An Introduction To Differential Geometry Through Computation PDF
An Introduction To Differential Geometry Through Computation PDF
DigitalCommons@USU
Textbooks Open Texts
Winter 1-26-2016
Recommended Citation
Fels, Mark E., "An Introduction to Differential Geometry through Computation" (2016). Textbooks. 2.
http://digitalcommons.usu.edu/oer_textbooks/2
This Book is brought to you for free and open access by the Open Texts at
DigitalCommons@USU. It has been accepted for inclusion in Textbooks
by an authorized administrator of DigitalCommons@USU. For more
information, please contact [email protected].
An Introduction to Differential Geometry
through Computation
Mark E. Fels
c Draft date March 9, 2017
Contents
Preface v
1 Preliminaries 1
1.1 Open sets . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Smooth functions . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Smooth Curves . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.4 Composition and the Chain-rule . . . . . . . . . . . . . . . . . 6
1.5 Vector Spaces, Basis, and Subspaces . . . . . . . . . . . . . . . 10
1.6 Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
1.7 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2 Linear Transformations 25
2.1 Matrix Representation . . . . . . . . . . . . . . . . . . . . . . 25
2.2 Kernel, Rank, and the Rank-Nullity Theorem . . . . . . . . . 30
2.3 Composition, Inverse and Isomorphism . . . . . . . . . . . . . 34
2.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3 Tangent Vectors 43
3.1 Tangent Vectors and Curves . . . . . . . . . . . . . . . . . . . 43
3.2 Derivations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.3 Vector fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.4 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
i
ii CONTENTS
4.5 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
7 Hypersurfaces 133
7.1 Regular Level Hyper-Surfaces . . . . . . . . . . . . . . . . . . 133
7.2 Patches and Covers . . . . . . . . . . . . . . . . . . . . . . . . 135
7.3 Maps between surfaces . . . . . . . . . . . . . . . . . . . . . . 138
7.4 More General Surfaces . . . . . . . . . . . . . . . . . . . . . . 139
7.5 Metric Tensors on Surfaces . . . . . . . . . . . . . . . . . . . . 139
7.6 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146
Bibliography 213
iv CONTENTS
Preface
This book was conceived after numerous discussions with my colleague Ian
Anderson about what to teach in an introductory one semester course in
differential geometry. We found that after covering the classical differential
geometry of curves and surfaces that it was difficult to make the transition to
more advanced texts in differential geometry such as [4], or to texts which use
differential geometry such as in differential equations [9] or general relativity
[11], [13]. This book aims to make this transition more rapid, and to prepare
upper level undergraduates and beginning level graduate students to be able
to do some basic computational research on such topics as the isometries of
metrics in general relativity or the symmetries of differential equations.
This is not a book on classical differential geometry or tensor analysis,
but rather a modern treatment of vector fields, push-forward by mappings,
one-forms, metric tensor fields, isometries, and the infinitesimal generators
of group actions, and some Lie group theory using only open sets in IR n .
The definitions, notation and approach are taken from the corresponding
concept on manifolds and developed in IR n . For example, tangent vectors
are defined as derivations (on functions in IR n ) and metric tensors are a field
of positive definite symmetric bilinear functions on the tangent vectors. This
approach introduces the student to these concepts in a familiar setting so
that in the more abstract setting of manifolds the role of the manifold can
be emphasized.
The book emphasizes liner algebra. The approach that I have taken is to
provide a detailed review of a linear algebra concept and then translate the
concept over to the field theoretic version in differential geometry. The level of
preparation in linear algebra effects how many chapters can be covered in one
semester. For example, there is quite a bit of detail on linear transformations
and dual spaces which can be quickly reviewed for students with advanced
training in linear algebra.
v
CONTENTS 1
Preliminaries
1
2 CHAPTER 1. PRELIMINARIES
U = { (x, y) ∈ IR 2 | y > 0 }
and is open.
V = { (x, y) ∈ IR 2 | y ≥ 0 }
is not open. Any point (x, 0) ∈ V can not satisfy the open ball condition.
S n = { x ∈ IR n+1 | d(x, 0) = 1 }
and S n is not open. No point x ∈ S n satisfies the open ball condition. The
set S n is the boundary of the open ball B1 (0) ⊂ IR n+1 .
f (x) = f (x1 , x2 , . . . , xn ).
