formula (1.32) is the formula for induced Riemannian metric on the
surface r = r(u, v). It is First Quadratic Form of this surface.
16

If X, Y are two tangent vectors in the tangent plane Tp C then G(X, Y)


at the point p is equal to scalar product of vectors X, Y:
(X, Y) = (X 1 r1 + X 2 r2 , Y 1 r1 + Y 2 r2 ) =

(1.33)

X 1 (r1 , r1 )Y 1 + X 1 (r1 , r2 )Y 2 + X 2 (r2 , r1 )Y 1 + X 2 (r2 , r2 )Y 2 =


X (r , r )Y = X g Y = G(X, Y)
We can come to this formula just transforming differentials. In Cartesian
coordinates hX, Yi = X 1 Y 1 + X 2 Y 2 + X 3 Y 3 , i.e. the Riemannian structure
of Euclidean space in Cartesian coordinates is given by
GE3 = (dx)2 + (dy)2 + (dz)2 .

(1.34)

The condition that Riemannian metric (1.32) is induced by Euclidean scalar


product means that


GE3 r=r(u,v) = (dx)2 + (dy)2 + (dz)2 r=r(u,v) = GM = g du du

(1.35)


i.e. ((dx)2 + (dy)2 + (dz)2 ) r=r(u,v) =


2 
2 
2
x(u, v)
x(u, v)
x(u, v)
x(u, v)
x(u, v)
x(u, v)
du +
du +
du +
du +
du +
du =
u
u
u
u
u
u
(x2u + yu2 + zu2 )du2 + 2(xu xv + yu yv + zu zv )dudv + (x2v + yv2 + zv2 )dv 2

We see that
GM = g du du = g11 du2 + 2g12 dudv + g22 dv 2 ,

(1.36)

where g11 = guu = (x2u + yu2 + zu2 ) = hru , ru iE3 , g12 = g21 = guv = gvu =
(xu xv + yu yv + zu zv ) = hru , rv iE3 , g22 = gvv = (x2v + yv2 + zv2 ) = hrv , rv iE3 . We
come to same formula (1.32).
(See the examples of calculations in the next subsection.)
Remark Sometimes it is convenient to denote Cartesian coordinates of
Euclidean space by xi , (i = 1, 2, 3). Let surface M be given in local parameterisation xi = xi (u ). Riemannian metric of Euclidean space (1.34) has
appearance
GE = dxi ik dxk .
(1.37)
17

and induced metric (1.35) has appearance



xi (u) xk (u)
GM = dxi ik dxk xi =xi (u ) =
ik
du du = g (u)du du
u
u
(1.38)
Why this representation is useful? it is easy to see that formulae (1.37) ,(1.38) work for
arbitrary dimensions, i.e. if we have m-dimensional manifold embedded in n-dimensional
Euclidean space. We just have to suppose that in this case i = 1, . . . , n and = 1, . . . , m;
manifold is given by parameterisation xi = xi (u ) ( = 1, . . . , m). Moreover in the face if
manifold is embedded not in Euclidean space but in an arbitrary Riemannian space then
one can see comparing formulae (1.30) and (1.37) we come to the induced metric

xk (u)
xi (u)
gik ((x(u)))
du du = g (x(u))du du
GM = dxi gik ((x(u))) dxk xi =xi (u ) =

u
u

Check explicitly again that length of the tangent vectors and curves on
the surface calculating by External observer (i.e. using Euclidean metric
(1.34)) is the same as calculating by Internal Observer, ant (i.e. using the
induced Riemannian metric (1.32), (1.36)). Let X = X r = aru + brv be a
vector tangent to the surface M . The square of the length |X| of this vector
calculated by External observer (he calculates using the scalar product in
E3 ) equals to
|X|2 = hX, Xi = hru + brv , aru + brv i = a2 hru , ru i + 2abhru , rv i + b2 hrv , rv i
(1.39)
where h , i is a scalar product in E3 . The internal observer will calculate the
length using Riemannian metric (1.32) (1.36):

  
 g11 g12
a
G(X, X) = a, b

= g11 a2 + 2g12 ab + g22 b2


(1.40)
g21 G22
b
External observer (person living in ambient space E3 ) calculate the length
of the tangent vector using formula (1.39). An ant living on the surface
calculate length of this vector in internal coordinates using formula (1.40).
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 = ru + rv where = ut , = vt :
18

v=

dr(t)
dt

= ut ru + vt rv .

The length of the curve is equal to


Z bp
Z bp
Z b
|v(t)|dt =
hv(t), v(t)iE3 dt =
hut ru + vt rv , ut ru + vt rv iE3 dt =
L=
a

(1.41)
Z bq
hru , ru iE3 u2t + 2hru , rv iE3 ut vt + hrv , rv iE3 vt2 d =
a

Z bp
g11 u2t + 2g12 ut vt + g22 vt2 dt

(1.42)

An external observer will calculate the length of the curve using (1.41).
An ant living on the surface calculate length of the curve using (1.42) using
Riemannian metric on the surface. They will come to the same answer.
1.4.3

Induced Riemannian metrics. Examples.

We consider here examples of calculating induced Riemannian metric on some


quadratic surfaces in E3 . using calculations for tangent vectors (see (1.32))
or explicitly in terms of differentials (see (1.35) and (1.36)).
First of all consider the general case when a surface M is defined by the
equation z F (x, y) = 0. One can consider the following parameterisation
of this surface:

x = u
r(u, v) :
(1.43)
y=v

z = F (u, v)
Then

1
ru = 0
Fu

0
rv = 1
Fv

(1.44)

,
(ru , ru ) = 1 + Fu2 ,

(ru , rv ) = Fu Fv ,

(rv , rv ) = 1 + Fv2

and induced Riemannian metric (first quadratic form) (1.32) is equal to



 
 

g11 g12
(ru , ru ) (ru , rv )
1 + Fu2 Fu Fv
||g || =
=
=
(1.45)
g12 g22
(ru , rv ) (rv , rv )
Fu Fv 1 + Fv2
GM = ds2 = (1 + Fu2 )du2 + 2Fu Fv dudv + (1 + Fv2 )dv 2
19

(1.46)

and the length of the curve r(t) = r(u(t), v(t)) on C (a t b) can be


calculated by the formula:
Z Z bq
(1 + Fu2 )u2t + 2Fu Fv ut vt + (1 + Fv )2 vt2 dt
L=
a

One can calculate (1.46) explicitly using (1.35):



GM = dx2 + dy 2 + dz 2 x=u,y=v,z=F (u,v) = (du)2 + (dv)2 + (Fu du + Fv dv)2 =
= (1 + Fu2 )du2 + 2Fu Fv dudv + (1 + Fv2 )dv 2 .

(1.47)

Cylinder
Cylinder is given by the equation x2 + y 2 = a2 . One can consider the
following parameterisation of this surface:

x = a cos
r(h, ) :
(1.48)
y = a sin

z=h

We have Gcylinder = (dx2 + dy 2 + dz 2 ) x=a cos ,y=a sin ,z=h =
= (a sin d)2 + (a cos d)2 + dh2 = a2 d2 + dh2
The same formula in terms of scalar product of tangent vectors:

0
a sin
rh = 0 r = a cos
1
0

(1.49)

(1.50)

,
(rh , rh ) = 1,
and


||g || =

(rh , r ) = 0,

(r , r ) = a2

 

(ru , ru ) (ru , rv )
1 0
=
,
(ru , rv ) (rv , rv )
0 a2
G = dh2 + a2 d2

(1.51)

and the length of the curve r(t) = r(h(t), (t)) on the cylinder (a t b)
can be calculated by the formula:
Z bp
h2t + a2 t dt
(1.52)
L=
a

20

Cone
Cone is given by the equation x2 + y 2 k 2 z 2 = 0. One can consider the
following parameterisation of this surface:

x = kh cos
r(h, ) :
(1.53)
y = kh sin

z=h
Calculate induced Riemannian metric:
We have

Gconus = dx2 + dy 2 + dz 2 x=kh cos ,y=kh sin ,z=h =
(k cos dh kh sin d)2 + (k sin dh + kh cos d)2 + dh2


1 + k2
0
2 2
2
2
2
Gconus = k h d + (1 + k )dh , ||g || =
0
k 2 h2

(1.54)

The length of the curve r(t) = r(h(t), (t)) on the cone (a t b) can be
calculated by the formula:
Z bq
(1 + k 2 )h2t + k 2 h2 2t dt
(1.55)
L=
a

Circle
Circle of radius R is given by the equation x2 + y 2 = R2 . Consider
standard parameterisation of this surface:
(
x = R cos
r() :
y = R sin
Calculate induced Riemannian metric (first quadratic form)

GS 2 = dx2 + dy 2 x=R cos ,y=R sin =
(R sin d)2 + (a cos d)2 = (R2 cos2 + R2 sin2 )d2 = R2 d2 .
One can consider stereographic coordinates on the circle (see Example in
the subsection 1.1) A point x, y : x2 + y 2 = R2 has stereographic coordinate
21

t if points (0, 1) (north pole), the point (x, y) and the point (t, 0) belong to
, i.e.
the same line, i.e. xt = Ry
R
(
2
x = R2tR
Rx
2 +t2
t=
,
. since x2 + y 2 = R2 .
2 R2
t
Ry
y = t2 +R2 R
Induced metric in coordinate t is
2   2
 
2

t R2
2tR2
2
2
+ d
G = (dx + dy ) x=x(t),y=y(t) = d
R
=
R2 + t2
R2 + t2
2 
2

4R2 tdt
4R4 dt2
2R2 dt
4t2 R2 dt
+

.
R2 + t2 (R2 + t2 )2
(t2 + R2 )2
(R2 + t2 )2
(See for detail Homework 2)
Remark Stereographic coordinates very often are preferable since they
define birational equivalence between circle and line.
Sphere
Sphere of radius R is given by the equation x2 + y 2 + z 2 = R2 . Consider
the following (standard ) parameterisation of this surface:

x = R sin cos
r(, ) :
(1.56)
y = R sin sin

z = R cos
Calculate induced Riemannian metric (first quadratic form)

GS 2 = dx2 + dy 2 + dz 2 x=R sin cos ,y=R sin sin ,z=R cos =
(R cos cos dR sin sin d)2 +(R cos sin d+R sin cos d)2 +(R sin d)2 =

R2 cos2 d2 + R2 sin2 d2 + R2 sin2 d2 =


 2

R
0
2
2
2
2
2
= R d + R sin d ,
||g || =
0 R2 sin2

(1.57)

One comes to the same answer calculating scalar product of tangent vectors:

R cos cos
R sin sin
r = R cos sin r = R sin cos
R sin
0
22

,
(r , r ) = R2 ,
and

 2
R
0

(rh , r ) = 0,

(r , r ) = R2 sin2



(ru , ru ) (ru , rv )
||g|| =
=
(ru , rv ) (rv , rv )

0
, GS 2 = ds2 = R2 d2 + R2 sin2 d2
R2 sin2

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:
Z

L=

t2 + sin2 2t dt

(1.58)

One can consider on sphere as well as on a circle stereographic coordinates:

2uR2
(

x = R2 +u
2 +v 2

Rx
u = Rz
2vR2
(1.59)
,
y = R2 +u2 +v2
Ry

v = Rz

u2 +v 2 R2
z = u2 +v2 +R2 R
In these coordinates Riemannian metric is

G = (dx2 + dy 2 + dz 2 ) x=x(u,v),y=y(u,v),z=z(u,v) =
 
d

2uR2
R2 + u2 + v 2

2  
+ d

2vR2
R 2 + u2 + v 2

2  
+ d 1

2R2
R2 + u2 + v 2

 2
R =

4R4 (du2 + dv 2 )
.
(R2 + u2 + v 2 )2

(see in detail Homework 2 and Solutions)


Notice that we show that metric on sphere is conformally Euclidean.
Saddle (paraboloid)2
Consider paraboloid z = x2 y 2 . It can be rewritten as z = axy and
it is called sometimes saddle (rotation on the angle = /4 transforms
z = x2 y 2 onto z = 2xy.) Paraboloid and saddle they are ruled surfaces which are
formed by lines. We considered this surface in the course of Geometry.
2

This example was not considered on lectures. It could be useful for learning purposes.

23

Consider the following (standard ) parameterisation of this surface:

x = u
r(u, v) :
(1.60)
y=v

z = uv
Calculate induced metric:

Gsaddle = dx2 + dy 2 + dz 2 x=u cos ,y=v sin ,z=uv = du2 +dv 2 +(udv +vdu)2 =
Gsaddle = (1 + v 2 )du2 + 2uvdudv + (1 + u2 )dv 2 .
One-sheeted and two-sheeted hyperboloids.
These examples were mostly considered on tutorials.
Consider surface given by the equation
x2 + 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 E3 .
If c < 0 it is two-sheeted hyperboloid a surface with two sheets: upper
sheet z > 0 and another sheet: z < 0.
Consider these cases separately.
1) One-sheeted hyperboloid: x2 + y 2 z 2 = a2 . It is ruled surface.
Exercise Find the lines on two-sheeted hyperboloid
One-sheeted hyperboloid is given by the equation x2 + y 2 z 2 = a2 . It is
convenient to choose parameterisation:

x = a cosh cos
r(, ) :
(1.61)
y = a cosh sin

z = a sinh
x2 + y 2 z 2 = a2 cosh2 a2 sinh2 = a2 .
(Compare the calculations with calculations for sphere! We changed functions cos, sin on cosh, sinh.)
Induced Riemannian metric (first quadratic form) is equal to

GHyperbolI = dx2 + dy 2 + dz 2
=
x=a cosh cos ,y=a cosh sin ,z=a sinh

24

(a sinh cos da cosh sin d)2 +(a sinh sin d+a cosh cos d)2 +(a cosh d)2 =

a2 sinh2 d2 + a2 cosh2 d2 + a2 cosh2 d2 =



 2
a (1 + 2 sinh2 )
0
2
2
2
2
2
2
= a (1+2 sinh )d +a cosh d ,
||g || =
0
a2 cosh2

2) Two-sheeted hyperboloid: z 2 x2 y 2 = a2 . It is not ruled surface!


For two-sheeted hyperboloid calculations will be very similar.
In the same way as for one-sheeted hyperboloid (see equation (1.61)) it
is convenient to choose parameterisation:

x = a sinh cos
(1.62)
r(, ) :
y = a sinh sin

z = a cosh
z 2 x2 y 2 = a2 cosh2 a2 sinh2 = a2
(Compare the calculations with calculations for sphere and one-sheeted hyperboloid.
Induced Riemannian metric (first quadratic form) is equal to

GHyperbolI = dx2 + 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 =

a2 cosh2 d2 + a2 sinh2 d2 + a2 sinh2 d2 =


 2

a (1 + 2 sinh2 )
0
2
2
2
2
2
2
= a (1+2 sinh )d +a sinh d ,
||g || =
0
a2 sinh2
(1.63)

We calculated examples of induced Riemannian structure embedded in Euclidean space


almost for all quadratic surfaces.
Quadratic surface is a surface defined by the equation
Ax2 + By 2 + Cz 2 + 2Dxy + 2Exz + 2F yz + ex + f y + dz + c = 0
One can see that any quadratic surface by affine transformation can be transformed to
one of these surfaces
cylinder (elliptic cylinder) x2 + y 2 = 1
hyperbolic cylinder: x2 y 2 = 1)

25

parabolic cylinder z = x2
paraboloid x2 + y 2 = z
hyperbolic paraboloid x2 y 2 = z
cone x2 + y 2 z 2 = 0
sphere x2 + y 2 + z 2 = 1
one-sheeted hyperboloid x2 + y 2 z 2 = 1
two-sheeted hyperboloid z 2 x2 y 2 = 1
(We exclude degenerate cases such as point x2 + y 2 + z 2 = 0, planes, e.t.c.)

1.4.4

Induced metric on two-sheeted hyperboloid embedded in


pseudo-Euclidean space.

Consider the same two-sheeted hyperboloid z 2 x2 y 2 = 1 embedded R3 (See


equation (1.62). For simplicity we assume now that a = 1.) Now we consider the
ambient space R3 not as Euclidean space but as pseudo-Euclidean space, i.e. in
R3 instead standard scalar product
hX, Yi = X 1 Y 1 + X 2 Y 2 + X 3 Y 3
we consider pseudo-scalar product defined by bilinear form
hX, Yipseud = X 1 Y 1 + X 2 Y 2 X 3 Y 3
The pseudoscalar product is bilinear, symmetric. It is defined by non-degenerate
matrix. But it is not positive-definite. E.g. The pseudo-length of vectors X =
(a cos , a sin , a) is equals to zero (such vectors are called null vectors):
X = (a cos , a sin , a) hX, Xipseudo = 0,
The corresponding pseudo-Riemannian metric is:
Gpseudo = dx2 + dy 2 dz 2

(1.64)

It turns out that the following remarkable fact occurs:


Proposition The pseudo-Riemannian metric (1.64) in the ambient 3-dimensional
pseudo-Euclidean space induces Riemannian metric on two-sheeted hyperboloid
x2 + y 2 z 2 = 1.
Remark This is not the fact for one-sheeted hyperboloid (see problem 7 in
Homework 2)

26

Show it. (See also problems 5 and 6 in Homework 2. ) Repeat the calculations
above for two-sheeted hyperboloid changing in the ambient space Riemannian
metric G = dx2 + dy 2 + dz 2 on pseudo-Riemannian dx2 + dy 2 dz 2 :
Using (1.62) and (1.64) we come now to

G = dx2 + 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 =
a2 cosh2 d2 + a2 sinh2 d2 a2 sinh2 d2


1
0
2 2
2
2
2
GL = a d + a sinh d ,
||g || =
0 sinh2

(1.65)

The two-sheeted hyperboloid equipped with this metric is called hyperbolic or


Lobachevsky plane.
Now express Riemannian metric in stereographic coordinates. (We did it in
detail in homework 2)
Calculations are very similar to the case of stereographic coordinates of 2sphere x2 + y 2 + z 2 = 1. (See homework 1). Centre of projection (0, 0, 1): For
1
stereographic coordinates u, v we have ux = yv = 1+z
. We come to
(
u=
v=

x
1+z
y
1+z

x =
y=

z=

2u
1u2 v 2
2v
1u2 v 2
2
u +v 2 +1
1u2 v 2

(4)

The image of upper-sheet is an open disc u2 + v 2 = 1 since u2 + v 2 =


z 2 1
(1+z)2

z1
z+1 .

Since for upper sheet z > 1 then 0

z1
z+1


G = (dx + dy dz ) x=x(u,v),y=y(u,v),z=z(u,v) =
 
d

2v
1 u2 v 2

2

x2 +y 2
(1+z)2

< 1.

 
d

2u
1 u2 v 2

2
+

2
  2
4(du)2 + 4(dv)2
u + v2 + 1
d
=
.
1 u2 v 2
(1 u2 v 2 )2

These coordinates are very illuminating. One can show that we come to so called
hyperbolic plane (see in detail Homework 2)

27

1.5

Isometries of Riemannian manifolds.

Let (M1 , G(1) ), (M2 , G(2) ) be two Riemannian manifolds manifolds equipped
with Riemannian metric G(1) and G(2) respectively.
Loosely speaking isometry is the diffeomorphism of Riemannian manifolds
which preserves the distance.
Definition Let F be a diffeomorphims (one-one smooth map with smooth
inverse) of manifold M1 on manifold M2 .
We say that diffeomorphism F is an isometry of Riemannian manifolds (M1 , G(1) )
and (M2 , G(2) ) if it preserves the metrics, i.e. G(1) is pull-back of G(2) :
F G(2) = G(1) .