∂kf
∂xi1 ∂xi2 . . . ∂xik
where 1 ≤ i1 , i2 , . . . , ik ≤ n. We say for a function f : U → IR , U an
open set in IR n , that f ∈ C k (U ) if all the partial derivatives up to order k
exist at every point in the open set U and they are also continuous at every
point in U . A function f : U → IR is said to be smooth or f ∈ C ∞ (U ) if
f ∈ C k (U ) for all k ≥ 0. In other words a function is smooth if its partial
derivatives exist to all orders at every point in U , and the resulting functions
are continuous.
3. (f g)(x) = f (x)g(x) ∈ C k (U ),
f (x)
4. ( fg )(x) = g(x)
∈ C k (V ), where V = { x ∈ U | g(x) 6= 0 }.
Φ(x, y) = (x + y, x − y, x2 + y 2 )
has components
Therefore Φ ∈ C ∞ (IR 2 , IR 3 ).
pPn
where ||σ̇|| = i )2 .
i=1 (σ̇
dσ
σ̇(t) = = (− sin t, cos t, 1).
dt
π
When t = 4
we have the tangent vector (− √12 , √12 , 1), which looks like,
diagram
y a = Φa (x1 , . . . , xn ) 1 ≤ a ≤ m,
uα = Ψα (y 1 , . . . , y m ) 1 ≤ α ≤ l.
uα = Ψα (Φ(x1 , . . . , xn )) 1 ≤ α ≤ l.
Example 1.4.1. Let σ : IR → IR 3 be the helix from example 1.3.1, and let
Φ : IR 3 → IR 2 be
Example 1.4.3. We verify the chain-rule for the functions Ψ and Φ in ex-
ample 1.4.1. For the left side of equation 1.5, we have using equation 1.4,
(1.6) ∂x (Ψ ◦ Φ) = (y − 1, 2xy + 2yz + y 2 ).
While for the right side we need
(1.7)
∂Ψ
= (∂u Φ1 , ∂u Φ2 )|(u,v)=Φ(x,y,z) = (1, v)|(u,v)=Φ(x,y,z) = (1, x + y)
∂u (u,v)=Φ(x,y,z)
∂Ψ
= (∂u Φ1 , ∂u Φ2 )|(u,v)=Φ(x,y,z) = (−1, u)|(u,v)=Φ(x,y,z) = (−1, xy + 2yz)
∂v (u,v)=Φ(x,y,z)
and
(1.8) ∂x Φ = (y, 1).
Therefore the two terms on the right side of 1.5 for α = 1, 2 can be computed
from equations 1.7 and 1.8 to be
∂Ψ1 ∂Φ1 ∂Ψ1 ∂Φ2
+ =y−1
∂u ∂x ∂v ∂x
∂Ψ2 ∂Φ1 ∂Ψ2 ∂Φ2
+ = (x + y)y + (xy + 2yz)
∂u ∂x ∂v ∂x
which agrees with 1.6.
8 CHAPTER 1. PRELIMINARIES
f (x, y, z) = log(x + y + z)
and where
∂f
(1.11) gi (p) = .
∂xi p
1.4. COMPOSITION AND THE CHAIN-RULE 9
Therefore let
1
Z
∂f
(1.15) gi (x) = dt,
0 ∂ξ i ξ=p+t(x−p)
which satisfy 1.10 on account of equation 1.14. The smoothness property of
the functions gi (x) follows by differentiation under the integral sign (see [12]
where this is justified).
Finally substituting x = p into equation 1.15 gives
Z 1
∂f ∂f
gi (p) = dt =
0 ∂ξ i ∂xi
ξ=p x=p
If A ∈ Mm×n (IR ), B ∈ Mn×p (IR ) then the product of A and B is the matrix
AB ∈ Mm×p (IR ) defined by
n
X
(AB)as = Aaj Bsj , 1 ≤ a ≤ m, 1 ≤ s ≤ p.
j=1
Note that the zero-vector 0 will always satisfy this condition with c1 = 0, v1 ∈
S. The set of all vector which are a linear combination of S is called the span
of S and denoted by span(S) .
A subset S ⊂ V is linearly independent if for every choice {vi }1≤i≤k ⊂ S,
the only combination
k
X
(1.17) ci v i = 0
i=1
i
is c = 0, 1 ≤ i ≤ k. The empty set is taken to be linear independent.