(1.66)

In local coordinates this means the following: Let p1 be an arbitrary point on


manifold M1 and p2 M2 be its image:F (p1 ) = p2 . Let {xi } be arbitrary
oordinates in a vicinity of a point p1 M1 and {y a } be arbitrary coordinates
in a vicinity of a point p2 M2 . Let Riemannian metrics G(1) on M1 has local
expression G(1) = g(1)ik (x)dxi dxk in coordinates {xi } and respectively Riemannian
metrics G(2) has local expression G(2) = g(2)ab (y)dy a dy b in coordinates {y a } on M2 .
Then the formula (1.66) has the following appearance in these local coordinates:



F g(2) ab (y)dy a dy b = g(2) ab (y)dy a dy b y=y(x) =
g(2) ab (y(x))

y a (x) i y b (x) k
dx
dx = g(1) ik (x)dxi dxk ,
xi
xk

(1.67)

i.e.

y b (x)
y a (x)
(y(x))
g
,
(1.68)
ab
xi (2)
xk
where y a = y a (x) is local expression for diffeomorphism F . We say that diffeomorphism F is isometry of Riemanian manifolds (M1 , G(1) ) and M2 , G(2) .
Diffeomorphism F establishes one-one correspondence between local coordinates on manifolds M1 and M2 . The left hand side of equation (1.67) can be
considered as a local expression of metric G(2) in coordinates xi on M2 and the
right hand side of this equation is local expression of metric G(1) in coordinates xi
on M1 . Diffeomorphism F identifies manifolds M1 and M2 and it can be considered
as changing of coordinates.
g(1) ik (x) =

Example Consider surface of cylinder C, x2 + y 2 = a2 in E3 with induced


Riemannian metric GC = a2 d2 + dh2 (see equations (1.48) and (1.49)). If we
remove the line l : x = a, y = 0 from the cylinder surface C we come to surface

28

C 0 = C\l. Consider a map F of this surface in Euclidean space E 2 with Cartesian


coordinates u, v (with standard Euclidean metric GEucl = du2 + dv 2 ):
(
u = a
F:
0 < < 2 .
(1.69)
v=h
One can see that F is the diffeomorphism of C 0 on the domain 0 < u < 2a
in E2 and this diffeomorphism is an isometry: it transforms the metric GEucl on
Euclidean space in metric GC on cylinder, i.e. pull-back condition (1.66) is obeyed:


F GEucl = F du2 + dv 2 = du2 + dv 2 u=a,v=h = a2 d2 + dh2 = G1 .
We see that cylinder surface with removed line is isometric to domain in E2 and
the map F establishes this isometry.
Remark Let F be diffeomorphism of manifold M1 on a manifold M2 . Let
a manifold M2 be equipped with Riemannian metric G(2) . Then consider the
pull-back of this metric, Riemannian metric G(1) = F G(2) on M1 . We see that
diffeomorphism F is an isometry of Riemannian manifold (M1 , G(1) ) on Riemannian manifold (M2 , G(2) ).

1.5.1

Isometries of Riemannian manifold (on itself )

Definition Let (M, G) be a Riemannian manifold. We say that a diffeomorphism


F is an isometry of Riemannian manifold on itself if it preserves the metric, i.e.
F G = G. In local coordinates this means that
gik (x) = gpq (x0 (x))

xp (x0 ) xq (x0 )
,
xi
xk

(1.70)

where x0 = x0 (x) is a local expression for diffeomorphism F . Example Let E2 be


Euclidean plane with metric dx2 + dy 2 in Cartesian coordinates x, y. Consider the
transformation
(
x0 = p + ax + by
y 0 = q + cx + dy


a b
is isometry if and only if the matrix A =
is an orthogonal matrix, i.e. if
c d
the trasformation above is combination of translation, rotation and reflection.

29

1.6

Infinitesimal isometries of Riemannian manifold


(Killing vector fields)

Let X be an arbitrary vector field on Riemannian manifold M . It induces infinitesimal diffeomorphism


0

F : xi = xi + X i (x),

where 2 = 0 .

(the condition 2 = 0 reflects the fact that we ignore terms of order 2 over
.) Find a condition which guarantees that infinitesimal diffeomorphism is an
0
isometry. If xi = xi + X i (x), then one can see that the inverse infinitesimal
0
diffeomorphism is defined by the equation xi = xi X i (x0 ) and equation (1.70)
implies that



pX p (x)
pX q (x)
xp (x0 ) xq (x0 )
p
q
i
i
0
= gpq (x +X ) i +
k +
=
gik (x) = gpq (x (x))
xi
xi
xk
xk


gik (x)
X q (x)
X p (x)
p
gik (x) + X (x)
+ giq (x)
+ gpk (x)
,
xp
xi
xk
Here we consider only terms of first and zero order over since 2 = 0 (this
is related with the fact that transformation is infinitesimal). The last relation
implies that
X p (x)

gik (x)
X q (x)
X p (x)
+
g
(x)
= 0.
+
g
(x)
iq
pk
xp
xi
xk

(1.71)

Left hand side of this relation we denote LX G Lie derivative of Riemannian


metric along vector field X. Vector field X induces isometry if Lie derivative of
metric along this vector field vanishes. We come to
Proposition Vector field X on Riemannian manifold (M, G) induces infinitesimal isometry if LX G = 0:
LX G = X p (x)

gik (x)
X q (x)
X p (x)
+ giq (x)
+ gpk (x)
= 0.
p
k
x
xi
x

(1.72)

Definition) We call vector field X Killing vector field) if it preserves the metric,
i.e. if equation (1.72) is obeyed.
Example Consider plane (x, y) with Riemannian metric G = (x, y)(dx2 +
2
dy ). Find differential equation for infinitesimal isometries of this metric, i.e. write
down equations (1.72) for this
 metric.

(x, y)
0
We have ||gik (x, y)|| =
.
0
(x, y)

30

Let X = A(x, y)x + B(x, y)y . Write down equations (1.72) for components
g1 1, g1 2, g21 and g22 : We will have the following three equations

A(x,y)

A(x, y) x + B(x, y) y + 2 x = 0 for component g11


B(x,y)

(1.73)
A(x, y)
x + B(x, y) y + 2 y = 0 for component g22

B(x,y) A(x,y
+ y = 0 for components g12 and g21
x
Practically for sphere, Lobachevsky plane, e.t.c. it is much easier to find the
Killing fields not solving these equations, but considering the usual isometries (see
examples in solutions of Coursework and in the Appendix about Killing vector
fields for Lobachevsky plane.))
Another simple and interesting exercise: How look Killing vectors for Euclidean
space En . In this case we come from (1.72) to equation
LK G = iq (x)
i.e.

K q (x)
K p (x)
+

(x)
= 0,
pk
xi
xk

K i (x) K k (x)
= 0.
+
xi
xk

(1.74)

Solve this equation. Differentiating by x we come to


2 K i (x) K k (x)
+ m i =0
xm xk
x x
Consider tensor field

2K i
xm xk
It follows from equation (1.74) that
i
Tmk
=

i
i
Tmk
= Tkm
= Tikm .

(1.75)

(1.76)

i
It is easy to see that this implies that Tmk
0!!!:
i
k
k
i
i
i
i
Tmk
= Tikm = Tkim = Tmi
= Tim
= Tkm
= Tmk
Tmk
= Tmk
,
i
i.e. Tmk
=

2 K i (x)
xm xk

= 0. This implies that


K i (x) = C i + Bki xk

We come to
Theorem All infinitesimal isometries of En are translations and infinitesimal rotations.
What happens in general case?
31

1.6.1

Locally Euclidean Riemannian manifolds

It is useful to formulate the local isometry condition between Riemannian


manifold and Euclidean space. A neighbourhood of every point of n-dimensional
manifold is diffeomorphic to Rn . Let as usual En be n-dimensional Euclidean space, i.e. Rn with standard Riemannian metric G = dxi ik dxk =
(dx1 )2 + + (dxn )2 in Cartesian coordinates (x1 , . . . , xn ).
Definition We say that n-dimensional Riemannian manifold (M, G) is
locally isometric to Euclidean space En , i.e. it is locally Euclidean Riemannian manifold, if for every point p M there exists an open neighboorhood
D (domain) containing this point, p D such that D is isometric to a domain in Euclidean plane. In other words in a vicinity of every point p there
exist local coordinates u1 , . . . , un such that Riemannian metric G in these
coordinates has an appearance
G = dui ik duk = (du1 )2 + + (dun )2 .

(1.77)

Consider examples.
Example Consider again cylinder surface..
We know that cylinder is not diffeomorphic to plane (there are plenty
reasons for this). In the previous subsection we cutted the line from cylindre. Thus we came to surface diffeomorphic to plane. We established that
this surface is isometric to Euclidean plane. (See equation (1.69) and considerations above.) Local isometry of cylinder to the Euclidean plane, i.e.
the fact that it is locally Euclidean Riemannian surface immediately follows
from the fact that under changing of local coordinates u = a, v = h in
equation (1.69), the standard Euclidean metric du2 + dv 2 transforms to the
metric Gcylinder = a2 d2 + dh2 on cylinder.
Example Now show that cone is locally Euclidean Riemannian surface,
i.e, it is locally isometric to the Euclidean plane. This means that we have to
find local coordinates u, v on the cone such that in these coordinates induced
metric G|c on cone would have the appearance G|c = du2 + dv 2 . Recall
calculations of the metric on cone in coordinates h, where

x = kh cos
r(h, ) : y = kh sin ,

z=h
32

x2 + y 2 k 2 z 2 = k 2 h2 cos2 + k 2 h2 sin2 k 2 h2 = k 2 h2 k 2 h2 = 0. We have


that metric Gc on the cone in coordinates h, induced with the Euclidean
metric G = dx2 + dy 2 + dz 2 is equal to

Gc = dx2 + dy 2 + dz 2 x=kh cos ,y=kh sin ,z=h = (k cos dh kh sin d)2 +
(k sin dh + kh cos d)2 + dh2 = (k 2 + 1)dh2 + k 2 h2 d2 .
In analogy
with polar coordinates try to find new local coordinates u, v such
(
u = h cos
that
, where , are parameters. We come to du2 + dv 2 =
v = h sin
( cos dh h sin d)2 +( sin dh + h cos d)2 = 2 dh2 +2 2 h2 d2 .
Comparing with the metric on the cone Gc = (1 + k 2 )dh2 + k 2 h2 d2 we see
k
2
2
2
2
2 2 2
2
that if we put = k and = 1+k
2 then du + dv = dh + h d =
(1 + k 2 )dh2 + k 2 h2 d2 .
Thus in new local coordinates
(

u = k 2 + 1h cos kk2 +1

v = k 2 + 1h sin kk2 +1
induced metric on the cone becomes G|c = du2 + dv 2 , i.e. cone locally is
isometric to the Euclidean plane
Of course these coordinates are local. Cone and plane are not homeomorphic, thus they are not globally isometric.
Example and counterexample
Consider domain D in Euclidean plane with two metrics:
G(1) = du2 + sin2 vdv 2 ,

and G(2) = du2 + sin2 udv 2

(1.78)

Thus we have two different Rimeannian manifolds (D, G(1) ) and (D, G(2) ).
Metrics in (1.78) look similar. But.... It is easy to see that the first one is
locally isometric to Euclidean plane, i.e. it is locally Euclidean Riemannian
manifold since sin2 vdv 2 = d( cos v)2 : in new coordinates u0 = u, v 0 = cos v
Riemannian metric G(1) has appearance of standard Euclidean metric:
(du0 )2 + (dv 0 )2 = (du)2 + (d(cos v))2 = du2 + sin2 vdv 2 = G(1) .
33

This is not the case for second metric G(2) . If we change notations u 7 ,
v 7 then G(2) = d2 + sin2 d2 . This is local expression for Riemannian
metric induced on the sphere of radius R = 1. Suppose that there exist
coordinates u0 = u0 (, ) v 0 = v 0 (, ) such that in these coordinates metric
has Eucldean appearance. This means that locally geometry of sphere is as a
geometry of Euclidean plane. On the other hand we know from the course of
Geometry that this is not the case: sum of angles of triangels on the sphere
is not equal to , sphere cannot be bended without shrinking. Later in this
course we will return to this question....
There are plenty other examples:
4 (dx2 +dy 2 )
is isometric to the sphere with radius R.
2) Plane with metric 4R
(R2 +x2 +y 2 )2
3) Disc with metric

du2 +dv 2
(1u2 v 2 )2

is isometric to half plane with metric

dx2 +dy 2

.
(see also exercises in Homeworks and Coursework.)

4y 2

1.6.2

Conformally Euclidean Riemannian manifolds

We have also considered not only locally Euclidean Riemannian manifolds,


but also Riemannian manifolds with conformally Euclidean metrics (see subsection 1.3.1 and equation (1.25) there.) Now we would like again to repeat
the definition of conformally Euclidean manifolds,, to compare it with definition of locally Euclidean Rimeanniam manifolds.
Definition We say that n-dimensional Riemannian manifold (M, G) is
conformally Euclidean, in a vicinity of every point p there exist local coordinates u1 , . . . , un such that Riemannian metric G in these coordinates has
an appearance
G = (x)dui ik duk = (du1 )2 + + (dun )2 .

(1.79)

(Compare with definition of locally Euclidean manifolds (see definition (1.77)).)


Examples of conformally Euclidean manifolds
sphere S n , (one can consider stereographic coordinates to see that in
these coordinates metric is prortional to Euclidean),
Lobachevsky plane
In fact one can show that every two dimensional Riemannian manifold is
conformally Euclidean.
34

1.7

Volume element in Riemannian manifold

The volume element in n-dimensional Riemannian manifold with metric G =


gik dxi dxk is defined by the formula
p
det gik dx1 dx2 . . . dxn .
(1.80)
If D is a domain in the n-dimensional Riemannian manifold with metric
G = gik dxi then its volume is equal to to the integral of volume element over
this domain.
Z p
V (D) =
det gik dx1 dx2 . . . dxn .
(1.81)
D

Remark Students who know the concept of exterior forms can read the
volume element as n-form
p
det gik dx1 dx2 dxn .
Note that in the case of n = 1 volume is just the length, in the case if
n = 2 it is area.
1.7.1

Motivation: Gramm formula for volume of parallelepiped

Explain how formulae (1.80), (1.81) are related with basic formulae of linear
algebra.
Recall the formulae for calculation of volume of n-dimensional parallelepiped. Let En be Euclidean vector space equipped with orthonormal
basis {ei }. i
Let {ai } = {v1 , . . . , an } be an arbitrary basis in this vector space. Consider n-dimensional parallelepiped formed by vectors {ai }:
ai : r = ti vi , 0 ti 1.
We know that the volume of this parallelepiped is equal to
V ol(ai ) = det ||am
i || ,

(1.82)

where matrix A is defined by expansion of vectors {ai } over orthonormal


basis {ei } ai = em aim . (We know this formula at least for n = 1, 2length of
interval, are of parallelogram).
(Volume vanishes if {ai } is not a basis.)
35

Now consider the scalar product (Riemannian metric) in En in the basis


a1 , . . . , an :
gik = hai , ak i ,
(1.83)
where h , i is scalar product in En : hei , ej i = ij . We see that in (1.83)
X
X
T
2
n
gij = hai , ak i = h
am
e
,
anj en i = am
i mn ai = (A A)ij det G = (det A) ,
i m
m

where G = ||gij ||. Comparing with formula (1.82) we come to Gramm formula:
p
V ol(ai ) = det gik
(1.84)
The matrix G = ||gik || is called Gramm matrix for the vectors {ai }. Gramm
formula states that volume of parallelogram formed by the vectors {a1 , . . . , m }is
equal to the square root of the determinant of Gramm matrix.
Remark One can easy see that formula (1.84) works for arbitrary n-dimensional
parallelogramm in m-dimensional space. Indeed if 1 , . . . , n are just arbitrary n vectors
in m-dimensional Euclidean space then if n < m, the formula (1.82) s failed (matrix A is
m n matrix), but formula (1.84) works. For example the area of parallelogram formed
by arbitrary vectors a1 , a2 in En is equal to
s



 s
g11 g12
ha1 , a1 i ha1 , a1 i
det
.
= det
ha1 , a1 i ha1 , a1 i
g21 g22

One can say that n-dimensional parallelepiped {ai } in new coordinates


ti corresponding to the basis {ai } becomes n-dimensional cube, Standard
Euclidean metric G = dxi ik dxk (in orthonormal basis {ei }) transforms to
G = dxi ik dxk = (aim dtm )ik (akn dtn ) = (AT A)mn dtm dtn
and
Z

Volume{ai } =

Gdt1 . . . dtn = det G = det AT A .

dx . . . dx =
x

1.7.2

0ti 1

Invariance of volume element under changing of coordinates

Check straighforwardly that volume element is invariant under coordinate


transformations, i.e. if y 1 , . . . , y n are new coordinates: x1 = x1 (y 1 , . . . , y n ), x2 =
x2 (y 1 , . . . , y n )...,
xi = xi (y p ), i = 1, . . . , n , p = 1, . . . , n
36

and gpq (y) matrix of the metric in new coordinates:


gpq (y) =

xi
xk
g
(x(y))
.
ik
y p
y q

(1.85)

Then
q
p
det gik (x) dx1 dx2 . . . dxn = det gpq (y) dy 1 dy 2 . . . dy n

(1.86)

This follows from (1.85). Namely


s
 i

p
x
xk
1
2
n
det gik (y) dy dy . . . dy = det
gik (x(y)) q dy 1 dy 2 . . . dy n
y p
y
 i
x
=
Using the fact that det(ABC) = det A det B det C and det y
p
 k
3
det x
we see that from the formula above follows:
y q
s
 i

p
x
xk
1
2
n
det gik (y) dy dy . . . dy = det
gik (x(y)) q dy 1 dy 2 . . . dy n =
y p
y
s
 i 2
p
x
det gik (x(y))dy 1 dy 2 . . . dy n =
det
y p
 i
p
x
det gik (x(y)) det
dy 1 dy 2 . . . dy n =
(1.87)
y p
Now note that
 i
x
det
dy 1 dy 2 . . . dy n = dx1 . . . dxn
p
y
according to the formula for changing coordinates in n-dimensional integral
4
. Hence
 i
p
p
x
1
2
n
det gik (x(y)) det
dy
dy
.
.
.
dy
=
det gik (x(y))dx1 dx2 . . . dxn
y p
(1.88)
Thus we come to (1.86).
3

determinant of matrix does not change if we change the matrix on the adjoint, i.e.
change columns on rows.
 i
x
4
Determinant of the matrix y
of changing of coordinates is called sometimes Jap
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.

37

1.7.3

Examples of calculating volume element

Consider first very simple example: Volume element of plane in Cartesian


coordinates, metric g = dx2 + dy 2 . Volume element is equal to
s


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

If we go to polar coordinates:
x = r cos , y = r sin

(1.89)

Then we have for metric:


G = dr2 + r2 d2
because
dx2 + dy 2 = (dr cos r sin d)2 + (dr sin + r cos d)2 = dr2 + r2 d2
(1.90)
Volume element in polar coordinates is equal to
s


p
1 0
det gdrd = det
drd = rdrd .
0 r2
Lobachesvky plane.
2
2
In coordinates x, y (y > 0) metric G = dx y+dy
, the corresponding matrix
2
 2


1/y
0
G=
. Volume element is equal to det gdxdy = dxdy
.
2
y2
1/y
0
Sphere in stereographic coordinates In stereographic coordinates
G=

4R4 (du2 + dv 2 )
(R + u2 + v 2 )2

(1.91)

(It is isometric to the sphere of the radius R without North pole in stereographic coordinates (see the Homeworks.))
38

Calculate its volume element and volume. It is easy to see that:


!
4R4
0
16R8
2
2
2
2
(R +u +v )
det
g
=
G=
(1.92)
4
4R
(R + u2 + v 2 )4
0
(R2 +u2 +v 2 )2

4 dudv
and volume element is equal to det gdudv = (R4R
2 +u2 +v 2 )2
One can calculate volume in coordinates u, v but it is better to consider
homothety u Ru, v Rv and polar
= Rr cos , v =
coordinates: u4 dudv
4R2 rdrd
Rr sin . Then volume form is equal to det gdudv = (R4R
2 +u2 +v 2 )2 = (1+r 2 )2 .
Now calculation of integral becomes easy:
Z
Z
Z
4R2 rdrd
rdr
du
2
2
V =
= 8R
= 4R
= 4R2 .
2
2
2
2
2
(1 + r )
(1
+
r
)
(1
+
u)
0
0
Segment of the sphere.
Consider sphere of the radius a in Euclidean space with standard Riemannian metric
a2 d2 + a2 sin2 d2
This metric is nothing but first quadratic form on the sphere (see (1.4.3)).
The volume element is
s
 2

p
a
0
det gdd = det
dd = a2 sin dd
0 a2 sin
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:

Z
Z 1 Z 2


2
2
V =
a sin dd =
a sin d d =
0

1 0

2a2 sin d = 2a2 (cos 0 cos 1 ) = 2a(a cos 0 acos1 ) = 2ah

(1.93)
E.g. for all the sphere h = 2a. We come to S = 4a2 . 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)
39

Covariant differentiaion. Connection. Levi


Civita Connection on Riemannian manifold

2.1

Differentiation of vector field along the vector field.


Affine connection

How to differentiate vector fields on a (smooth )manifold M ?


Recall the differentiation of functions on a (smooth )manifold M .

Let X = Xi (x)ei (x) = x


i be a vector field on M . Recall that vector
field 5 X = Xi ei defines at the every point x0 an infinitesimal curve: xi (t) =
xi0 + tX i (More exactly the equivalence class [(t)]X of curves xi (t) = xi0 +
tX i + . . . ).
Let f be an arbitrary (smooth) function on M and X = X i x i . Then
derivative of function f along vector field X = X i x i is equal to
X f = X f = X i

f
xi

The geometrical meaning of this definition is following: If X is a velocity


vector of the curve xi (t) at the point xi0 = xi (t) at the time t = 0 then
the value of the derivative X f at the point xi0 = xi (0) is equal just to the
derivative by t of the function f (xi (t)) at the time t = 0:

dxi (t)
X i (x) x0 =x(0) =
,
dt t=0



d
then X f xi =xi (0) = f xi (t) t=0
dt
(2.1)
Remark In the course of Geometry and Differentiable Manifolds the
operator of taking derivation of function along the vector field was denoted
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 fields
along the vector field.
One can see that the operation X on the space C (M ) (space of smooth
functions on the manifold) satisfies the following conditions:
if

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


)
5

here like always we suppose


Pn by default the summation over repeated indices. E.g.X =
X ei is nothing but X = i=1 X i ei
i

40

hX+gY (f ) = hX (f ) + gY (f ) (linearity over the space of functions)


X (f g) = f X (g) + gX (f ) (Leibnitz rule)
(2.2)
Remark One can prove that these properties characterize vector fields:operator
on smooth functions obeying the conditions above is a vector field. (You will
have a detailed analysis of this statement in the course of Differentiable Manifolds.)
How to define differentiation of vector fields along vector fields.
The formula (2.1) cannot be generalised straightforwardly because vectors at the point x0 and x0 + tX are vectors from different 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
different vector space. One have to define an operation of transport of vectors from the space Tx0 M to the point Tx0 +tX M defining the transport from
the point Tx0 M to the point Tx0 +tX M ).
Try to define the operation on vector fields such that conditions (2.2)
above be satisfied.
2.1.1

Definition of connection. Christoffel symbols of connection

Definition Affine connection on M is the operation which assigns to every


vector field 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 of vector fields on M :
X (Y + Z) = X Y + X Z,

for every , R

(2.3)

(Compare the first condition in (2.2)).


which satisfies the following conditions:
for arbitrary (smooth) functions f, g on M
f X+gY (Z) = f X (Z) + gY (Z)
(compare with second condition in (2.2))
41

(C(M )-linearity)

(2.4)

for arbitrary function f


X (f Y) = (X f ) Y + f X (Y)

(Leibnitz rule)

(2.5)

Recall that X f is just usual derivative of a function f along vector


field: X f = X f .
(Compare with Leibnitz rule in (2.2)).
The operation X Y is called covariant derivative of vector field Y along
the vector field X.
Write down explicit formulae in a given local coordinates {xi } (i =
1, 2, . . . , n) on manifold M .
Let

X = X i ei = X i i Y = Y i ei = Y i i
x
x

The basis vector fields xi we denote sometimes by i sometimes by ei


Using properties above one can see that

(2.6)
X Y = X i i Y k k = X i i Y k k ,
where i = i
Then according to (2.4)


i Y k k = i Y k k + Y k i k
Decompose the vector field i k over the basis i :
i k = m
ik m
and

(2.7)

 Y k (x)
i Y k k =
k + Y k m
ik m ,
xi
Y m (x)
X Y = X i
m + X i Y k m
ik m ,
xi

(2.8)
(2.9)

In components
m

(X Y) = X

Y m (x)
+ Y k m
ik
xi


(2.10)

i
Coefficients {m
ik } are called Christoffel symbols in coordinates {x }. These
coefficients define covariant derivativeconnection.

42

If operation of taking covariant derivative is given we say that the connection is given on the manifold. Later it will be explained why we us the
word connection
We see from the formula above that to define covariant derivative of vector
fields, connection, we have to define Christoffel symbols in local coordinates.
2.1.2

Transformation of Christoffel symbols for an arbitrary connection

Let be a connection on manifold M . Let {ikm } be Christoffel symbols


of this connection in given local coordinates {xi }. Then according (2.7) and
(2.8) we have

Y i
X Y = X m m i + X m imk Y k i ,
x x
x
and in particularly
imk i = m k
Use this relation to calculate Christoffel symbols in new coordinates xi

0 k 0
im0 k0 i0 = m

We have that m0 =
(2.4), (2.5) we have
0
im0 k0 i0

xm0

=


0
= m0 k0 = m

xk
xk0

xm
0
xm

xm
xm0 xm

xk
k
xk0


=

xm
.
xm0 m

xk
xk0

Hence due to properties

0 k +
m
xm0

xk
xk0


k =

2 xk
2 xk
xk xm
k + m0 k0 k
k =
m k +
xm0 xk0
xk0 xm0 m
x x
0

2 xk
xk xm i xi
2 xk xi
xk xm i
0 +

i0
i
k
i
0
0
0
0
0
0
xk xm mk
xm xk
xk xm mk xi
xm0 xk0 xk
Comparing the first and the last term in this formula we come to the transformation law:
If {ikm } are Christoffel symbols of the connection in local coordinates
0
{xi } and {ik0 m0 } are Christoffel symbols of this connection in new local
0
coordinates {xi } then
0

0
ik0 m0

2 xr xi
xk xm xi i
=

+
xk0 xm0 xi km xk0 xm0 xr
43

(2.11)

Remark Christoffel 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 flat connections) then the transformation
 rule
 above is the the
1
same as a transformation rule for tensors of the type
(see the formula
2
(1.1)). In general case this is not true. Christoffel symbols do not transform as tensor under arbitrary non-linear coordinate transformation: see the
second term in the formula above.
Remark On the other hand note that difference of two arbitrary connections is a
i are corresponding Chrstoffel symbols then it follows from (1.1))
tensor. If ikm and
km
i
i transforms as a tensor:
that their difference Skm
= ikm
km
0
Ski 0 m0

0
ik0 m0

0
0

xk xm xi i
xk xm xi  i
i
i0

km km =
S
k 0 m 0 =
xk0 xm0 xi
xk0 xm0 xi km

(See for detail the Homework 4.)

2.1.3

Canonical flat affine connection

It follows from the properties of connection that it is suffice to define connection at vector fields which form basis at the every point using (2.7), i.e.
to define Christoffel symbols of this connection.
Example Consider n-dimensional Euclidean space En with Cartesian
coordinates {x1 , . . . , xn }.
Define connection such that all Christoffel symbols are equal to zero in
these Cartesian coordinates {xi }.
ei ek = m
ik em = 0,

m
ik = 0

(2.12)

Does this mean that Christoffel symbols are equal to zero in an arbitrary
Cartesian coordinates if they equal to zero in given Cartesian coordinates?
Does this mean that Christoffel symbols of this connection equal to zero
in arbitrary coordinates system?
it follows from transformation rules (??) for Christoffel symbols that
0
Christofel symbols vanish also in new coordinates xi if and only if
2 xi
i
i
i k
0
0 = 0, i.e. x = b + ak x
m
i
x x

(2.13)

i.e. the relations between new and old coordinates are linear. We come to
simple but very important
44

Proposition Let all Christoffel symbols of a given connection be equal to


zero in a given coordinate system {xi }. Then all Christoffel symbols of this
0
connection are equal to zero in an arbitrary coordinate system {xi } such that
the relations between new and old coordinates are linear:
0

xi = bi + aik xk

(2.14)
2 i

If transformation to new coordinate system is not linear, i.e. xm0xxi0 6= 0


then Christoffel symbols of this connection in general are not equal to zero in
0
new coordinate system {xi }.
Definition We call connection flat if there exists coordinate system
such that all Christoffel symbols of this connection are equal to zero in a
given coordinate system.
In particular connection (2.12) has zero Christoffel symbols in arbitrary
Cartesian coordinates.
Corollary Connection has zero Christoffel symbols in arbitrary Cartesian
coordinates if it has zero Christoffel symbols in a given Cartesian coordinates.
Hence the following definition is correct:
Definition A connection on the Euclidean space En which Christoffel
symbols vanish in Cartesian coordinates is called canonical flat connection.
Remark Canonical flat connection in Euclidean space is uniquely defined,
since Cartesian coordinates are defined globally. On the other hand on arbitrary
manifold one can define flat connection locally just choosing any arbitrary local
coordinates and define locally flat connection by condition that Christoffel symbols
vanish in these local coordinates. This does not mean that one can define flat
connection globally. We will study this question after learning transformation law
for Christoffel symbols.

Remark One can see that flat connection is symmetric connection.


Example Consider a connection (2.12) in E2 . It is a flat connection.
Calculate Christoffel symbols of this connection in polar coordinates
(
(
p
x = r cos
r = x2 + y 2
(2.15)
= arctan xy
y = y sin

45

Write down Jacobians of transformationsmatrices of partial derivatives:

y

 



x

2
2
x +y
xr yr
cos
sin
rx x
x2 +y 2

=
,
= y
x
x y
r sin r cos
ry y
2
2
x +y
2
2
x +y

(2.16)
According (2.11) and since Chrsitoffel symbols are equal to zero in Cartesian
coordinates (x, y) we have
0

0
ik0 m0

2 xr xi
=
,
xk0 xm0 xr

(2.17)

where (x1 , x2 ) = (x, y) and (x1 , x2 ) = (r, ). Now using (2.16) we have
rrr =
rr = rr =

2 y r
2 x r
+
=0
rr x rr y

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

r =

2 y r
x
y
2 x r
+
= x y = r .
x r y
r
r
rr =

r = r =
=

2 x
2 y
+
= 0.
rr x rr y

2 x
2 y
y
x
1
+
= sin 2 + cos 2 =
r x r y
r
r
r
2 y
y
x
2 x
+
= x 2 y 2 = 0 .
x y
r
r

(2.18)

Hence we have that the covariant derivative (??) in polar coordinates has
the following appearance
r = rr r + r =

r r = rrr r + rr = 0 , ,

,
= r r + = rr (2.19)
r
Remark Later when we study geodesics we will learn a very quick method
to calculate Christoffel symbols.
r = rr r + r =

46

2.1.4

Global aspects of existence of connection

We defined connection as an operation on vector fields obeying the special axioms


(see the subsubsection 2.1.1). Then we showed that in a given coordinates connection is defined by Christoffel symbols. On the other hand we know that in
general coordinates on manifold are not defined globally. (We had not this trouble
in Euclidean space where there are globally defined Cartesian coordinates.)
How to define connection globally using local coordinates?
Does there exist at least one globally defined connection?
Does there exist globally defined flat connection?
These questions are not naive questions. Answer on first and second questions
is Yes. It sounds bizzare but answer on the first question is not Yes 6
Global definition of connection
The formula (2.11) defines the transformation for Christoffer symbols if we go
from one coordinates to another.
Let {(xi ), U } be an atlas of charts on the manifold M .
If connection is defined on the manifold M then it defines in any chart (local
0
coordinates) (xi ) Christoffer symbols which we denote by () ikm . If (xi ), (xi() )
are different local coordinates in a vicinity of a given point then according to (2.11)
0

i0
() k0 m0

i
xk() xm
() x()
0

()i
mk

xi()
xk() xm
()

()

2 xk()
0

xi()
0

k xk
xm
()
() x()

(2.20)

Definition Let {(xi ), U } be an atlas of charts on the manifold M


()i
We say that the collection of Christoffel symbols {km } defines globally a
connection on the manifold M in this atlas if for every two local coordinates
(xi() ), (xi() ) from this atlas the transformation rules (2.20) are obeyed.
Using partition of unity one can prove the existence of global connection constructing it in explicit way. Let {(xi ), U } ( = 1, 2, . . . , N ) be a finite atlas on
the manifold M and let { } be a partition of unity adjusted to this atlas. Denote
by () ikm local connection defined in domain U such that its components in these
()
coordinates are equal to zero. Denote by () ikm Christoffel symbols of this local
6

Topology of the manifold can be an obstruction to existence of global flat connection.


E.g. it does not exist on sphere S n if n > 1.

47

()

connection in coordinates (xi() ) (() ikm = 0). Now one can define globally the
connection by the formula:
0

i
() km (x)

(x)

() i
() km (x)

(x)

xi() 2 xi() (x)


0

xi() xk() xm
()

(2.21)

This connection in general is not flat connection7

2.2

Connection induced on the surfaces

Let M be a manifold (surface) embedded in Euclidean space8 . Canonical flat


connection on EN induces the connection on surface in the following way.
Let X, Y be tangent vector fields to the surface M and can.flat a canonical
flat connection in EN . In general
Z = can.flat
Y
X

is not tangent to manifold M

(2.22)

Consider its decomposition on two vector fields:




can.flat
can.flat
Z = Ztangent + Z , can.flat
, Y = X
Y tangent + X
Y , (2.23)
X
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. Define an induced
connection M on the surface M by the following formula

can.flat
M :
M
Y tangent
(2.24)
X Y : = X
Remark One can imply this construction for an arbitrary connection in
E .
N

2.2.1

Calculation of induced connection on surfaces in E3 .

Let r = r(u, v) be a surface in E3 . Let can.flat be a flat connection in E3 .


Then

can.flat
can.flat
M :
M
Y || = X
Y n(can.flat
Y, n), (2.25)
X Y : = X
X
7

See for detail the text: Global affine connection on manifold in my homepage:
www.maths.mancheser.ac.uk/khudian in subdirectory Etudes/Geometry
8
We know that every n-dimensional manifodl can be embedded in 2n + 1-dimensional
Euclidean space

48

where n is normal unit vector field to M . Consider a special example


Example (Induced connection on sphere) Consider a sphere of the radius
R in E3 :

x = R sin cos
r(, ) : y = R sin sin

z = R cos
then

R cos cos
R sin sin
sin cos
r = R cos sin , r = R sin cos , n = sin sin ,
R sin
0
cos
, r = r(,)
are basic tangent vectors and n is normal unit
where r = r(,)

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


r
= (r )tangent .
=
tangent

Rsin cos
On the other hand one can see that r = Rsin sin = Rn is
R cos
proportional to normal vector, i.e. (r )tangent = 0. We come to
= (r )tangent = 0 = = 0 .

(2.26)

Now calculate and .






r
r
= (r )tangent , =
= (r )tangent
=
tangent
tangent
We have

= = (r )tangent

R cos sin
= R cos cos
.
0
tangent

We see that the vector r is orthogonal to n:


hr , ni = R cos sin sin cos + R cos cos sin sin = 0.
49

Hence

= = (r )tangent = r

R cos sin
= R cos cos = cotan r .
0

We come to
= = cotan = = 0, = = cotan (2.27)
Finally calculate

= (r )tangent

R sin cos
= R sin sin
0
tangent

Projecting on the tangent vectors to the sphere (see (2.25)) we have


= (r )tangent = r nhn, r i =

R sin cos
sin cos
R sin sin sin sin (R sin cos sin cos R sin sin sin sin ) =
0
cos

R cos cos
sin cos R cos sin = sin cos r ,
R sin
i.e.
= sin cos r = sin cos , = = 0 .

2.3
2.3.1

(2.28)

Levi-Civita connection
Symmetric connection

Definition. We say that connection is symmetric if its Christoffel symbols


ikm are symmetric with respect to lower indices
ikm = imk
50

(2.29)

The canonical flat connection and induced connections considered above are
symmetric connections.
Invariant definition of symmetric connection
A connection is symmetric if for an arbitrary vector fields X, Y
X Y Y X [X, Y] = 0

(2.30)

If we apply this definition to basic fields k , m which commute: [k , m ] = 0 we


come to the condition
k m m k = imk i ikm i = 0
and this is the condition (2.29).

2.3.2

Levi-Civita connection. Theorem and Explicit formulae

Let (M, G) be a Riemannian manifold.


Definition. Theorem
A symmetric connection is called Levi-Civita connection if it is compatible with metric, i.e. if it preserves the scalar product:
X hY, Zi = hX Y, Zi + hY, X Zi

(2.31)

for arbitrary vector fields X, Y, Z.


There exists unique levi-Civita connection on the Riemannian manifold.
In local coordinates Christoffel symbols of Levi-Civita connection are given
by the following formulae:


1 ij gjm gjk gmk
i
+ m
.
(2.32)
mk = g
2
xk
x
xj
where G = gik dxi dxk is Riemannian metric in local coordinates and ||g ik || is
the matrix inverse to the matrix ||gik ||.
Proof
Suppose that this connection exists and imk are its Christoffel symbols. Consider vector fields X = m , Y = i and Z = k in (2.31). We have that
m gik = hrmi r , k i + hi , rmk r i = rmi grk + gir rmk .
for arbitrary indices m, i, k.

51

(2.33)

Denote by mik = rmi grk we come to


m gik = mik + mki , i.e.
Now using the symmetricity mik = imk since kmi = kim we have
mik = m gik mki = m gik kmi = m gik (k gmi kim ) =
m gik k gmi + kim = m gik k gmi + ikm = m gik k gmi + (i gkm imk ) =
m gik k gmi + i gkm mik .
Hence
1
1
mik = (m gik + i gmk k gmi ) kim = g kr (m gir + i gmr r gmi ) (2.34)
2
2
We see that if this connection exists then it is given by the formula(2.32).
On the other hand one can see that (2.32) obeys the condition (2.33). We
prove the uniqueness and existence.
since i k = m
ik m .

Consider examples.
2.3.3

Levi-Civita connection of En

For Euclidean space En in standard Cartesian coordinates


GEucl = (dx1 )2 + + (dxn )2 = ik dxi dxk
Components of metric are constants (they are equal to 0 or 1). Hence obviously Christoffel symbols of Levi-Civita connection in Cartesian coordinates
according formula (2.32) vanish:
Ikm = 0 in Cartesian coordiantes
Recalling canonical flat connection (see 2.1.3) we come to simple but important observation:
Observation Levi-Civita connection coincides with canonical flat connection in Euclidean space En . They have vanishing Cristoffel symbols in
Cartesian coordinates.

52

2.3.4

Levi-Civita connection on 2-dimensional Riemannian manifold with metric G = adu2 + bdv 2 .

Example Consider 2-dimensional manifold with Riemannian metrics



 

g11 g12
a(u, v)
0
2
2
G = a(u, v)du + b(u, v)dv ,
G=
=
g21 g22
0
b(u, v)
Calculate Christoffel symbols of Levi Civita connection.
Using (2.34) we see that:
111

= 21 (1 g11 + 1 g11 1 g11 ) = 12 1 g11 =

1
a
2 u

211 = 121

= 21 (1 g12 + 2 g11 1 g12 ) = 12 2 g11 =

1
a
2 v

221

= 21 (2 g12 + 2 g12 1 g22 ) = 12 1 g22 =

1
b
2 u

112

= 21 (1 g12 + 1 g12 2 g11 ) = 12 2 g11 =

1
a
2 v

122 = 212

= 21 (2 g21 + 1 g22 2 g21 ) = 12 1 g22 =

1
b
2 u

222

= 21 (2 g22 + 2 g22 2 g22 ) = 12 2 g22 =

1
b
2 v

(2.35)

To calculate ikm = g ir kmr note that for the metric a(u, v)du2 + b(u, v)dv 2
!
 11 12 
1
0
g
g
= a(u,v)
G1 =
1
g 21 g 22
0
b(u,v)
Hence
111 = g 11 111 =

au
,
2a

211 = g 22 112 =

av
,
2b

2.3.5

g 11 121 =

av
,
2a

122 = g 11 221 =

bu
2a

221 = 212 = g 22 122 =

bu
,
2b

222 = g 22 222 =
(2.36)

bv
2b

121 = 112 =

Example of the sphere again

Calculate Levi-Civita connection on the sphere.


On the sphere first quadratic form (Riemannian metric) G = R2 d2 +
2
R sin2 d2 . Hence we use calculations from the previous example with
53

a(, ) = R2 , b(, ) = R2 sin2 (u = , v = ). Note that a = a = b = 0.


Hence only non-trivial components of will be:


b
sin 2
R2 sin 2

=
=
,
=
,
(2.37)
2a
2
2


b
cos
R2 sin 2

=
=
(2.38)
= =
2b
sin
2
All other components are equal to zero:
= = = = = 0
Remark Note that Christoffel symbols of Levi-Civita connection on the
sphere coincide with Christoffel 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 E3

We know already that canonical flat connection of Euclidean space is the LeviCivita connection of the standard metric on Euclidean space. (see section
2.3.3.) Now we show that Levi-Civita connection on surfaces in Euclidean
space coincides with the connection induced on the surfaces by canonical flat
connection. We perform our analysis for surfaces in E3 .
Let M : r = r(u, v) be a surface in E3 . Let G be induced Riemannian
metric on M and LeviCivita connection of this metric.
We know that the induced connection (M ) is defined in the following
way: for arbitrary vector fields X, Y tangent to the surface M , M
X Y equals
can.flat
to the projection on the tangent space of the vector field X
Y:

can.flat
M
Y
=

Y
,
X
X
tangent
where can.flat is canonical flat connection in E3 (its Christoffel symbols
vanish in Cartesian coordinates). We denote by Atangent a projection of
the vector A attached at the point of the surface on the tangent space:
A = A n(A, n) , (n is normal unit vector field to the surface.)

54

Theorem Induced connection on the surface r = r(u, v) in E3 coincides


with Levi-Civita connection of Riemannian metric induced by the canonical
metric on Euclidean space E3 .
Proof
Let M be induced connection on a surface M in E3 given by equations
r = r(u, v). Considering this connection on the basic vectors rh , rv we see
that it is symmetric connection. Indeed
M
u
u
v
v
M
u v = (ruv )tangent = (rvu )tangent = v u . uv = vu , uv = vu .