12 CHAPTER 1. PRELIMINARIES
1 0 21 | 32
1 1 0 | 2
2 0 1 | 3 rref → 0 1 − 12 | 12 .
1 −1 1 | 1 0 0 0 | 0
equations
1 1 0 a
c1 2 + c2 0 + c3 1 = b
1 −1 1 c
for c1 , c2 , c3 ∈ IR where the right hand side is any vector in IR 3 . This is the
matrix equation
1
1 1 0 c a
2
(1.21) 2 0 1 c = b ,
1 −1 1 c3 c
The augmented matrix and a reduced form are,
1 1 0 | a 1 1 0 | a
2 0 1 | b → 0 −2 1 | b − 2a .
1 −1 1 | c 0 0 0 | c−b+a
Therefore if c − b + a 6= 0, the system has no solution and S is not a spanning
set.
Lastly, a subset β ⊂ V is a basis if β is linearly independent and a
spanning set for V . We will always think of a basis β as an ordered set.
Example 1.5.8. The set S in equation 1.18 of example 1.5.6 is not a basis
for IR 3 . It is not linearly independent a linearly independent set, nor is it a
spanning set.
Example 1.5.9. Let V = IR n , and let
β = {e1 , e2 , . . . , en },
where
0
0
.
.
.
0
(1.22) ei = 1 in the ith row, 0 otherwise.
1
0
.
..
0
The set β is the standard basis for IR n , and the dimension of IR n is n.
1.5. VECTOR SPACES, BASIS, AND SUBSPACES 15
The collection {Eji }1≤i≤m,1≤j≤n forms a basis for Mm×n (IR ) called the stan-
dard basis. We order them as β = {E11 , E21 , . . . , E12 , . . . , Enm }
1. 0 ∈ W ,
2. u + v ∈ W, f or allu, v ∈ W ,
3. cu ∈ W, f or allu ∈ W, c ∈ IR .
1.6. ALGEBRAS 17
1.6 Algebras
Definition 1.6.1. An algebra (over IR ) denoted by (V, ∗), is a vector space
V (over IR ) together with an operation ∗ : V × V → V satisfying
2. v ∗ (aw1 + bw2 ) = av ∗ w1 + bv ∗ w2 .
commutative if
(1.25) v1 ∗ v2 = v2 ∗ v1 ,
18 CHAPTER 1. PRELIMINARIES
and anti-commutative if
v1 ∗ v2 = −v2 ∗ v1
for all v1 , v2 , v3 ∈ V .
Example 1.6.2. Let V = IR 3 with its usual vector-space structure. Let
the multiplication on V be the cross-product. Then (V, ×) is an (anti-
commutative) algebra which is not associative.
Example 1.6.3. Let n ∈ Z + and let V = Mn×n (IR ) be the vector-space
of n × n matrices with ordinary matrix addition and scalar multiplication
defined in equation 1.16. Let ∗ be matrix multiplication. This is an algebra
because of the following algebraic properties of matrix multiplication:
(cA + B) ∗ C = cA ∗ C + B ∗ C
A ∗ (cB + C) = cA ∗ B + A ∗ C
for all c ∈ IR , A, B, C ∈ Mn×n (IR ). These are properties 1 and 2 in Def-
inition 1.6.1. This algebra is associative because matrix multiplication is
associative.
Example 1.6.4. Let V = C k (U ), where U is an open set in IR n . This is
vector-space (see example 1.5.5). Define multiplication of vectors by f ∗ g =
f · g by the usual multiplication of functions. Part 3) in Lemma 1.2.4 implies
f ∗ g ∈ C k (U ). Therefore C k (U ) is an algebra for any k (including k = ∞).
This algebra is associative and commutative.
Example 1.6.5. Let V = IR 2 with the standard operations of vector addi-
tion and scalar multiplication. We define vector multiplication by consider-
ing V as the complex plane. The multiplication is determined by multiplying
complex numbers,
(x + iy)(u + iv) = xu − yv + i(xv + yu).
Therefore on V we define (x, y) ∗ (u, v) = (xu − yv, xv + yu) which makes
V = IR 2 into a commutative and associative algebra.