Prove that this connection preserves scalar product on M . For arbitrary


tangent vector fields X, Y, Z we have
flat
flat
X hY, ZiE3 = hcan.
Y, ZiE3 + hY, can.
ZiE3 .
X
X

since canonical flat connection in E3 preserves Euclidean metric in E3 (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
Atangent is its projection on the tangent space to the surface, then for every tangent vector B scalar product hA, BiE3 equals to the scalar product
hAtangent , BiE3 = hAtangent , BiM since vector A Atangent is orthogonal to
the surface. Hence we deduce from (2) that X hY, ZiM =


flat
can. flat
M
h can.
Y tangent , ZiE3 +hY, X
Z tangent iE3 = hM
X
X Y, ZiM +hY, X ZiM .
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 (2.32).
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.

3
3.1
3.1.1

Parallel transport and geodesics


Parallel transport
Definition

Let M be a manifold equipped with affine connection .


55

Definition Let C : x(t), t0 t t1 be a curve on the manifold M ,


starting at the point p0 = x(t0 ) and ending at the point p1 = x(1 ) ((with
coordinates xi = xi (t))). Let X = X(t0 ) be an arbitrary tangent vector
attached at the initial point p0 = x0 (with coordinates xi (t0 )) of the curve
C, i.e. X(t0 ) Tp0 M is a vector tangent to the manifold M at the point
p0 with coordinates xi (t0 ). (The vector X is not necessarily tangent to the
curve C)
We say that X(t), t0 t t1 is a parallel transport of the vector X(t0 )
T0 M along the curve C : xi = xi (t), t0 t t1 if
For an arbitrary t, t0 t t, vector X = X(t), (X(t)|t=t0 = X(t0 )) 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
= v X = 0 .
(3.1)
dt
In components: if X m (t) are components of the vector field X(t) and
v m (t) are components of the velocity vector v of the curve C ,
X(t) = X m (t)

|x(t) ,
xm

v=

dx(t)
dxi
=
|x(t)
dt
dt xm

then the condition (3.1) can be rewritten as


dX i (t)
+ v k (t)ikm (xi (t))X m (t) 0 .
dt

(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 flat connection then in Cartesian
coordinates Christoffel symbols vanish and parallel transport is nothing but
dX
= v X = 0, X(t) is constant vector.
dt
Remark Compare this definition of parallel transport with the definition 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 E3 and define parallel transport by the condition, that only
orthogonal component of the vector changes during parallel transport, i.e.
dX(t)
is a vector orthogonal to the surface
dt
56

3.1.2

Parallel transport is a linear map. Parallel transport with


respect to Levi-Civita connection

We usualy consider parallel transport on Riemannian manifold with respect


to Levi-Civita connection. If (M, G) is Riemannian manifold then we consider parallel transport with respect to connection which is Levi-Civita
connection of the Riemannian metric G.
Consider again curve C : x = x(t), t0 t t1 on manifold M starting
at the point p0 and ending at the point p1 (see above). Let X T0 be an
arbitrary tangent vector at the point 0 , and X(t) be parallel transport (3.1)
of this vector along the curve C:

X(t) t=t0 = X ,

X(t)
= 0.
dt

Taking value of X(t)


at the final point p1 of the curve C we come to the new
0
vector X = X(t) t=t1 tangent to the manifold M at the point p1 . Thus we
define the map between tangent vectors at the initial point p0 of the curve
C and tangent vectors at the ending point p1 of this curve:
PC :

Tp0 M 3 X PC (X) = X0 Tp1 M .

(3.3)

Sure this map depends on the curve C which joins starting and ending points
(if we are not in Euclidean space).
Proposition
Let C be a an arbitrary curve with starting point p0 and ending point
p1 . Then the map (3.3) defines linear operator PC which does not depend
on parameterisation of the curve9 :
PC (X1 + X2 ) = PC (X1 ) + PC (X2 ) .

(3.4)

In the case if connection is Levi-Civita connection, then PC is an orthogonal operator: for two arbitrary vectors X, Y T0 M
hX, Yip0 = hX0 , Y0 ip1

X0 = PC (X) Tp1 M ,

Y0 = PC (Y) Tp1 M ,
(3.5)

we consider parameterisatons with the same initial and ending points, i.e. all reparameterisations do not change orientation. One can say that linear operator PC is an
operator defined for orineted curve, since we fix initial and ending poitns of the curve.

57

where as usual h , ip is the scalar product at the point p.


In particular the length of the vector is preserved during parallel transport.
Proof The fact that it is a linear map follows immediately from the fact
that differential equations (3.2) are linear. If vector fields X(t), Y(t) are
covariantly constant along the curve C, i.e. they obey differential equation
(3.2), then their linear combination X(t) + Y(t) obeys this equation, also,
This implies (3.4).
The fact that the map (3.3) does not depend on the parameterisation
(if it does not change the orinetation) also follows from differential equation
(3.2). Indeed let t = t( ), 0 1 , t(0 ) = t0 , t(1 ) = 1 be another
parameterisation of the curve C, which does not change parameterisation, i.e.
initial and ending points of the curve do not intechange. Then multiplying
dt
and using the fact that velocity v0 ( ) = t v(t) we
the equation (3.2) on d
come to differential equation:
dX i (t( ))
+ v 0k (t( ))ikm (xi (t( )))X m (t( )) 0 .
d

(3.6)

The functions X ( t( )) with the same initial conditions are the solutions
of this equation.
It remains to prove that PC is orthogonal operator.
It follows immediately from the definition (3.1) of a parallel transport and
the definition (2.31) of Levi-Civita connection that during parallel transport
the scalar product hX(t), Y(t)ix(t) is preserved:
d
hX(t), Y(t)i = v hX(t), Y(t)i = hv X(t), Y(t)i+hX(t), v Y(t)i = h0, Y(t)i+hX(t), 0i = 0 .
dt
(3.7)
This implies (3.5).

3.2
3.2.1

Geodesics
Definition. Geodesic on Riemannian manifold

Let M be manifold equipped with connection .


Definition A parameterised curve C : xi = xi (t) is called geodesic if
i
velocity vector v(t) : v i (t) = dxdt(t) is covariantly constant along this curve,

58

i.e. it remains parallel along the curve:


v v =

v
dv i (t)
=
+ v k (t)ikm (x(t))v m (t) = 0, i.e.
dt
dt

(3.8)

d2 xi (t) dxk (t) i


dxm (t)
(x(t))

= 0.
(3.9)
+
km
dt2
dt
dt
These are linear second order differential equations. One can prove that
this equations have solution and it is unique10 for an arbitrary initial data
(xi (t0 ) = xi0 , x i (t0 ) = x i0 . )
In other words the curve C : x(t) is a geodesic if parallel transport of
velocity vector along the curve is a velocity vector at any point of the curve.
Geodesics defined with Levi-Civita connection on the Riemannian manifold 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 above (see equation (3.4), (3.5) and (3.7)) 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 (2.31)) in particularly the length of
the velocity vector v:
Example 1 Geodesics of Euclidean space. In Cartesian coordinates
Christoffel symbols of Levi-Civita connection vanish, and differential equation (3.8), (3.9) are reduced to equation
dxi (t)
d2 xi (t)
= 0,
= v i xi = xi0 + v i t .
2
dt
dt

(3.10)

We come to straight lines.


Example 2 Geodesics of cylindrical surface One can see that if Riemannian metric G = Gik dui dv k have constant coefficients in coordinates ui then
Christoffel symbols of Levi-Civita connection vanish in these coordinates,
10

this is true under additional technical conditions which we do not discuss here

59

(see formula (2.32)) and according to (3.10) geodesics are straight lines in
coordinates ui . In particular
this is a case for cylinder: If surface of cylin
x = a cos
der is given by equation y = a sin
then Riemannian metric is equal to

z=h
2
2
2
G = a d + dh and we come to equations:
(
(
( 2
d(t)
(t)
d (t)
=

= 0 + t
=
0
2
dt
dt
dt
dh(t)
dh(t)
.
(3.11)
d2 h(t)
=c
= h0 + tc
=0
dt
dt
dt2
In general case we come to helix:

x = a cos (t) = a cos (0 + t)


y = a sin (t) = a sin (0 + t)

z = h(t) = h0 + ct

(3.12)

If c = 0 then geodesics are circles x2 + y 2 = a2 , z = h0 . If angular velocity


= 0 then geodesics are vertical lines x = x0 , y = y0 , z = h0 + ct.
3.2.2

Geodesics and Lagrangians of free particle on Riemannian


manifold.

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 differential equations


L
d L
(3.13)
=
i
dt x
xi
In detail


d L
2L m
2 L m
L
x

+
x =
.
(3.14)
=
i
m
i
m
i
dt x
x x
x x
xi
These equations are called Euler-Lagrange equations of the Lagrangian
L. We will explain later the variational origin of these equations 11 .
11

To every mechanical system one can put in correspondence a Lagrangian on configuration 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.

60

Lagrangian of free particle


Let (M, G), G = gik dxi dxk be a Riemannian manifold.
Definition We say that Lagrangian L = L(x, x)
is the Lagrangian of a a
free particle on the Riemannian manifold M if
L=

gik x i x k
2

(3.15)

Example Free particle in Euclidean space. Consider E3 with standard


metric G = dx2 + dy 2 + dz 2
gik x i x k
x 2 + y 2 + z 2
L=
=
2
2

(3.16)

Note that this is the Lagrangian that describes the dynamics of a free particle.
Example A free particle on a sphere.
The metric on the sphere of radius R is G = R2 d2 + R2 sin2 d2 . Respectively for the Lagrangian of free particle we have
gik x i x k
R2 2 + R2 sin2 2
L=
=
2
2

(3.17)

Equations of geodesics and Euler-Lagrange equations


Theorem. Euler-Lagrange equations of the Lagrangian of a free particle
are equivalent to the second order differential equations for geodesics.
This Theorem makes very easy calculations for Christoffel indices.
This Theorem can be proved by direct calculations.
Calculate Euler-Lagrange equations (3.13) for the Lagrangian (3.15):


gmk x m x k




2
d
d L
= d gik x k = gik x k + gik x m x k
=
dt x i
dt
x i
dt
xm
and
L
=
xi

gmk x m x k
2

xi

1 gmk m k
x x .
2 xi

Hence we have


d L
gik
L
1 gmk m k

= gik x k + m x m x k =
=
x x ,
i
i
dt x
x
x
2 xi
61

i.e.

gik x k + m gik x m x k = i gmk x m x k .


2

1
m k
m k
Note that m gik x x = 2 m gik x x + k gim x m x k . Hence we come to
equation:
d 2 xk 1
gik 2 + (m gik + k gim i gmk ) x m x k
dt
2
Multiplying on the inverse matrix g ik we come


d2 xi 1 ij gjm gjk gmk dxm dxk
= 0.
(3.18)
+ g
+ m
dt2
2
xk
x
xj
dt dt

We recognize here Christoffel symbols of Levi-Civita connection (see (2.32))


and we rewrite this equation as
d2 xi dxm i dxk
+

= 0.
dt2
dt mk dt

(3.19)

This is nothing but the equation (3.8).


Applications of this Theorem: calculation of Christoffel symbols of LeviCivita connection.
3.2.3

Calculations of Christoffel symbols and geodesics using the


Lagrangians of a free particle.

It turns out that equation (3.19) is the very effective tool to caluclatie
Christoffel symbols of Levi-Civita connection.
Consider two examples: We calculate Levi-Civita connection on sphere
in E3 and on Lobachevsky plane using Lagrangians and find geodesics.
1) Sphere of the radius R in E3 :
Lagrangian of free particle on the sphere is given by (3.17):
L=

R2 2 + R2 sin2 2
2

Euler-Lagrange equations defining geodesics are


 



d L
L
d
2

=
R R2 sin cos 2 sin cos 2 = 0 ,
dt

dt
(3.20)
62

 

d L
L
d

=
R2 sin2 = 0 + 2cotan = 0 .
dt

dt
Comparing Euler-Lagrange equations with equations for geodesic in terms of
Christoffel symbols:

+ 2 + 2 + 2 = 0,

+ 2 + 2 + 2 = 0
we come to
= = = 0 , = sin cos ,

(3.21)

= = 0, = = cotan .

(3.22)

(Compare with previous calculations for connection in subsections 2.2.1 and


2.3.4)
We know already and we will prove later in a elegant way that geodesics on the sphere
are great circles. (see subsection 3.2.8 below). Consider another technically more difficult
but straightforward proof of this fact. To find geodesics one have to solve second order
differential equations (3.19)
One can see that the great circles: = 0 , = 0 + t are solutions of second order
differential equations (3.20) with initial conditions






(t) t=0 = 0 , (t)
= 1, (t) t=0 = 0 , (t)
= 0.
(3.23)
t=0
t=0
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 t1 be geodesic. Rotating the sphere we can come to the curve = 0 (t), =
0 (t), 0 t t1 such that velocity vector at the initial time is direccted along meridian,
i.e. initial conditions are




0 (t) t=0 = 0 , 0 (t) t=0 = a, 0 (t) t=0 = 0 , 0 (t) t=0 = 0 .
(3.24)
(Compare with initial conditions (3.23)) Second order differential equations with boundary
conditions for coordinates and velocities at t = 0 have unique solution. The solutions of
second order differential equations (3.20) with initial conditions (3.24) is a curve 0 (t) =
0 + at, 0 (t) = 0 . It is great circle. Hence initial curve the geodesic = (t), = (t),
0 t t1 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 =
dx2 +dy 2
is
y2
1 x 2 + y 2
L=
.
2 y2
63

Euler-Lagrange equations are


L
d L
d
=0=
=
x
dt x
dt

L
x 2 + y 2
d
d L
=
=
=
3
y
y
dt y
dt

x
y2


x
2x y
2x y

= 2 3 , i.e. x
= 0,
y
y
y


y
y
2y 2
x 2 y 2

= 0.

,
i.e.
y
+
y2
y2
y3
y
y
i

Comparing these equations with equations for geodesics: x x k ikm x m = 0


(i = 1, 2, x = x1 , y = x2 ) we come to
1
1
1
xxx = 0, xxy = xyx = , xyy = 0, yxx = , yxy = yyx = 0, yyy = .
y
y
y
In a similar way as for a sphere one can find 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.

See another examples in Homework 6 and 7.


3.2.4

Magnitudes preserved along geodesicsIntegrals of motion

It is very useful to find magnitudes which are preserved along geodesics,


functions F = F (x, x)
such that for geodesic C : xi = xi (t) the magnitude
dI(t)
= 0.
dt
(3.25)
Geodesics are solutions of equations of motions for the Lagrangian of a free
i k
for an
particle L = gik (x)2 x x . One can consider such functions F = F (x, x)
i
arbitrary Lagrangian L. In this case x (t) is a solution of the Lagrangian L.
These magnitudes which are preserved along solutions of equations of
motion (in particular along geodesics in the case if L is the Lagrangian of a
free particle) are called integrals of motion (See in detail about integrals of
motion in Appendix to this lectures).
There is the following very useful criterion to find magnitudes, which are
preserved on equations of motions, i.e. integrals fo motion.
Proposition Let Lagrangian L(xi , x i ) in coordinates {xi } does not depend, say on the coordinate x1 . L = L(x2 , . . . , xn , x 1 , x 2 , . . . , x n ). Then the
function
L
F1 (x, x)
=
x 1

I(t) = F (x, x)
xi = xi (t) is preserved along geodesic xi (t),

64

is integral of motion. (In the case if L(xi , x i ) does not depend on the coordiL
nate xi . the function Fi (x, x)
= x
i will be integral of motion.)
Proof is simple. Check the condition (3.25): Euler-Lagrange equations of
motion are:


d L
L
i = 0 (i = 1, 2, . . . , n)
i
dt x
x
L
In particular for first coordinate x1 , x
1 = 0 and




d L
L
d L
1 =
= 0,
1
1
dt x
x
dt x
L
i.e. the magnitude I(t) = F (x, x)
is preserved if F = x
1 . We see that
exactly first equation of motion is


d L
d
L
= 0 since x
= F1 (q, q)
1 = 0, .
1
dt x
dt

(if L(xi , x i ) does not depend on the coordinate xi then the function
is integral of motion since i th equation is exactly the condiFi (x, x)
= L
x 1

tion Fi = 0.)
L
The integral of motion Fi = x
i is called sometimes generalised momentum.
Consider examples of calculation of preserved mangnitudes along geodesics.
Example (sphere)
3
Sphere of the
metric: G = Rd2 +R2 sin2 d2
 radius R in E . Riemannian

and Lfree = 21 R2 2 + R2 sin2 2 . Lagrangian does not depend explicitly
on coordinate . The integral of motion is
F =

Lfree
= R2 sin2 .

It is preserved along geodesics, i.e. along great circles.


Example (cone)

x = kh cos
Consider cone y = kh sin

z=h

. Riemannian metric:

G = d(kh cos )2 + d(kh sin )2 + (dh)2 = (k 2 + 1)dh2 + k 2 h2 d2 .


65

and free Lagrangian


Lfree

(k 2 + 1)h 2 + k 2 h2 2
.
=
2

Lagrangian does not depend explicitly on coordinate h. The integral of motion is


Lfree
F =
= k 2 h2 .


It is preserved along geodesics.
Remark One has to note that for the Lagrangian of a free particle F =
L = gik x i x k , kinetik energy, is integral of motion preserved along geodesic:it
is nothing that square of the length of velocity vector which is preserved
along the geodesic.
See these and other examples in Homework 6.
Using integral of motions to calculate geodesics
Integrals of motions may be very useful to calculate geodesics. The equations for
geodesics are second order differential equations. If we know integrals of motions they
help us to solve these equations. Consider just an example.
2
2
For Lobachevsky plane the free Lagrangian L = x 2y+2y . We already calculated geodesics
in the subsection 3.3.4. Geodesics are solutions of second order Euler-Lagrange equations
2
2
for the Lagrangian L = x 2y+2y (see the subsection 3.3.4)
(
x

y+

2x y
y =0
y 2
x 2
y y

=0

It is not so easy to solve these differential equations.


For Lobachevsky plane we know two integrals of motions:
E=L=

x 2 + y 2
,
2y 2

and F =

x
L
= 2.
x
y

These both integrals are preserved in time: if x(t), y(t) is geodesics then
(
(
x(t)

F = y(t)
x = C1 y 2
2
p

x(t)
2 +y(t)
2
y = 2C2 y 2 C12 y 4
E = 2y(t)2 = C2

(3.26)

(3.27)

These are first order differential equations. It is much easier to solve these equations in
general case than initial second order differential equations.

66

3.2.5

Variational principe and Euler-Lagrange equations

Here very briefly we will explain how Euler-Lagrange equations follow from
variational principe.
Let M be a manifold (not necessarily Riemannian) and L = L(xi , x i ) be
a Lagrangian on it.
x ,t
Denote my Mx21 ,t21 the space of curves (paths) such that they start at the
point x1 at the time t = t1 and end at the point x2 at the time t = t2 :
x ,t

Mx21 ,t21 = {C : x(t), t1 t t2 , x(t1 ) = x1 , x(t2 ) = x2 } .

(3.28)

x ,t

Consider the following functional S on the space Mx21 ,t21 :


Z

t2

S [x(t)] =


L xi (t), x i (t) dt .

(3.29)

t1
x ,t

for every curve x(t) Mx21 ,t21 .


This functional is called action functional.
Theorem Let functional S attaints the minimal value on the path x0 (t)
x2 ,t2
Mx1 ,t1 , i.e.
x ,t

x(t) Mx21 ,t21

S[x0 (t)] S[x(t)] .

(3.30)

Then the path x0 (t) is a solution of Euler-Lagrange equations of the Lagrangian L:




L
d L
if x(t) = x0 (t) .
(3.31)
=
i
dt x
xi
Remark The path x(t) sometimes is called extremal of the action functional (3.29).
We will use this Theorem to show that the geodesics are in some sense
shortest curves 12 .
12

The statement of this Theorem is enough for our purposes. In fact in classical mechanics another more useful statement is used: the path x0 (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 (3.29), i.e.
S[x0 (t) + x(t)] S[x0 (t) + x(t)] = 0(x(t))
for an arbitrary infinitesimal variation of the path x0 (t): x( t1 ) = x( t2 ) = 0.

67

(3.32)

3.2.6

Un-parameterised geodesic

We defined 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?
Definition We call un-parameterised curve geodesic if under suitable
parameterisation it obeys the equation (3.8) 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. Parallel
transport does not depend on parametersiation of the curve (see subsection
3.1.2, equation (3.3)). Choose a suitable parameterisation xi = xi (t) such
that xi (t) obeys the equations (3.8) for geodesics, i.e. the velocity vector v(t)
is covariantly constant along the curve: v v = 0. If A(t0 ) = cv(t0 ) at the
given point p (c is a scalar coefficient) 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). In the
c(t)v(t) choose a reparameterisation t = t( ) such that dt(
d
dt( )
0
new parameterisation the velocity vector v ( ) = 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 following important observation:
Let C is un-parameterised geodesic, xi (t) be its arbitrary parameterisation and v(t) be velocity vector in this parameterisation. Then the velocity
vector remains parallel to the curve since it is a tangent vector.
68

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 xi = xi (t) is geodesic in an arbitrary parameterisation and s = s(t) is a natural parameter (which defines the length
of the curve) then xi (t(s)) is parameterised geodesic.
3.2.7

Parallel transport of vectors along geodesics

We already now that during parallel transport along curve with respect to
Levi-Civita connection scalar product of vectors, i.e. lengths of vectors and
angle between them does not change (see subsection 3.1.2, equations (3.3)
and (3.5)). This remark makes easy to calculate parallel transport of vectors
along geodesics in Riemannian manifold. Indeed let C a geodesic (in general
un-parameterised) and a vectors X(t) is attached to the point p1 C on
the curve C. In the special case if X is a tangent vector to geodesic C then
during parallel transport it remains tangent, i.e. proportional to velocity
vector:
X(t) = a(t)v(t) .
(3.33)
and r = r(t) is an arbitrary parameterisation of geodesic C.
Here v(t) = dr(t)
dt
Note that in general t is not parameter such that r = r(t) is parameterised
geodesic; t is an arbitrary parameter. In the special case if t is a parameter
such that r = r(t) is parameterised geodesic then velocity vector remains
velocity vector during parallel transport, i.e. X(t) = av(t) where a is not
dependent on t.
To calculate the dependence of coefficient a on t in (3.34) we note that
the length of the vector is not changed (see the section ??, i.e.
hX(t), X(t)i = ha(t)v(t), a(t)v(t)i = a2 (t)|v(t)|2 = constant
3.2.8

(3.34)

Geodesics on surfaces in E3

Let M : r = r(u, v) be a surface in E3 . Let GM be induced Riemannian


metric and a Levi-Civita connection on M . We consider on M Levi-Civita
connection of the metric GM .
the velocity vector. AcLet C be an arbitrary geodesic and v(t) = dr(t)
dt
cording to the definition of geodesic v v = 0. On the other hand we know
that Levi-Civita connection coincides with the connection induced on the
69

surface by canonical flat connection in E3 (see the Theorem in subsection


2.4). Hence

can.flat
v v = 0 = M
v tangent
(3.35)
v v = v
2

= a.
In Cartesian coordinates vcan.flat v = v v = dtd v(u(t), v(t)) = d dtr(t)
2
Hence according to (3.35) the tangent component of acceleration equals
to zero.
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 (3.35) v v = 0.
We come to very beautiful observation:
Theorem 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

x = a cos
Let r(h(t), (t)) be a geodesic on the cylinder y = a sin . We have

z=h

a sin
a cos and for acceleration:
=
v = dr
dt
h

a sin

a cos

a 2 cos
dv
a=
=
+ a 2 sin
dt

0
h
|
{z
}
|
{z
}
normal
acceleration
tangent acceleration
2

Since tangential acceleration equals to zero hence ddt2h = 0 and h(t) = h0 + ct


Normal acceleration is centripetal acceleration of the rotation over circle
with constant speed (projection on the plane OXY ). The geodesic is helix.
(Compare these calculations with calculations of geodesics of cylinder in the
last example of section 3.2.1: see (3.12).)
Geodesics on sphere
70

x = a sin cos
Let r = r((t), (t)) be a geodesic on the sphere of the radius a: r(, ) : 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
d
M(t) =
(r(t) v(t)) = (v(t) v(t)) + (r(t) a(t)) = 0
dt
dt

(3.36)

We have M(t) = M0 . 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
M0 = 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 surface.
Remark The vector M = r(t) v(t) is the torque. The torque is integral of motion in isotropic space.This is the core of the considerations for
geodesics on the sphere.
3.2.9

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 x1 and x2 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 x1 C. Then for an arbitrary point x2 C
which is close to the point x1 the arc of geodesic joining the points x1 , x2 is
a shortest curve between these points13 .
13
More precisely: for every point x1 C there exists a ball B (x1 ) such that for an
arbitrary point x2 C B (x1 ) the arc of geodesic joining the points x1 , x2 is a shortest
curve between these points.

71

This Theorem makes a bridge between two different 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 Lfree =
gik (x)x i x k
and the length Lagrangian
2
q
p
Llength (x, x)
= gik (x)x i x k = 2Lfree .
If C : xi (t), t1 t t2 is a curve on M then

Length of the curve C =


Z t2 q
t2
i
i
Llength (x (t), x (t))dt =
gik (x(t))x i (t)x k (t)dt .

t1

(3.37)

t1

The proof of the Theorem follows from the following observation:


Observation Euler-Lagrange equations for the length functional (3.37) are equivalent
to the Euler-Lagrange equations for action functional (3.29). 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 functional. We showed before
that Euler-Lagrange equations for action functional (3.29) define geodesics. Hence the
shortest curves are geodesics.
One can check the observation p
by direct calculation:
Calculate Euler-Lagrange equa
tions for the Lagrangian Llength = gik (x)x i x k = 2Lfree :
!


d Llength
gkm x k x m
d
1
1
Llength
k
p
p
=
g
x

ik
dt
x i
xi
dt
xi
gik x i x k
2 gik x i x k
d
=
dt

Lfree
i
Llength x
1

1 Lfree
= 0.
Llength xi

(3.38)

To facilitate calculations note that the length functional (3.37) is reparameterisation


invariant. Choose the natural parameter s(t) or a parameter proportional to the natural
parameter on the curve xi (t). We come to Llength = const and it follows from (3.38) that


 


d Llength
Llength
1
d Lfree
Lfree

=
= 0.

i
dt
x i
xi
Llength dt
xi
x
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.
72

3.2.10

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 find geodesics
using the fact that they are shortest. The fact that geodesics are locally the
shortest curves
Consider again sphere in E3 with the radius R: Coordinates , , induced
Riemannian metrics (first quadratic form):
G = R2 (d2 + sin2 d2 ) .

(3.39)

Consider two arbitrary points A and B on the sphere. Let (0 , 0 ) be coordinates of the point A and (1 , 1 ) be coordinates of the point B
Let CAB be a curve which connects these points: CAB : (t), (t) such
that (t0 ) = 0 , (t1 ) = 1 , (t0 ) = 0 , (t1 ) = 1 then:
Z q
(3.40)
LCAB = R t2 + sin2 (t)2t dt
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:
(t0 ) = (t1 ) = 0 , (t0 ) = (t1 ) = 0 . Compare its length with the length
of the meridian which connects the points A, B:
Z t1 p
Z t1
Z t1 q
2
2
2
2
t dt = R
t dt = R(1 0 ) (3.41)
R t + sin t dt R
t0

t0

t0

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=

dx2 + dy 2
y2
73

(3.42)

The length of the curve : x = x(t, y = y(t)) is equal to


Z s 2
xt + yt2
L=
dt
y 2 (t)
In particularly the length of the vertical interval [1, ] tends to infinity if
0:
Z 1r
Z s 2
xt + yt2
1
1
dt =
dt = log
L=
2
2
y (t)
t

One can see that the distance from every point to the line y = 0 is equal to
infinity. This motivates the fact that the line y = 0 is called absolute.
Consider two points A = (x0 , y0 ), B = (x1 , y1 ) 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 = x0 . Let CAB be an arc of
the ray x = x0 which joins these points: CAB : x = x0 , y = y0 + t Then
it is easy to see that the length of the curve CAB is less or equal than the
length
curve x = x(t), y = y(t)
which joins these points:
of the arbitrary


x(t) t=0 = x0 , y(t) t=0 = y0 , x(t) t=t1 = x0 , y(t) t=t1 = y1 :
Z ts
0

x2t

yt2

+
dt
2
y (t)

Z ts
0

yt2
dt =
y 2 (t)

y1

y0

dt
y1
dt = log
= length of CAB
t
y0

Hence CAB is shortest. We prove that vertical rays are geodesics.


Consider now the case if x0 6= x1 . 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 coordinates (r, ):
x = a + r cos , y = r sin
(3.43)
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, dx2 + dy 2 = dr2 + r2 d2 . Hence:
G=

dx2 + dy 2
dr2 + r2 d2
d2
dr2
=
==
+
y2
r2 sin2
sin2 r2 sin2

74

(3.44)

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:
Z t1 s 2
Z t1 s 2
t
t
rt2
+ 2 2 dt
dt =
(3.45)
LAB =
2
sin r sin
sin2
t0
t0
Z

t1

t0

t
dt =
sin

d
tan 1
= log
sin
tan 1

The proof is finished.

Surfaces in E3

Now equipped by the knowledge of Riemannian geometry we consider surfaces in E3 . We reconsider again conceptions of Shape (Weingarten) operator, Gaussian and mean curvatures, focusing attention on the the fact what
properties are internal and what properties are external. In particular we
consider again gaussian curvature and derive its internal meaning. We will
consider again Theorema Egregium.

4.1

Parallel transport of the vector. Formulation of


result.

We formulate here very important theorem about parallel transport of vectors


over closed curve and deduce Theorema Egregium form this theorem.
We will consider also the formula for Gaussian curvature in isothermal
coordinates14
4.1.1

Theorem of parallel transport over closed curve

Let M be a surface in Euclidean space E3 . Consider a closed curve C on M ,


M : r = r(u, v), C : r = r(u(t, v(t)), 0 t t1 , x(0) = x(t1 ). (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: Recall that we did it in the section 3.1.2 where we
already considered parallel transport over curve for arbitrary connection and
for Levi-Civita connection for arbitrary Riemannian manifold (see equations
14

This year this is not compulsory

75

(3.3) and (3.5)). Now we will repeat these considerations for this special
case.)




xi


.
= X (t)r r(u(t),v(t)) , r =
X(t) = X (t)
u u (t)
u xi
|
{z
}
|
{z
}
Internal observer External observer
X(t) :

X(t)
= 0, 0 t t1 ,
dt

i.e.
du (t)
dX (t)
+ X (t) (u(t))
= 0, 0 t t1 ,
dt
dt

, , = 1, 2 , (4.1)

where is the connection induced on the surface M by canonical flat


connection (see (2.25)), or (it is the same) the Levi-Civita connection (2.32)
of the induced Riemannian metric on the surface M and its Christoffel
symbols:


1 g g g
xi xi

= g
,
where
g
=
hr
,
r
i
=


2
u
u
u
u u
(4.2)

are components of induced Riemannian metric GM = g du du /


Let r(0) = p be a starting (and ending) point of the curve C: r(0) =
r(t1 ) = p. The differential equation (4.1) defines the linear operator
PC : Tp M Tp M

(4.3)

For any vector X Tp M , its image the vector RC X as


the solution of the
differential equation (4.1) with initial condition X(t) t=0 = X. (See also
section 3.1.2, equations (3.3) and (3.5)). )
On the other hand we know that parallel transport is orthogonal operator,
it does not change the scalar product of two vectors, and it does not chnage
lengths of vectors (see (3.5) in the subsection 3.1.2):
hX, Xi = hPC X, PC Xi

(4.4)

We see that PC is an orthogonal operator in the 2-dimensional vector space


Tp M . We know that orthogonal operator preserving orientation is the operator of rotation on some angle .
76

One can see that PC preserves orientation 15 then the action of operator
PC on vectors is rotation on the angle, i.e. the result of parallel transport
along closed curve is rotation on the . This angle depends depends on
the curve. The very beautiful question arises: How to calculate this angle
(C)
Theorem Let M be a surface in Euclidean space E3 . 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
Z
= (X, PC X) =
Kd ,
(4.5)
D

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.
Remark Here we assume that a curve C is smooth curve. E.g. if C is a smooth closed
geodesics then it follwos from this Theorem and properties of geodesics that rotation angle
is equal to 2n (where n is integer). In
Pgeneral if C is piecewise smooth curve, then one
can see that rotation angle is equal to
i (n 1) where n is number of smooth arcs,
and i angles between them.

Example Consider the closed curve, latitude C0 : = 0 on the sphere


of the radius R. Calculations show that
(C0 ) = 2(1 cos 0 )

(4.6)

(see also the Homework 8). On the other hand the latitude C0 is the
boundary of the segment D with area 2RH where H = R(1cos 0 ). Hence
Z
1
2RH
= 2 area of the segment =
Kd
(X, RC X) =
R2
R
D
since Gaussian curvature is equal to R12
(See the proof of this Theorem in Appendix (section 6.2.6)).
15

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 by continuity arguments
that operator RC preserves an orientation.

77

4.1.2

Gau Theorema Egregium

We defined Gaussian curvature in terms of Weingarten (shape) operator as


a product of principal curvatures. This definition was in terms of External
Observer. On the other hand the right hand side of the formula (??) depends
on metric on the surface, i.e. Gaussian curvature maybe independently calculated by Internal Observer. We come to remarkable corollary:
Corollary Gau Egregium Theorema
Gaussian curvature of the surface can be expressed in terms of induced
Riemannian metric. It is invariant of isometries.
We may come to the same corollary form the Theorem about transport
over closed curve. Indeed let D be a small domain around a given point p,
let C its boundary and (D) be an 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
Z
if S(D) 0 then (D) =
Kd K(p)S(D), i.e.
D

(D)
.
(4.7)
S(D)0 S(D)
Now notice that left hand side od this equation defining Gaussian curvature K(p) depends only on Riemannian metric on the surface C. Indeed
numerator of LHS is defined by the solution of differential equation (4.1)
which depends on Levi-Civita connection depending on the induced Riemannian metric, and denominator is an area depending on Riemannian metric
too. Thus we come again to GauTheorema Egregium.
K(p) = lim

In next subsections we develop the technique which itself is very interesting. One of the applications of this technique is the proof of the Theorem
(4.5) and Theorem (??). Thus we will prove Theorema Egregium too.
Later in the fifth section we will give another proof of the Theorema
Egregium.
4.1.3

Formula for Gaussian curvature in isothermal (conformal)


coordinates

Now we will consider one very beautiful and illuminating formula to calculate
Gaussian curvature for surfaces.
78

Let M : r = r(u, v) be a surface in E3 .


Definition We say that coordinates (parameters) u, v are isothermal (or
conformal) if the induced Riemannian metric 1.4.2 is equal to
G = (u, v)(du2 + dv 2 )
Consider examples. For example if (u, v) are stereographic coordinates for
sphere of radius R we know that
G=

4R4 (du2 + dv 2 )
(R2 + u2 + v 2 )2 )

i.e. stereographic coordiantes are conformal coordiantes:


G = (u, v)(du2 + dv 2 ) with =

4R4
.
(R2 + u2 + v 2 )2

Another example: an arbitrary locally Euclidean surface, i.e. surface with


induced Riemannian metric G = du2 + dv 2 in some local coordinates u, v.
One can show that locally one can always find conformal (isothermal)
coordinates on surface in E3 16
Theorem Let surface M is given in conformal coordinates: r = r(u, v)
such that induced Riemannian metric is equal to G = (u, v)(du2 + dv 2 ).
Then Gaussian curvature K of the surface M is given by the formula
1 (log )
2
2
(4.8)
,
where = u
2 + v 2 .
2

In particular this formula states that Gaussian curvature is expressed in terms


of Riemannian metric. This implies GauTheorema Egregium.
If u, v are conformal coordinates, G = (du2 + dv 2 ), then it is often
convenient to introduce a function (u, v) such that = log , i.e. =
e(u,v) . Then
G = e(u,v) (du2 + dv 2 ) .
(4.9)
K=

Then formula (4.8) will have the following appearance:




1
1 2 (u, v) 2 (u, v)
K = e () = e
+
.
2
2
u2
v 2

(4.10)

What about existence of conformal (isothermal) coordinates? Proposition For surface M in E3


16

The existence of local isothermal coordiantes is a part of famous Gauss theorem, which
can be formulated in modern terms in the following way: every surface has a canonical
complex structure (z = u + iv, z = u iv). We will consider this question later.

79

in a vicinity of an arbitrary point there exist isothermal coordinates i.e. coordinates


such that induced metric G = e (du2 + dv 2 ) = e dzd
z.
If (u, v) and (u0 , v 0 ) are two arbitrary isothermal coordinates then the function
z = f (w) is holomorphic function or anti-holomorphic function,
We denote u + iv = z, u iv = z. Recall that if z = u + iv then




1

F
1

F
Fz =
=
i
F, and Fz =
=
+i
F
z
2 u
v
z
2 u
v

(4.11)

Function F = f + ig is holomorphic Fz = 0 (fu + igu ) + i(fv + igv ) = 0 fu = gv


and fv = gv (Cauchy Riemann conditions). Function F = f + ig is anti-holomorphic
Fz = 0 (fu + igu ) i(fv igv ) = 0. E.g. F = z 2 = (u + iv)2 = u2 v 2 + 2iuv is
holomorphic function, F = ez = euiv = eu (cos v i sin v) is anti-holomorphic function
(See also ??).
This Proposition immediately implies the following important corollary-Theorem:
Two-dimensional surface in E3 has canonical complex structure, i.e. one can consider
a canonical atlas of local complex charts such that transition funtctions are analytic.
Idea of Proof of this Proposition
Let (, ) be arbitrary parameters of surface and G = Ad 2 + 2Bdd + Cd 2 (g11 =
A, g12 = g21 = B, g22 = C). The positive-definiteness of the metric implies that G =
where = df + idg is 1-form. Use the fact that an arbitrary 1-form up to a mulitplier
dF .
function is an exact form: = dF . We come to isothermal coordinates: G = dF
To prove the second part of Proposition we just perform straightforward calculation. Let
G = e (du2 + dv 2 ) = e dzd
z in local coordinates z = u + iv, and in new local coordinates
0
0
w = u0 + iv 0 G = e (du02 + dv 02 ) = e dwdw,
where Let w = F (z). Then

G = e dzdz = e (Fw dw + Fw dw)
Fw dw + Fw dw
=


e Fw Fw dw2 + |Fw |2 + |Fw |2 | dwdw
+ Fw Fw dw
2
(4.12)
The condition that new coordinates are isothermal too means that Fw Fw = 0 , i.e. Fw = 0,
i.e. F is anti-holomorphic function or Fw = 0, i.e. F is holomorphic function.
Illustrate this idea on the example: Let G = dx2 + f 2 (x)dy 2 be a metric on a domain
of Riemannian manifold (e.g. for sphere x = , y = , f (x) = sin2 x, for cone x = h, y =
, f (x) = x). Then G = (dx + if dy)()dx if (x)dy). For 1-form = dx + if (x)dy we have
that dx + if (x)dy = f (x) (dG(x) + idy) = f (x)d (L(x) + iy), where L(x) is antiderivative
of a function f1 and dx2 + f 2 (x)dy 2 = f 2 (x)d (L(x) + iy) d (G(x) iy) = e (du2 + dv 2 ),
where e = f 2 (x), u = L(x), v = y 17

4.2

Derivation formula

Let M be a surface embedded in Euclidean space E3 , M : r = r(u, v).


17

in general case we use essentially the condition of analiticity. This proof was done by
Gauss. The general smooth case was proved only in the beginning of XX century.

80

Let e, f , n be three vector fields defined 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 surface18 .
Vector fields 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
f
f
n
n
e
du +
dv, df =
du +
dv , dn =
du +
dv
u
v
u
v
u
v

These 1-forms take values in the vectors in E3 , i.e. they are vector valued
1-forms. Any vector in E3 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.
e
e
for de = u
du+ e
dv expanding vectors u
and e
over basis vectors we come
v
v
to
e
e
= A1 (u, v)e+B1 (u, v)f +C1 (u, v)n,
= A2 (u, v)e+B2 (u, v)f +C2 (u, v)n
u
v
thus
de =

e
e
du +
dv = (A1 e + B1 f + C1 n) du + (A2 e + B2 f + C2 n) dv =
u
v

= (A1 du + A2 dv) e + (B1 du + B2 dv) f + (C1 du + C2 dv) e,


|
{z
}
|
{z
}
|
{z
}
M11

M12

(4.13)

M11

i.e.
de = M11 e + M12 f + M13 n,
where M11 , M12 and M13 are 1-forms on the surface M defined by the relation
(4.13).
In the same way we do the expansions of vector-valued 1-forms df and
dn we come to
de = M11 e + M12 f + M13 n
df = M21 e + M22 f + M23 n
dn = M31 e + M32 f + M33 n
18

One can say that {e, f , n} is an orthonormal basis in Tp E3 at every point of surface
p M such that {e, f } is an orthonormal basis in Tp E3 at every point of surface p M .

81

This equation can be rewritten in the following way:




e
M11 M12 M13
e

f
d f = M21 M22 M23
n
M31 M32 M33
n

(4.14)

Proposition The matrix M in the equation (4.14) is antisymmetrical matrix, i.e.


M11 = M22 = M33 = 0
M12 = M21 = a
(4.15)
M13 = M31 = b
M23 = M32 = b
i.e.



e
0
a b
e
d f = a 0 c f ,
n
b c 0
n

(4.16)

where a, b, c are 1-forms on the surface M .


Formulae (4.16) are called derivation formula.
Prove this Proposition. (Here I give the detailed proof, but later in Remark, very short proof in condensed notations.)
Recall that {e, f , n} is orthonormal basis, i.e. at every point of the surface
he, ei = hf , f i = hn, ni = 1, and he, f i = he, ni = hf , ni = 0
Now using (4.14) we have
he, ei = 1 dhe, ei = 0 = 2he, dei = he, M11 e + M12 f + M13 ni =
M11 he, ei + M12 he, f i + M13 he, ni = M11 M11 = 0 .
Analogously
hf , f i = 1 dhf , f i = 0 = 2hf , df i = hf , M21 e+M22 f +M23 ni = M22 M22 = 0 ,
hn, ni = 1 dhn, ni = 0 = 2hn, dni = hn, M31 e+M32 f +M33 ni = M33 . M33 = 0 .
We proved already that M11 = M22 = M33 = 0. Now prove that M12 =
M21 , M13 = M31 and M13 = M31 .
he, f i = 0 dhe, f i = 0 = he, df i + hde, f i =
82

he, M21 e + M22 f + M23 ni + hM11 e + M12 f + M13 n, f i = M21 + M12 = 0.


Analogously
he, ni = 0 dhe, ni = 0 = he, dni + hde, ni =
he, M31 e + M32 f + M33 ni + hM11 e + M12 f + M13 n, ni = M31 + M13 = 0
and
hf , ni = 0 dhf , ni = 0 = hf , dni + hdf , ni =
hf , M31 e + M32 f + M33 ni + hM21 e + M22 f + M23 n, ni = M32 + M23 = 0.
Remark This proof may be be performed much more shortly in condensed notations. Derivation formula (4.16) in condensed notations are
dei = Mik ek

(4.17)

Orthonormality condition means that hei , ek i = ik . Hence


dhei , ek i = 0 = hdei , ek i+hei , dek i = hMim em , ek i+hei , Mkn en i = Mik +Mki = 0
(4.18)
Much shorter, is not it?
4.2.1

Gauss condition (structure equations)

Derive the relations between 1-forms a, b and c in derivation formula.


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
ddf = 0 and the fact that for 1-form = 0.) We have from derivation
formula (4.16) that
d2 e = 0 = d(af + bn) = daf a df + dbn b dn =
da f a (ae + cn) + db n b (be cf ) =
(da + b c)f + (a a + b b)e + (db a c)n = (da + b c)f + (db + c a)n = 0 .
We see that
(da + b c)f + (db + c a)n = 0

(4.19)

Hence components of the left hand side equal to zero:


(da + b c) = 0 (db + c a) = 0 .
83

(4.20)

Analogously
d2 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
4.2.2

(4.21)

Geometrical meaning of derivation formula. Weingarten operator (shape oeprator) in terms of derivation formula.

Let M be a surface in E3 .
Let e, f , n be three vector fields defined 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 formula (4.16).
Induced Riemannian metric
If G is the Riemannian metric induced on the surface M then since e, f
is orthonormal basis at every tangent space Tp M then
G(e, e) = G(f , f ) = 1, G(e, f ) = G(f , e) = 0
The matrix of the Riemannian metric in

1
G=
0

the basis {e, f } is



0
1

(4.22)

(4.23)

Induced connection Let be the connection induced by the canonical flat connection
on the surface M .
Then according equations (2.25) and derivation formula (4.16) for every tangent vector
X
X e = (X e)tangent = (de(X))tangent = (a(X)f + b(X)n)tangent = a(X)f .

84

(4.24)

and
X f = (X f )tangent = (df (X))tangent = (a(X)e + c(X)n)tangent = a(X)e .

(4.25)

In particular
e e = a(e)f
f e = a(f )f
e f = a(e)e f f = a(f )e

(4.26)

We know that the connection is Levi-Civita connection of the induced Riemannian


metric (4.24) (see the subsection 4.2.1)19 .
Second Quadratic form Second quadratic form is a bilinear symmetric function A(X, Y)
on tangent vectors which is well-defined 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 formula (4.16)
we have
A(e, e) = he e, ni = hde(e), ni = ha(e)f + b(e)n, ni = b(e) ,
A(f , e) = hf e, ni = hde(f ), ni = ha(f )f + b(f )n, ni = b(f ) ,
A(e, f ) = he f , ni = hdf (e), ni = ha(e)f + c(e)n, ni = c(e) ,
A(f , f ) = hf f , ni = hdf (f ), ni = ha(f )f + c(f )n, ni = c(f ) ,
The matrix of the second quadratic form in the basis {e, f } is

 

A(e, e) A(f , e)
b(e) b(f )
A=
=
A(e, f ) A(f , f )
c(e) c(f )

(4.28)

This is symmetrical matrix (see the subsection 4.3.2):


A(f , e) = b(f ) = A(e, f ) = c(e) .

(4.29)

Weingarten (Shape) operator


Let S be Weingarten operator: SX = X n (see the subsection 6.4 in
Appendix, or Geometry lectures). Then it follows from derivation formula
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
19

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 .

85

(4.27)

and the matrix of the Weingarten operator in the basis {e, f } is




b(e) b(f )
S=
c(e) c(f )

(4.30)

Remark According to the condition (4.29) the matrix S is symmetrical. The relations
A = GS, S = G1 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.

4.2.3

Gaussian and mean curvature in terms of derivation formula

Now we are equipped to express Gaussian and mean curvatures in terms of


derivation formula. Using (4.30) we have for Gaussian curvature
K = det S = b(e)c(f ) c(e)b(f ) = (b c)(e, f )

(4.31)

and for mean curvature


H = Tr S = b(e) + c(f )

(4.32)

What next? We will study in more detail formula (4.31) later.


Now consider some examples of calculation of Weingarten operator, e.t..c.
for using derivation formula.

4.3
4.3.1

Calculations with use of derivation formulae


Examples of calculations of derivation formulae and curvatures for cylinder, cone and sphere

Last year we calculated Weingarten operator and curvatures for cylinder,


cone and sphere (see also the subsection 6.4 in Appendices.). Now we do the
same but in terms of derivation formula.
Cylinder
We have to define three vector fields e, f , n on the points of the cylinder
surface x2 + y 2 = a2 :

x = a cos
r(h, ) :
(4.33)
y = a sin

z=h
86

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 fields rh , r and normal unit
vector field:

0
a sin
cos
rh = 0 , r = a cos , n = sin .
(4.34)
1
0
0
Vectors rh , r and n are orthogonal to each other but not all of them have
unit length. One can choose


0
sin
cos
r
cos , n = sin
=
(4.35)
e = rh = 0 , f =
a
0
0
1
These vectors form an orthonormal basis and e, f form an orthonormal basis
in tangent space.
Derive for this basis derivation formula (4.16). For vector fields e, f , n in
(4.35) we have

sin
cos
de = 0, df = d cos = sin d = nd,
0
0

cos
sin
dn = d sin = cos = f d,
0
0
i.e.



e
e
0
a b
e
0 0
0

f = 0 0 d
f ,
d f = a 0 c
n
b c 0
n
0 d
0
n
i.e. in derivation formula 1-forms a, b vanish a = b = 0 and c = d.
The matrix of Weingarten operator in the basis {e, f } is


 
 
b(e) c(e)
0 d(e)
0 0
S=
=
=
b(f ) c(f )
0 d(f )
0 R1
87

(4.36)

According to (4.31) and(4.32) Gaussian curvature K = b(e)c(f )b(e)c(f ) = 0


and mean curvature
r 
1

=
H = b(e) + c(f ) = d(f ) = d
R
a
Remark We denote by the same letter a the radius of the cylinder surface
(4.33) and 1-form a in derivation formula. I hope that this will not lead to
the confusion. (May be it is better to denote the radius of the cylindrical
surface by the letter R.)
Cone
For cone:

x = kh cos
r(h, ) :
,
y = kh sin

z=h

kh sin
k cos
cos
1
sin
rh = k sin , r = kh cos , n =
2
1
+
k
0
1
k

Tangent
vectors rh , r are orthogonal to each other. The length of the vector rh 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 rh , r :

k cos
sin
cos
rh
1
k sin , f = r = cos , n = 1
sin
e=
=
hk
1 + k2
1 + k2
1 + k2
1
0
k
Calculate de, df and dn:

k cos
sin
kd
cos = kd f ,
de = d k sin =
2
1+k
1 + k2
1
0

sin
cos
df = d cos = sin d =
0
0

k cos
cos
k
d d sin = kd e d n ,
k
sin

1 + k2
1 + k2
1 + k2
1 + k2
1
k
and

cos
sin
1
d
cos .
dn =
d sin =
1 + k2
1
+ k2
k
0

88

We come to

e
0
d f = a
n
b


0
a b
e
kd

0 c
f =
1+k2
c 0
n
0

kd
1+k2

0
d
1+k2

d f ,
2
1+k
n
0
0

(4.37)

i.e. in derivation formula for 1-forms a = kd


, b = 0 and and c = d
.
1+k2
1+k2
Remark Note that calculation of df are little bit hard. On the other hand the answer
for df follows from answers for de and dn since the matrix in (4.37) is antisymmetric. So
we can omit the straightforward calculations of df .
The matrix of Weingarten operator in the basis {e, f } is
! 




0 d(e)
0
0
b(e) c(e)
2
1+k
.
S=
=S=
=
)
0 kh1
b(f ) c(f )

0 d(f
1+k2
1+k2
r 
1
1
since d(f ) = d kh
d( ) = kh
.
= kh
According to (4.31), (4.32) Gaussian curvature

K = b(e)c(f ) b(e)c(f ) = 0
and mean curvature
H = b(e) + c(f ) = d(f ) = d

r 

.
kh 1 + k 2

Sphere
For sphere

x = R sin cos
(4.38)
r(, ) :
y = R sin sin

z = R cos

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

R sin
0

sin cos
r
n(, ) =
= sin sin .
R
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 R sin . Hence we

89

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

cos cos
sin
sin cos
r
r
r
= cos sin , f (, ) =
= cos , n(, ) =
= sin sin .
e(, ) =
R
R sin
R
sin
0
cos
Calculate de, df and dn:

cos cos
de = d cos sin =
sin

cos sin
cos cos
cos cos d cos sin d = cos df dn,
0
sin

sin
cos
df = d cos = sin d =
0
0

cos cos
sin cos
cos d cos sin sin d sin sin = cos de sin dn ,
sin
cos

sin cos
cos cos
sin sin
dn = d sin sin = cos sin d + sin cos d
cos
sin
0
= de + sin df .
i.e.



e
0
a b
e
0
cos d
d
e
0
sin d f ,
d f = a 0 c f = cos d
n
b c 0
n
d
sin d
0
n
(4.39)
i.e. in derivation formula a = cos d, b = d, c = sin d.
Remark The same remark as for cone: equipped by the properties of
derivation formula we do not need to calculate df . The calculation of de and
dn and the property that the matrix in derivation formula is antisymmetric
gives us the answer for df .
90

The matrix of Weingarten operator in the basis {e, f } is



 
  1

0
b(e) c(e)
d(e) sin d(e)
R
S=
=
=
0 R1
b(f ) c(f )
d(f ) sin d(f )


since d(e) = d R = R1 d( ) = R1 , d(e) = d R = R1 d( ) = 0.
According to (4.31) and(4.32) Gaussian curvature
K = b(e)c(f ) b(e)c(f ) =

1
R2

and mean curvature

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 ).
Mean curvature is define up to a sign. If we change n n mean
curvature H R1 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.
H = b(e) + c(f ) =

4.3.2

Curvatures for surface z = F (x, y)

.
Now using derivation formulae we calculate curvature for arbitrary surface
z = F (x, y) and later we will calculate curvatures for surfaces in conformal
(isometric coordinates).
We will caclulate curvature not at an arbitrary point but only at the
points of extrema of function F (x, y). (In fact this condition is not very
demanding.)
Derivation formulae become very useful tool for solving these questions
20
.
A surface z = F (x, y) can be defined by parameterisation

x = u
r(u, v) :
y=v

z = F (u, v)
20

In the previous section we calculated curvatures of cylinder, and sphere using derivation formulae. These calculations may be even easier to perform using just usual methods
which we studied in the course of Geometry.

91

Conisder coordinate vector fields of the surface




1
0

ru = 0
rv = 1 ,
,
u
v
Fu
Fv
and unit normal vector field

Fu
1
Fv .
n(u, v) = p
2
2
1 + Fu + Fv
1
It is obviously orthogonal to ru , rv and
it has
unbit length.
1
One can see that e = |rruu | = 1 2 0 is unit vector field tangent to
1+Fu
Fu
0
rv
1

surface. The vector field |rv | = 2 1 is also tangent to surface, it is


1+Fv
Fv
also unit, but in general it is not orthogonal to vector field e, hru , rv i = Fu Fv .
To find a second tangent vector field orthgonal to e we may consider the
vector field f which is vector product f = n e:


Fu
1
1
1

Fv p
0=
f =ne= p
2
2
2
1 + Fu + Fv
1 + Fu F
1
u

Fu Fv
1
1 + Fv2
p
(1 + Fu2 + Fv2 )(1 + Fu2 )
Fv

Since vector field n is orthogonal to surface, and it is unit, hence vector


fields {e, f , n} form orthonormal basis, e, f are tangent. Thus we constructed
orthonormal basis {e, f , n} attached to the surface. We want to calculate
curvatures at the origin a point p with coordinates u = v = 0 Put the
following condition: the surface z = F (x, y) has extremum at the origin, i.e.


Fu p = Fv p = 0 , (p it has coordinates u = v = 0).
(4.40)
This conidition is not demanding. For every point A on the surface on
the surface one can find adjusted Cartesian coordinates such that in these
92

coordinates this surface will have extremum at the point A. On the other
hand this condition drastically simplifies calculations. Note that if condition
(4.40) is obeyed then at the point p vector fields e, f , n look in a very simple
way:



1
0
0





e p = e(u, v)u=v=0 = 0 , f p = f (u, v)u=v=0 = 1 , n p = n(u, v)u=v=0 = 0 ,


0
0
1
Now obtain derivation formulae (at the point p) We will calculate everything
just at the point p. Note that if condition (4.40) is obeyed then

1
0

1
0 = 0
= (ndFu )p ,
de p = d p
1 + Fu2 F
dF
u

u=v=0

F
F
0
u
v

1
1 + Fv2 = 0
= (ndFv )p ,
df p = d p
2
2
2
(1 + Fu + Fv )(1 + Fu )
Fv
dFu u=v=0
p

and

Fu
0

1
Fv = 0
dn p = d p
= (edFu f dFv )p ,
1 + Fu2 + Fv2
1
dF
u u=v=0
p
since all other terms vanish at u = v = 0. Comparing with derivation formula


e
0
a b
e
d f = a 0 c d f
n
b c 0
n
we see that forms a, b, c at origin are equal to





a p = 0 , b p = dFu p = (Fuu du + Fuv dv)p , c p = dFv p = (Fvu du + Fvv dv)p ,
Now calculate the values of these forms on vectors e, f at origin. We have
that b(e) = dFu (e)p = (Fuu du + Fuv dv)(ru ) = Fuu . Analogously b(f ) =
dFu (f )p = (Fuu du+Fuv dv)(rv ) = Fuv , c(e) = dFv (e)p = (Fvu du+Fvv dv)(ru ) =
Fvu , and c(f ) = dFv (f )p = (Fvu du + Fvv dv)(rv ) = Fvv ,
93

Hence we have that matrix of Weintegarten (shape) operator at the origin


is equal to

 

b(e) c(e)
Fuu Fuv
2
S=
=
, K = det S = Fuu Fvv Fuv
, H = Tr S = Fuu +Fvv
b(f ) c(f )
Fvu Fvv p
Theorem For surface z = F (x, y), Wengarten (Shape) operator in extremum point is
given by quadratic form (Hessian) of function.

Example
Consider surface defined by equation z = Ax2 +2Bxy +y 2 , The point x =
y = 0 is extremum point. All derivatives Fu = 2Au + 2Bv, Fv = 2Bu + 2Cv
vanish t origin.
Then Gaussian curvature at point x = y = 0 is equal to
2
K = Fxx Fyy Fxy

Gaussian curvature at arbitrary point of surface z = F (x, y) is equal to


K=

4.3.3

2
Fxx Fyy Fxy
2
2
(1 + Fx + Fy )3/2

Proof of the Theorem on curvature of surfaces given in


conformal coordinates using derivation formulae

We return here to subsection 4.1.3 where we formulated the Theorem about


Gaussian curvature(??) of surface r = r(u, v) in E3 defined in conformal
coordinates. Let (u, v) be local conformal coordinates, and metric G =
(u, v)(du2 + dv 2 ) = e (du2 + dv 2 ).
Consider vectors

e = e 2

,
u

f = e 2

,
v

n=ef.

It is evident that {e, f , n} form orthonoromal basis:


he, ei = 1 , he, f i = 0 , he, ni = 0, hf , f i = 1 hf , ni = 0 , hn, ni = 1 .
Consider derivation formula (4.16) for

e
0
d f = a
n
b

this basis:

e
a b
0 c f ,
c 0
n

94

(4.41)

To calculate Gaussian curvature we need to calculate 1-form a in this equations since accordind equations (4.31) and (4.21) K = b c(e, f ) and da +
b c = 0, i.e. K = da(e, f ). Now calculate 1-form a. We have
 
de = d e 2 ru = af + bn .
Taking scalar product of this equation of f we come to

 

a = hde, f i = d e 2 ru , e 2 rv .

(4.42)

Calculate it. Since hru , rv i = 0 then

 
d e 2 ru , e 2 rv = e hdru , rv i = e hruu , rv idu + e hruv , rv idv .
Now using the fact that hrv , rv i = hru , ru i = e and hru , rv i = 0 calculate
hruv , rv i and hruu , rv i. We have
hruv , rv i =

 1
1
1
hrv , rv i =
e = u e
2 u
2 u
2

and
hruu , rv i =

1
1 
1
hru , rv ihru , ruv i = 0hru , ruv i =
hrv , rv i =
e = v e
u
2 v
2 v
2

Hence we see that 1-form a in (4.42) is equal to

   
1
a = hde, f i = d e 2 ru , e 2 rv
= e hruu , rv idu+e hruv , rv idv = (u dvv du) ,
2
(4.43)
and 2-form


1
1
1
(u dv v du) = (uu dudvvv dvdu) = (uu + vv ) dvdu .
da = d
2
2
2
Now using Gauss formula (4.21) and (4.31) we come to


1

K = b c(e, f ) = da(e, f ) = (uu + vv ) du dv e 2


,e 2
=
2
u
v




e
e
e 2
(uu + vv ) du dv
,
=
(uu + vv ) =
,

2
u v
2
2
95

where Laplacian = u
2 + v 2 .
It is useful to write down this fiormula in complex coordinates Write down
the formula in holomorphic coordinates: z = u + iv, z = u iv. We have
that





2
e 2
e

2

K=

+
i

=
2e
,
+
2
u2 v 2
2
u
v
u
v
zz
(4.44)

(for definition of z and z see (4.11)). This expression is sometimes very


convenient for calculations.

Example Consider sphere of radius 1 in stereographic coordinates. Then G =


4dzd
z

In complex coordinates G = (1+z


z with e = (1+z4 z)2 , i.e.
z)2 = e dzd
2
= log 4 2 log(1 + z z). We see that z = 1+zz and zz = (1+z2 z)2 , i.e. K =
2e zz = 1.
Exercise Let z = f (w) be an holomorphic changing of complex coordinates. Due to
Theorem new coordinates u0 , v 0 (w = u0 + iv 0 , z = u + iv ) are isothermal coordinates too:
If
0
0
G = e (du2 + dv 2 ) = e dzd
z = e dwdw
= e (du02 + dv 02 ) .
4(du2 +dv 2 0
(1+u2 +v 2 )2 ) .

It is very illuminating to check straightforwardly that calculating of Gaussian curvature


in new coordinates we will come to the same answer. Do it. According to (4.12) we see
that e dzd
z = e fw fw dwdw,
i.e. 0 = + log fw + log fw . Hence

2 0
2
2
=
+
log fw + log fw .
ww

ww

ww

Notice that the function log fw is holomorphic function w log fw = 0 and the function

log fw is anti-holomorphic function w


log fw = 0 too. Hence

2
log fw + log fw = 0 .
ww

This implies that


2 0
2
=
.
ww

ww

Again using the fact that functions z = f (w) and zw = fw are holomorphic functions we
see that
2 0
2

z
=
=
(z fw ) =
fw = zz fw fw .
ww

ww

Finally we come to
0

K = 2e

2 0
2
= 2elog fw log fw zz fw fw = 2e
.
ww

zz

Thus we check by straightforward calculations that Gaussian curvature remains the same.In
these calculations In these calculations we used intensively properties (4.11) of holomorphic
and anti-holomorphic functions.

96

Curvature tensor

5.1

Curvature tensor for connection

Definition-Proposition Let manifold M be equipped with connection .


Consider the following operation which assigns to arbitrary vector fields X, Y
and Z on M the new vector field:

R(X, Y)Z = X Y Y X [X,Y] Z
(5.1)
This operation is obviously linear over the scalar coefficients.
One can show that this operation is C (M )-linear with respect to vector
fields X, Y Z, i.e. for an arbitrary functions f, g, h
R(f X, gY)(hZ) = f ghR(X, Y)Z .

(5.2)
 
1
This means that the operation defines the tensor field of the type
: If
3
X = X i i , X = X i i , X = X i i then according to (5.2)
i
R(X, Y)Z = R(X m m , Y n n )(Z r r ) = Z r Rrmn
X mY n
i
the components of the tensor R in the coordinate
where we denote by Rrmn
basis i
Rirmn i = R(m , n )r
(5.3)

This tensor is called curvature tensor of the connection .


Express components of the curvature tensor in terms of Christoffel symbols of the connection. If m n = rmn r then according to the (5.1) we
have:
Rirmn i = R(m , n )r = m n r n m r ,
Rirmn = m (pnr p ) n (pmr p ) =
m inr + imp pnr n imr inp pmr .

(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 + hY Z) Y (X hZ + hX Z) [X,Y] hZ h[X,Y] Z =
97

X Y hZ + Y hX Z + X hY Z + hX Y Z
Y X hZ X hY Z Y hX Z + hY X Z
[X,Y] hZ h[X,Y] Z =

 

h X Y Z Y X Z) [X,Y] Z + X Y h Y X h [X,Y]h Z =
hX 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 Rikmn is expressed trough derivatives of Christoffel symbols. In spite


this fact it is is much more pleasant object than Christoffel symbols, since
the latter is not the tensor.
i
It follows from the definition that the tensor Rkmn
is antisymmetrical
with respect to indices m, n:
i
i
Rkmn
= Rknm
.

(5.5)

One can prove that for symmetric connection this tensor obeys the following identities:
i
i
Rikmn + Rmnk
+ Rnkm
= 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 Christoffel symbols vanish in a vicinity of a given
point p in some chosen coordinate system then in general Christoffel symbols
do not vanish in arbitrary coordinate systems. E.g. Christoffel symbols of
canonical flat connection in E2 vanish in Cartesian coordinates but do not
vanish in polar coordinates. This unpleasant property of Christoffel symbols
is due to the fact that Christoffel symbols do not form a tensor.
In particular if a tensor vanishes in some coordinate system, then it vanishes in arbitrary coordinate system too. This implies very simple but important
Proposition If curvature tensor Rikmn vanishes in some coordinate system, then it vanishes in arbitrary coordinate systems.
We see that if Christoffel symbols vanish in a vicinity of a given point
p in some chosen coordinate system then its Riemannian curvature tensor

98

vanishes in a vicinity of the point p (see the formula (5.4)) and hence it
vanishes locally (in a vicinity of point p) in arbitrary coordinate system.
In fact one can prove
Theorem If a connection is symmetric then curvature tensor vanishes in
a vicinity of a point if and only there exist local Cartesian coordiantes,n a
vicinity of this point i.e. coordinates in which Christoffel symbol of connection vanish.

5.2

Riemann curvature tensor of Riemannian manifolds.

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 Riemann curvature tensor, or Riemann tensor.
Using Riemannian metric one can consider Riemann tensor with all low
indices
Rikmn = gij Rjkmn
(5.7)
In the subsection above we formulated very important Theorem that vanishing of curvature tensor means that connection is locally flat. For Riemann tensor one can formulate the analogous Theorem. If Riemannina manifold is locally Euclidean, i.e. there exist coordinates (x1 , . . . , xn ) such that
G = (dx1 )2 + + (dxn )2 then it is evident that Christoffel symbols of
Levi-Civita connection vanish in these coordinates, hence curvature tensor
vanishes also. The converse implication is true also:
Theorem Riemann curvature tensor vanishes if and only if Riemannian
i
manifold is locally Euclidean, i.e. if Rkmn
0 in a vicinity of the point
p of Riemannian manifold, then in a vicinity of this point there exist local
coordinates (x1 , . . . , xn ) such that Riemannian metric G = (dx1 )2 + +
(dxn )2 .
For Riemann tensor one can consider Ricci tensor,
i
Rmn = Rmin

(5.8)

which is symmetrical tensor: Rmn = Rnm .


One can consider scalar curvature:
R = Rikin g kn = g kn Rkn
99

(5.9)

where g kn is Riemannian metric with indices above (the matrix ||g ik || is


inverse to the matrix ||gil ||).
Ricci tensor and scalar curvature also play essential role for formulation
of famous Einstein gravity equations. In particular the space without matter
the Einstein equations have the following form:
1
Rik Rgik = 0 .
2

(5.10)

Due to identities (5.5) and (5.6) for curvature tensor Riemann tensor obeys the following identities:
Rikmn = Riknm ,

Rikmn + Rimnk + Rinkm = 0

(5.11)

Riemann curvature tensor which is curvature tensor for Levi-Civita connection obeys
also the following identities:
Rikmn = Rkimn ,

Rikmn = Rmnki .

(5.12)

These condition lead to the fact that for 2-dimensional Riemannian manifold the
Riemann curvature tensor of Levi-Civita connection has essentially only one non-vanishing
component: all components vanish or equal to component R1212 up to a sign.
Indeed consider for 2-dimensional Riemannian manifold Riemann tensor Rikmn , where
i, k, m, n = 1, 2. Since antisymmetricity with respect to third and fourth indices (Rikmn =
Riknm ), Rik11 = Rik22 = 0 and Rik12 = Rik21 . The same for first and second indices:
since antisymmetricity with respect to the the first and second indices (R12mn = R21mn ),
R11mn = R22mn = 0 and R12mn = Rik21 . If we denote R1212 = a then
R1212 = R2121 = a, R1221 = R2112 = a

(5.13)

and all other components vanish.

5.2.1

Curvature of surfaces in E3 .. Theorema Egregium again

For surfaces in E3 Gaussian curvature is equal to half of scalar curvature:


K=

R
,
2

(5.14)

i
where R = Rkin
g kn is scalar curvature of Riemann curvature tensor.
Equation (5.17) is the fundamental relation which claims that the Gaussian curvature (the magnitude defined in terms of External observer) equals
to the scalar curvature (up to a coefdficient), the magnitude defined in terms
of Internal Observer. This gives us another proof of Theorema Egregium.

100

Prove this formula.


Express Riemannian curvature of surfaces in E3 in terms of derivation formula (4.16).
Consider derivation formula (4.16) for the orthonormal basis {e, f , n, } adjusted to the
surface M :


e
0
a b
e
d f = a 0 c f ,
(5.15)
n
b c 0
n
where as usual e, f , n vector fields 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 defined by the formula (4.24) and (4.25):
Y e = (de(Y))tangent = a(X)f , Y f = (df (Y))tangent = a(X)e ,

(5.16)

According to the definition of curvature calculate


R(e, f )e = e f e f e e [e,f ] e .
Using these formulae one can calculate straightforwardly that for surfaces in E3 Gaussian
curvature is equal to half of the scalar curvature:
K=

R
2

Detailed calculations are following:


Note that since the induced connection is symmetrical connection then:
e f f e [e, f ] = 0 .
hence due to (5.16)
[e, f ] = e f f e = a(e)e a(f )f
Thus we see that R(e, f )e =
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 =
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 [a(e)e a(f )f ]] f =
= [e a(f )f f a(e)f a ([e, f ])] f = da(e, f )f .
Recall that we established in 6.41 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 .

101

(5.17)

This means that


2
R112
= K

(in the basis e, f ), i.e. the scalar curvature


R = 2R1212 = 2K
Thus we come to equation (5.17).
The proof of Theorema Egregium by straightforward calculations see in the last appendix.

5.2.2

Relation between Gaussian curvature and Riemann curvature tensor and Theorema Egregium

Let M be a surface in E3 and Rikmp be Riemann tensor, Riemann curvature tensor of Levi-Civita connection. Recall that this means that Rikmp
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 dx2 + dy 2 + dz 2 . Recall that Riemann curvature tensor is expressed
via Christoffel symbols of connection by the formula
Rikmn = m ink + imp pnk n imk inp pmk

(5.18)

(see he formula (5.4)) where Christoffel symbols of Levi-Civita connection


are defined by the formula


1 ij gjm gjk gmk
i
+ m
(5.19)
mk = g
2
xk
x
xj
(see Levi-Civita Theorem)
i
Recall that scalar curvature R of Riemann tensor equals to R = Rkim
g km ,
where g km is Riemannian metric tensor with upper indices (matrix ||g ik || is
inverse to the matrix ||gik ||).
Now consider 2-dimensional case. One can show that in this case scalar
curvature R can be expressed via the component R1212 = a by the formula
R=

2R1212
det g

(5.20)

2
where det g = det gik = g11 g22 g12
.

To see it note that as it was mentioned in the subsection above the formula for scalar
curvature becomes very simple in two-dimensional case (see formulae (5.11) and (5.13)
above) and in this case it is very easy to calculate Ricci tensor and scalar curvature R.

102

Indeed let R1212 = a. For 2-dimensional Riemannian surface all other components of
Riemann tensor equal to zero or equal to a (see (5.11) and (5.13)). Show it. Using
identities (5.11) and (5.11) we see that
R11 = Ri1i1 = R2121 = g 22 R2121 + g 21 R1121 = g 22 R1212 = g 22 a

(5.21)

R22 = Ri2i2 = R1212 = g 11 R1212 + g 12 R2221 = g 11 R1212 = g 11 a

(5.22)

R12 = R21 =

Ri1i2

R1112

12

12

12

= g R2112 = g R1212 = g a

(5.23)

Thus using the formula for inverse 2 2 matrix we come to the relation



  22

1
R11 R12
g11 a g12 a
g a g 12 a
Rik =
,
=
=
R21 R22
g 21 a g 11 a
det g g21 a g11 a
i.e. for 2-dimensional Riemannian manifold
Rik =

1
R1212 gik ,
det g

(5.24)

Hence for scalar curvatre of 2-dimension Riemannian manifold


R = Rikim g km = Rkm g km =

2
1
R1212 gik g ik =
R1212 .
det g
det g

(5.25)

Note that the relations (5.24) and (5.25) imply thatt


Rik =

1
Rgik .
2

(5.26)

One can say that gravity equation for n = 2 are trivial. The mathematical meaning
of
R this
formula is the following: equation (5.26) means that variation of functional S =
R det gd vanishes and this is one of corollaries of Gauss-Bonet Theorem (see later).

Formula (5.20) expresses scalar curvature for surface in terms of nontrivial component R1212 . On the other hand Gaussian curvature of 2-dimensional
surface is equal to the half of the scalar curvature (see equation (5.14)). Hence
we come to
Proposition For an arbitrary point of the surface M
K=

R
R1212
=
.
2
det g

(5.27)

i
where R = Rkim
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.). This is nothing but

103

Theorema Egregium! Theorema Egregium (see beginning of the section 4)


immediately follows from this Proposition which states that Gaussian curvature is equal (up to a coefficient) to scalar curvature whcih is expressed in
terms of Riemannian metric.
In the next section we will present straightforward calculations where we
check equation (5.27) just by brute force calculating Riemannian metric and
Riemanian curvature tensor.
Before going to calculations, just a simple example Example Let M = S 2
be sphere of radius R in E3 . Show that one cannot find local coordinates
u, v on the sphere such that induced Riemannian metric equals to du2 + dv 2
in these coordinates.
This immediately follows from the Proposition. Indeed suppose there exist local coordinates u, v on the sphere such that induced Riemannian metric
equals to du2 + dv 2 , i.e. Riemannian metric is given by unity matrix. Then
according to the formulae for Levi-Civita connection, the Christoffel 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 R12 .
(The straightforward proof see in the next paragraph)
5.2.3

Straightforward proof of the Proposition (5.27)

(Content of this paragraph is similar to the content of the solution of exercise


6 in Homework 6. It is useful to read them both.)
We prove this fact by direct calculations. The plan of calculations is
following:
Let M be a surface in E3 For an arbitrary point p of the surface M
we consider Cartesian coordinates x, y, z such that origin 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. (See the subsection 4.3.2) The surface M in these
Cartesian coordinates can be expressed by the equation

x = u
(5.28)
y=v

z = F (u, v)

104

where F (u, v) has local extremum at the point u = v = 0. We calculated in


subsection 4.3.2 Gaussian curvature at this point: Gaussian curvature at the
point p equals to
2
K = det S = Fuu Fvv Fuv
.
(5.29)
(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

 

hru , ru i hru , rv i
1 + Fu2 Fu Fv
G=
=
.
(5.30)
hrv , ru i hrv , rv i
Fv Fu 1 + Fv2
This immediately follows from the expression for basic vectors ru , rv
Note that Riemannian metric gik at the point u = v = 0 is defined by
unitymatrix guu = g
= gvu = 0 since p is extremum point:
vv = 1, guv 
2

1 + Fu Fu Fv
1 0
G=
=
since p is stationary point, extremum
Fv Fu 1 + Fv2 p
0 1
(Fu = Fv = 0). Hence the components of the tensor Rikmn and Rikmn =
gij Rjkmn at the point p are the same.
For 2-dimensional surface Riemann curvature tensor has oessentially only
1
one not-vanishing component R212
. All other components vanish all are equal
up to a sign to this component:
R1212 = R2112 = R2121 R1112 = = R2111 = 0 .
So in fact we need to calculate only one component, the component R1212 .
In our calculations we will use the fact that Riemannian metric at the
point p is defined by unity matrix, and that first derivatives of metric at p
vanish, i.e.Christoffel symbols in coordinates u, v vanish at the point p.
Recall that the components of Rikmn are defined by the formula
Rikmn = m ink + imp pnk n imk inp pmk .
(see equation (5.4)). Notice that at the point p not only the matrix of the
metric gik equals to unity matrix, but more: Christoffel 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 (5.30), see also in detail the file The solution of the problem in

105

i
at the point p one can consider
the coursework). Hence to calculate Rkmn
much more simple formula than formula (5.4):

Rikmn |p = m ink |p n imk |p


Try to continue calculations in a more economical way. Due to Levi-Civita
formula


1 ij gjm gjk gmk
i
mk = g
+ m
2
xk
x
xj


1 0
Since metric gik equals to unity matrix g =
at the point p hence g ij
0 1
is unity matrix also:


1 0
ik
g |p =
= ik .
0 1
(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 first derivatives of the metric vanish at the point p:
gik
|p = 0 .
xm
Hence it follows from the formulae above that for an arbitrary indices i, j, k, m, n


km
2
2




km gpr
= g gpr + g km gpr = km gpr .
g
p
p xi xj p
xi
xj
xi p xj p
xi xj p
Now using the Levi-Civita formula for the Christoffel symbols of connection:


1 ij gjm gjk gmk
i
mk = g
+ m
2
xk
x
xj



gjk
g
gmk
=
we come to n imk |p = xn 12 g ij xjm
+

k
xm
xj
p
1 ij

2 gjm
2 gjk
2 gmk
+

xn xk xn xm xn xj


.
p

(5.31)

since first derivatives of metric vanish at the point p.


Now using this formula we are ready to calculate Riemann curvature
tensor Rikmn . Remember that it is enough to calculate R1212 and R1212 =
106

R1212 at the point p since gik = ik at the point p. We have that at p


R1212 |p = 1 122 |p 2 112 |p . Now using equation (5.31) we come to
1
R212
|p

1 122 |p 2 112 |p

1
=
2 x1




g12 g22
1
g11 g12 g12
2 2

p=
+

x
x1
2 x2 x2
x1
x1

2 guv
1 2 gvv
1 2 guu
|p
|

|p
p
uv
2 u2
2 v 2

(5.32)

Now return to the surface (5.28) We have that guu = 1 + Fu2 , guv = Fu Fv
and gvv = 1 + Fv2 ,hence
2 guv
2
,
|p = (Fuu Fv + Fu Fuv )v = Fuu Fvv + Fuv
uv
2 gvv
2
|p = (2Fv Fvu )u = 2Fuv
,
2
u
2 guu
2
|p = (Fuu Fu )v = 2Fuv
,
v 2
Hence
2
2
R1212 |p = (Fuu Fvv + Fuv
) 2Fuv


p

2
= Fuu Fvv Fuv


p

= Kp .

1
The proof is finished: we showed by straightforward calculations that R212
is
equal to Gaussian curvature K. On the other hand at the point p, det g = 1.
Thus we come to the statement of Proposition5.27.
i
Repeat again: all other components of Riemann curvature tensor Rkmn
are equal to R1212 up to a sign or vanish. Hence we calculated Riemann
curvature tensor at the point p and showed that it is essentially is defined
by Gaussian curvature.
It is important to note that in our calculations of R1212 (see formula
(5.32)) we used only the fact that Riemannian metric at the point p is defined by unity matrix, and all first derivatives at this point vanish.(see also
Statement 1 in the solution of exercise 6 of Homework 9)

107

5.3

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:
Z
Kd = 2N
D

(Compare this formula with formula (4.5)).


What is the value of integer N ?
We present now one remarkable Theorem which answers this question
and prove this Theorem using the formula (4.5).
Let M be a closed orientable surface.21 All these surfaces can be classified up to a diffeomorphism. Namely arbitrary closed oriented surface M
is diffeomorphic 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

(5.33)

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 different way. The simplest appearance 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)
21

Closed means compact surface without boundaries. Intuitively orientability means


that one can define out and inner side of the surface. In terms of normal vectors orientability means that one can define the continuous field of normal vectors at all the
points of M . The direction of normal vectors at any point defines outward direction.
Orientable surface is called oriented if the direction of normal vector is chosen.

108

Denote by V number of vertices.


Then it turns out that
P E + V = (M )

(5.34)

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 diffeomorphic to
sphere. For every graph on M P E + V = 0 if M is diffeomorphic to torus.
Now we formulate Gau -Bonnet Theorem.
Let M be closed oriented surface in E3 .
Let K(p) be Gaussian curvature at any point p of this surface.
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
Z
1
Kd = (M ) = 2(1 number of holes)
(5.35)
2 M
In particular for the surface M diffeomorphic to the sphere (M ) = 2,
for the surface diffeomorphic 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 diffeomorphic to the sphere. If it is
sphere of the radius R then curvature is equal to R12 , area of the sphere is
equal to 4R2 and left hand side is equal to 4
= 2.
2
If surface M is an arbitrary surface diffeomorphic 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 diffeomorphic to sphere in E3 . Then there exists at least
one point where Gaussian curvature is positive.
R

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

109

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
3N
N
N
+V =V
= (M ).
2
2
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 (4.5) it is equal to
N 
X
i=1

Z
+

Kd
4i

We see that 2V = N +

= N +

N Z
X
i=1


Kd

Z
= N +

4i

Kd
M

Kd, i.e.


N
= 2(M )
Kd = 2V
2
M
M

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 T M ).
Definition We say that the function F = F (q, q)
on T M 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(3.13).

(6.1)

In other words
d
d
(F (x(t), x(t)))

= 0 if xi (t) :
dt
dt

110

L
i
x

L
= 0.
xi

(6.2)

6.1.2

Basic examples of Integrals of motion: Generalised momentum and Energy

Let L(xi , x i ) does not depend on the coordinate x1 . L = L(x2 , . . . , xn , x 1 , x 2 , . . . , x n ).


Then the function
L
F1 (x, x)
=
x 1
is integral of motion. (In the case if L(xi , x i ) does not depend on the coordinate xi . the
L
function Fi (x, x)
= x
i will be integral of motion.)
Proof is simple. Check the condition (6.2): Euler-Lagrange equations of motion are:


L
d L

= 0 (i = 1, 2, . . . , n)
i
dt x
xi
We see that exactly first equation of motion is


d L
d
= F1 (q, q)
=0
1

dt x
dt

since

L
x1

= 0, .

(if L(xi , x i ) does not depend on the coordinate xi then the function Fi (x, x)
=

integral of motion since i th equation is exactly the condition Fi = 0.)


L
The integral of motion Fi = x
i is called sometimes generalised momentum.

L
x 1

is

Another very important example of integral of motion is: energy.


E(x, x)
= x i

L
L.
x i

(6.3)

One can check by direct


that it is indeed integral of motion. Using Euler

 calculation
L
d
L
Lagrange equations dt x
i xi we have:
d
d
E(x(t), x(t))

=
dt
dt




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

L dxi
dL(x, x)

dL(x, x)

dL(x, x)

L dx i
+

= 0.
x i dt
xi dt
dt
dt
dt

6.1.3

Integrals of motion for geodesics

Apply the integral of motions for studying geodesics.


i k
The Lagrangian of free particle Lfree = gik (x)2 x x . For Lagrangian of free particle
solution of Euler-Lagrange equations of motions are geodesics.
i k
If F = F (x, x)
is the integral of motion of the free Lagrangian Lfree = gik (x)2 x x then
the condition (6.1) means that the magnitude I(t) = F (xi (t), x i (t)) is preserved along the
geodesics:
I(t) = F (xi (t), x i (t)) = const, i.e.

d
I(t) = 0 if xi (t) is geodesic.
dt

111

(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:


gpq (x)x p x q

2
gik (x)x i x k
L

=
E = x i i L = x i
x
x i
2
gik (x)x i x k
gik (x)x i x k
=
.
(6.5)
2
2
This is an integral of motion for an arbitrary Riemannian metric. It is preserved on an
arbitrary geodesic


dE(t)
d 1
i
k
=
gik (x(t))x (t)x (t) = 0 .
dt
dt 2
x i giq (x)x q

In fact we already know this integral of motion: Energy (6.5) is proportional to the square
of the length of velocity vector:
q

(6.6)
|v| = gik (x)x i x k = 2E .
We already proved that velocity vector is preserved along the geodesic (see the Proposition
in the subsection 3.2.1 and its proof (??).)
Example 2 Consider Riemannian metric G = adu2 + bdv 2 (see also calculations in
subsection 2.3.3) in the case if a = a(u), b = b(u), i.e. coefficients do not depend on the
second coordinate v:

1
G = a(u)du2 + b(u)dv 2 , Lfree =
a(u)u 2 + b(u)v 2
(6.7)
2
We see that Lagrangian does not depend on the second coordinate v hence the magnitude
F =

Lfree
= b(u)v
v

(6.8)

is preserved along geodesic. It is integral of motion because Euler-Lagrange equation for


coordinate v is
d Lfree
Lfree
d Lfree
d

=
= F = 0 since
dt v
v
dt v
dt

Lfree
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)
2
3
2
2
2
Sphere
 of the radius R in E . Riemannian metric: G = Rd + R sin d and
2
2
1
2 2
2
2
2
It is the case (6.7) for u = , v = , b(u) = R sin The
Lfree = 2 R + R sin
integral of motion is
Lfree
F =
= R2 sin2

112

Example (cone)

x = ah cos
Consider cone y = ah sin

z = bh

. Riemannian metric:

G = d(ah cos )2 + d(ah sin )2 + (dbh)2 = (a2 + b2 )dh2 + a2 h2 d2 .


and free Lagrangian
Lfree =

(a2 + b2 )h 2 + a2 h2 2
.
2

The integral of motion is


F =

Lfree
= a2 h2 .

Example (general surface of revolution)


Consider a surface of revolution in E3 :

x = f (h) cos
r(h, ) :
y = f (h) sin

z=h

(f (h) > 0)

(6.9)

(In
the case f (h) = R it is cylinder, in the case
f (h) = kh it is a cone, in the case
f (h) = R2 h2 it is a sphere, in the case f (h) = R2 + h2 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 0 (h) cos dh f (h) sin d)2 +
(f 0 (h) sin dh + f (h) cos d)2 + dh2 = (1 + f 02 (h))dh2 + f 2 (h)d2 .
The free Lagrangian of the surface of revolution is

1 + f 02 (h) h 2 + f 2 (h) 2
Lfree =
.
2
and the integral of motion is
F =

6.1.4

Lfree
= f 2 (h).

Using integral of motions to calculate geodesics

Integrals of motions may be very useful to calculate geodesics. The equations for geodesics
are second order differential equations. If we know integrals of motions they help us to
solve these equations. Consider just an example.

113

For Lobachevsky plane the free Lagrangian L = x 2y+2y . We already calculated geodesics
in the subsection 3.3.4. Geodesics are solutions of second order Euler-Lagrange equations
2
2
for the Lagrangian L = x 2y+2y (see the subsection 3.3.4)
(
x

y+

2x y
y =0
y 2
x 2
y y

=0

It is not so easy to solve these differential equations.


For Lobachevsky plane we know two integrals of motions:
E=L=

x 2 + y 2
,
2y 2

and F =

L
x
= 2.
x
y

(6.10)

These both integrals preserve in time: if x(t), y(t) is geodesics then


(
(
x(t)

F = y(t)
x = C1 y 2
2
p

2
2

+y(t)

y = 2C2 y 2 C12 y 4
= C2
E = x(t)
2y(t)2
These are first order differential equations. It is much easier to solve these equations in
general case than initial second order differential equations.

6.1.5

Killing vectors of Lobachevsky plane and geodesics

Killing vector field of Rimeannina manifold (M, G) is an infinitesimal isometry of the


Rimeannian metric G: under infinitesimal transform x x + K, xi xi + K I (x)
(2 = 0) metric does not change:
gik (x)dxi dxk = gik (xr + K i )(dxi + m K i dxm )(dxk + n K k dxn ) .

(1)

Expanding this formula by and using the fact that 2 = 0 we come to


K i i gkm + k K r grm + m K r grk = 0 ,

(1a)

(i.e. Lie derivative LK G = 0.)


Examples: Killings of plane, sphere, cylindre, Lobachevsky plane.....
Theorem Let V be a vector space of all Killing vector fields of Riemannian manifold
M . Then the dimension of V is less or equal than n(n+1
2) .
It means that for surfaces the number of independent Killing vector fields is less or
equal to 3.
One can prove that it is only for plane, sphere and Lobachevsky plane that number
of independent Killing vector fiels is equal to 3.
We calculate here Killing vector fields for Lobachevsky plane and use them for finding
geodesics.

114

x k x p

Theorem Let K be Killing tor field on Riemannian manifold (M, G) , and L = kp 2


Lagragian of free particle on M . We know that geodesics are solutions of its equations
of motions.
The magnitude
L(x, x)

I = IK = K i (x)
x i
is an integral of motion, i.e. it is preserved along geodesics.
The proof of the Theorem is obvious. The condition that K is Killing vector field
means that
L(xi + K i , x i + K i ) ,
(2)
i.e.
K i (x)

L
dK i L
+
= 0.
i
x
dt x i

(2a)

Hence




dK i L(x, x)
L(x, x)

L(x, x)

i
i d
=
=
+K
K (x)
x i
dt
x i
dt
x i


L
L(x, x)

dK i L
L(x, x)

i
i
K (x) i +
+K

= 0.
x {z dt x i}
xi
x i
|
{z
}
|
condition that K is Killing
equations of motion

d
d
IK =
dt
dt

Use this Theorem to find geodesics.


First find Killing vector fields, i.e. infinitesimal isometries.
Since the dimension is equal 2, the dimension of space of Killing vector fields is 3.
We will find three independent Killing vector fields.
2
2
is evidently invariant with
There are two evident Killing vectors: Metric G = dx y+dy
2
(
2
2
2
x x
+d(y)2
= dx y+dy
.
respect to translations x x + a and homothety:
: d(x)(y)
2
2
y 7 y

Infinitesimal translation is x0 = x + , y 0 = y, the vector field D1 = x


. Infinitesimal

0
0
homothety is x = x + x, y = y + y, the vector field D2 = x x + y y .
Now most
 interesting: find the third Killing vector field. Use the fact that inversion
y
x
O : (x, y) 7 x2 +y
preserves the metric. Consider the infinitesimal transforma2 , x2 +y 2
tion L = O T O (L0 = id):

L :

x
+
x2 +y 2

2
2
 
 x 
 x

y
x
x2 +y

+ x2 +y
x
2
2
2
2 +
2 +
2
.
y
x +y
x +yy
7
y

y
x2 +y 2
2
2
2
2

x +y
x +y
x
x2 +y 2

115

y
x2 +y 2

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 E3 then induced Riemannian metric GM is
defined by the formulae
hX, Yi = GM (X, Y) = G(X, Y) ,

(6.11)

where G is Euclidean metric in E3 :


3
3
X
X


GM = dx2 + dy 2 + dz 2 r=r(u,v) =
(dxi )2 r=r(u,v) =
i=1

i=1

xi
du
u

2

xi xi
du du ,
u u

i.e.
GM = g du , where g =

xi xi
du du .
u u

We use notations x, y, z or xi (i = 1, 2, 3) for Cartesian coordinates in E3 , u, v or u


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

= r =

i
u
u
|
{z x}
|{z}
Internal observer External observer
We have already plenty examples in the subsection 1.4. In particular for scalar product



g =
,
= xi xi .hr , r i .
(6.12)
u u

6.2.1

Recalling Weingarten operator

Continue to play with formulae 22 .


Recall the Weingarten (shape) operator which acts on tangent vectors:
SX = X n ,

(6.13)

where we denote by n-unit normal vector field at the points of the surface M : hn, r i = 0,
hn, r i = 1.
Now we realise that the derivative X R of vector field with respect to another vector
field is not a well-defined object: we need a connection. The formula X R in Cartesian
coordinates, is nothing but the derivative with respect to flat canonical connection: If
22

In some sense differential geometry it is when we write down the formulae expressing
the geometrical facts, differentiate these formulae then reveal the geometrical meaning of
the new obtained formulae e.t.c.

116

we work only in Cartesian coordinates we do not need to distinguish between X R and


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

So the rigorous definition of Weingarten operator is


SX = can.flat
n,
X

(6.14)

but we often use the former one (equation (6.16)) 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: hn, Xi = 0 X is tangent to the surface.
Indeed:
0 = X hn, ni = 2hX n, ni = 2hSX, ni = 0 SX is tangent to the surface
Recall also that normal unit vector is defined up to a sign, n n. On the other hand
if n is chosen then S is defined uniquely.
We use notations x, y, z or xi (i = 1, 2, 3) for Cartesian coordinates in E3 , u, v or u
( = 1, 2) for coordinates on the surface. We usually omit summation symbol over dummy
indices. For coordinate tangent vectors
xi

= r =

i
u
u
|
{z x}
|{z}
Internal observer External observer
We have already plenty examples in the subsection 1.4. In particular for scalar product



g =
,
= xi xi .hr , r i .
(6.15)
u u

6.2.2

Recalling Weingarten operator

Continue to play with formulae 23 .


Recall the Weingarten (shape) operator which acts on tangent vectors:
SX = X n ,

(6.16)

where we denote by n-unit normal vector field at the points of the surface M : hn, r i = 0,
hn, r i = 1.
23

In some sense differential geometry it is when we write down the formulae expressing
the geometrical facts, differentiate these formulae then reveal the geometrical meaning of
the new obtained formulae e.t.c.

117

Now we realise that the derivative X R of vector field with respect to another vector
field is not a well-defined object: we need a connection. The formula X R in Cartesian
coordinates, is nothing but the derivative with respect to flat canonical connection: If
we work only in Cartesian coordinates we do not need to distinguish between X R and
can.flat
R. Sometimes with some abuse of notations we will use X R instead can.flat
R,
X
X
but never forget: this can be done only in Cartesian coordinates where Christoffel symbols
of flat canonical connection vanish:
can.flat
X R = X
R in Cartesian coordinates .

So the rigorous definition of Weingarten operator is


SX = can.flat
n,
X

(6.17)

but we often use the former one (equation (6.16)) 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: hn, Xi = 0 X is tangent to the surface.
Indeed:
0 = X hn, ni = 2hX n, ni = 2hSX, ni = 0 SX is tangent to the surface
Recall also that normal unit vector is defined up to a sign, n n. On the other hand
if n is chosen then S is defined uniquely.

6.2.3

Second quadratic form

We define now the new object: second quadratic form


Definition. For two tangent vectors X, Y A(X, Y) is defined such that

can.flat
Y = A(X, Y)n
X

(6.18)

i.e. we take orthogonal component of the derivative of Y with respect to X.


This definition seems to be very vague: to evaluate covariant derivative we have to
consider not a vector Y at a given point but the vector field. 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 definition of second quadratic form and from the properties
of Weingarten operator that


A(X, Y) = can.flat
Y , n = can.flat
Y, n =
X
X


X hY, ni Y, can.flat
n = hS(X), Yi
(6.19)
X
We proved that second quadratic form depends only on value of vector field 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 tangent vectors X, Y in a given point.

118

A : Tp M Tp M R,

A(X, Y) = A(Y, X) = hSX, Yi .

(6.20)

In components
A = A du du = hr , ni =

2 xi
ni .
u u

(6.21)

and
S = g A = g xi ni ,

(6.22)

i.e.
A = GS, S = G1 A .
Remark The normal unit vector field is defined up to a sign.

6.2.4

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(G1 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 curvartures 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 Riemannian 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.

6.2.5

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 (1.51)).
Cylinder is given by the equation x2 + y 2 = R2 . One can consider the following
parameterisation of this surface:

0
R sin
x = R cos
r(h, ) :
, rh = 0 , r = R cos ,
(6.23)
y = R sin

1
0
z=h

119


Gcylinder = dx2 + dy 2 + dz 2 x=a cos ,y=a sin ,z=h =
2

||g || =

= (a sin d) + (a cos d) + dh = a d + dh ,

1
0

0
R2


.

cos
cos
Normal unit vector n = sin . Choose n = sin . Weingarten operator
0
0

cos
Sh = rcan.flat
n = rh n = h sin = 0 ,
h
0

cos
sin

n = r n = sin = cos = .
S = rcan.flat

R
0
0


  

0
0 0
h
S
.
= , S =

0 1
R
R

0
Calculate second quadratic form: rhh = h rh = 0 , rh = rh =
0

(6.24)

R sin
R sin
R cos
h R cos = 0, r = R cos = R sin = Rn .
0
0
0
We have

 

hrhh , ni hrh , ni
0
0
= hr , ni, A =
=
,
hrh , ni hr , ni
0 R

 


0
0
1 0
0
0
A = GS =
=
,
0 R2
0 R
0 R1

(6.25)

For Gaussian and mean curvatures we have



det A
0
= det
K = det S =
0
det G
and mean curvature


H = Tr S = Tr

0
0

0
R1

0
R1


=


= 0,

1
.
R

Mean curvature is define up to a sign. If we change n n mean curvature H


Gaussian curvature will not change.
Cone
We already calculated induced Riemannian metric on the cone (see (??).

120

(6.26)

(6.27)
1
R

and

Cone is given by the equation x2 + y 2 k 2 z 2 = 0. One can consider the following


parameterisation of this surface:

k cos
kh sin
x = kh cos
r(h, ) :
, rh = k sin , r = kh cos ,
(6.28)
y = kh sin

1
0
z=h

Gcone = dx2 + dy 2 + dz 2 x=kh cos ,y=kh sin ,z=h =
= (kh sin d + k cos dh)2 + (kh cos d + k sin dh)2 + dh2 =
 2

k +1
0
2 2
2
2
2
k h d + (k + 1)dh , ||g || =
.
0
k 2 h2

cos
One can see that N = sin is orthogonal to the surface: Nrh , r . Hence normal
k

cos
cos
1
sin . Choose n = 1 2 sin . Weingarten operator
unit vector n = 1+k
2
1+k
k
k

cos
1
sin = 0 ,
Sh = rcan.flat
n = rh n = h
h
1 + k2
k

cos
1
sin =
S = rcan.flat
n = r n =

1 + k2
k

sin
1
cos =

.
1 + k2
kh k 2 + 1
0
!
 


0
0
0
h
S
,
S
=
=
.

0 kh1
khk2 +1

k2 +1

0
Calculate second quadratic form: rhh = h rh = 0 , rh = rh =
0

(6.29)

kh sin
k sin
kh sin
kh cos
h kh cos = k cos , r = kh cos = kh sin .
0
0
0
0
We have

A = hr , ni, A =

hrhh , ni
hrh , ni

121

 
0
hrh , ni
=
0
hr , ni

0
kh
1+k
2


,

(6.30)


A = GS =

k2 + 1
0

0
k 2 h2


0
0


=

kh k2 +1

0
0

0
kh
k2 +1


,

For Gaussian and mean curvatures we have


K = det S =


0
det A
= det
0
det G


= 0,

(6.31)

.
kh k 2 + 1

(6.32)

kh k2 +1

and mean curvature



0
H = Tr S = Tr
0

0
1

kh k2 +1


=

Mean curvature is define up to a sign. If we change n n mean curvature H


Gaussian curvature will not change.

1
R

and

Sphere
Sphere is given by the equation x2 + y 2 + z 2 = a2 . Consider the parameterisation of
sphere in spherical coordinates

x = R sin cos
(6.33)
r(, ) :
y = R sin sin

z = R cos
We already calculated induced Riemannian metric on the sphere (see (6.2.5)). Recall
that

R sin sin
R cos cos
r = R cos sin , r = R sin cos
0
R sin
and

GS 2 = dx2 + dy 2 + dz 2 x=R sin cos ,y=R sin sin ,z=R cos =
(R cos cos d R sin sin d)2 + (R cos sin d + R sin cos d)2 +
(R sin d)2 = R2 cos2 d2 + R2 sin2 d2 + R2 sin2 d2 =
 2

R
0
2
2
2
2
2
= R d + R sin d ,
||g || =
.
0 R2 sin2
For the sphere
r(, )is orthogonal to the
surface. Hence

normal unit vector n(, ) =


sin cos
sin cos
r(,)
= sin sin . Choose n = Rr = sin sin . Weingarten operator
R
cos
cos
r
r
S = can.flat
n
=

n
=

= ,

r
R
R
r
r
S = rcan.flat
n = n =
= .

R
R

122


=

1


S=

1
R

(6.34)

R sin cos
For second quadratic form: r = r = R sin sin , r = r =
R cos

R sin sin
R cos sin
R sin sin
R sin cos
R sin cos = R cos cos , r = R sin cos = R sin sin .
0
0
0
0
We have


 

R
0
hr , ni hr , ni
=
,
hr , ni hr , ni
0
R sin2



  1
1
0
0
0
R
=
R
,
1
0
0 sin2
R2 sin2
R

A = hr , ni, A =

A = GS =

R2
0

(6.35)

For Gaussian and mean curvatures we have


K = det S =

 1
det A
R
= det
0
det G

and mean curvature


H = Tr S = Tr

R1
0

0
R1

0
R1


=

=
2
,
R

1
,
R2

(6.36)

(6.37)

Mean curvature is define up to a sign. If we change n n mean curvature H R1 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.

6.2.6

Proof of the Theorem of parallel transport along closed


curve

We are ready now to prove the Theorem. Recall that the Theorem states following:
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
Z
= (X, RC X) =
Kd ,
(6.38)
D

is the area element induced by


where K is the Gaussian curvature and d = det gdudv

the Riemannian metric on the surface M , i.e. d = det gdudv.


(see (4.5).
Recall that for derivation formula (4.16) we obtained structure equations
da + b c = 0
db + c a = 0
dc + a a = 0

123

(6.39)

We need to use only one of these equations, the equation


da + b c = 0 .

(6.40)

This condition sometimes is called Gau condition.


Let as always {e, f , n} be an orthonormal basis in Tp E3 at every point of surface
p M such that {e, f } is an orthonormal basis in Tp M at every point of surface p M .
Then the Gau condition (6.40) and equation (4.31) 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 )

(6.41)

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) = (u1 , 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 field 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 field X(t) is covariantly constant along the curve:
X(t)
=0
dt
To write this equation in components we usually expanded the vector field in the coordinate
basis {ru = u , rv = v } and used Christoffel symbols of the connection
: =


Now we will do it in different way: instead coordinate basis {ru = u , rv = 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

(6.42)

(see the equations (4.24) and (4.25))


Let
X = X(u(t)) = X 1 (t)e(u(t)) + X 2 (t)f (u(t))
Lbe an expansion of tangent vector field X(t) over basis {e, f }. Let v be velocity vector
of the curve C. Then the equation of parallel transport X(t)
= 0 will have the following
dt
appearance:

X(t)
= 0 = v X 1 (t)e(u(t)) + X 2 (t)f (u(t)) =
dt
1
dX (t)
dX 2 (t)
e(u(t)) + X 1 (t)v e(u(t)) +
f (u(t)) + X 2 (t)v f (u(t)) =
dt
dt
dX 1 (t)
dX 2 (t)
e(u(t)) + X 1 (t)a(v)f (u(t)) +
f (u(t)) X 2 (t)a(v)e(u(t)) =
dt
dt




dX 1 (t)
dX 2 (t)
2
1
X (t)a(v) e(u(t)) +
+ X (t)a(v) f (u(t)) = 0.
dt
dt

124

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) + iX 2 (t) = a(v(t))X 2 ia(v(t)X 1 = ia(v)Z(t),
i.e.

dZ(t)
= ia(v(t))Z(t)
dt

(6.43)

The solution of this equation is:


Z(t) = Z(0)ei
Calculate

R t1
0

Rt
0

a(v( ))d

(6.44)

a(v( ))d for closed curve u(0) = u(t1 ). Due to Stokes Theorem:
Z t1
Z
Z
a(v(t))dt =
a=
da
0

Hence using Gauss condition (6.40) we see that


Z t1
Z
Z
Z
a(v(t))dt =
a=
da =
bc
0

Claim

bc=

da =

Kd .

(6.45)

(6.46)

Theorem follows from this claim:


Z(t1 ) = Z(0)ei

= Z(0)ei D bC
R
R
Denote the integral i D b C by : = i D b C. We have

Z(t1 ) = X 1 (t1 ) + iX 2 (t1 ) = X 1 (0) + iX 2 (0) ei =
C

(6.47)

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
(6.48)

(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 (6.41). We see that
b c(e, f ) = da(e, f ) = Kd(e, f )
Hence 2-forms b c, da and volume form d coincide. Thus we prove (6.45).

125

You might also like