Now let V = IR 4 and consider points in V as pairs of complex numbers
(z1 , z2 ). We can define a multiplication on V = IR 4 in the following way,
(1.26) (z1 , z2 ) ∗ (w1 , w2 ) = (z1 w1 − z2 w̄2 , z1 w2 + z2 w̄1 )
where w̄1 and w̄2 are the complex conjugates. This make V = IR 4 into an
algebra called the quaternions.
1.6. ALGEBRAS 19
Now equation (1.28) holds for each choice 1 ≤ i, j ≤ n we can then write, for
each 1 ≤ i, j ≤ n there exists ckij ∈ IR , such that
n
X
(1.29) ei ∗ ej = ckij ek .
k=1
The real numbers ckij are called the structure constants of the algebra in the
basis β.
Example 1.6.9. Let β = {E11 , E21 , E12 , E22 } be the standard basis for M2×2 (IR ).
Then
(1.30)
E11 ∗ E11 = E11 , E11 ∗ E21 = E21 , E11 ∗ E12 = 0, E11 ∗ E22 = 0,
E21 ∗ E11 = 0, E21 ∗ E21 = 0, E21 ∗ E12 = E11 , E21 ∗ E22 = E21 ,
E12 ∗ E11 = E12 , E12 ∗ E21 = E22 , E12 ∗ E12 = 0, E12 ∗ E22 = 0,
E22 ∗ E11 = 0, E22 ∗ E21 = 0, E22 ∗ E12 = E12 , E22 ∗ E22 = E22 ,
Therefore the non-zero structure constants (equation 1.29) are found from
equation 1.30 to be
1.7 Exercises
1. Let Φ : IR 2 → IR 2 , and Ψ : U → IR 2 be
x
Φ(x, y) = (x + y, x2 + y 2 ), Ψ(x, y) = ( , x − y)
y
where U = {(x, y) | y 6= 0}.
2. Find the functions g1 (x, y), g2 (x, y) at p = (1, 2) for f (x, y) = ex−y in
Theorem 1.4.8 and verify equation 1.10.
d2 γ
κ = || ||
ds2
is called the curvature of γ.
(a) Compute the arc-length parameterization for the helix from ex-
ample 1.3.1 on I = (0, 2π).
22 CHAPTER 1. PRELIMINARIES
(b) Compute ||γ 0 || for the helix ( 0 is the derivative with respect to s).
(c) Prove ||γ 0 || = 1 for any curve σ.
(d) Compute κ (as a function of s) for the helix.
(e) Show that κ ◦ s(t) for a curve σ can be computed by
−1
ds dT
(1.32) κ(t) = || ||
dt dt
(a)
1 −1
1 −1
A= B = 2 3
2 1
1 1
(b)
1 3
1 −2 3
A = 2 2 B=
−1 1 1
3 1
Show that,
10. Consider the vector-space V = Mn×n (IR ) and define the function [ , ] :
V × V → V by
11. Compute the structure constants for gl(2, IR ) using the standard basis
for M2×2 (IR ).
Linear Transformations
25
26 CHAPTER 2. LINEAR TRANSFORMATIONS
m
X
T (v1 ) = Aa1 wa .
a=1
The role of the “extra index” 1 the coefficients A will be clear in a moment.
Repeating this argument with all of the basis vector vj ∈ β we find that for
each j, 1 ≤ j ≤ n there A1j , A2j , . . . , Am
j ∈ IR such that
m
X
(2.2) T (vj ) = Aaj wa .
a=1
and
3 1 0 0
0
T = 1 = 3 0 + 1 1 − 1 0 .
1
−1 0 0 1
Therefore
2 3
[T ]γβ = 1 1 .
1 −1
Note that the coefficients of T (e1 ) in the basis γ are in the first column,
and those of T (e2 ) are in the second column. We now compute [T ]γβ in the
following basis’
1 1 0
1 1
(2.4) β= , , γ = 2 , 0 , 1
−1 0
1 −1 0
We get,
−1 1 1 0
1 1 3
T = 0 =
2 − 0 − 1
−1 2 2
2 1 −1 0
and
2 1 1 0
1 3 1
T = 1 =
2 + 0 − 2 1 .
0 2 2
1 1 −1 0
Therefore
1 3
2 2
(2.5) [T ]γβ = − 32 1
2
−1 −2
Again note that the coefficients of T (v1 ) in the basis γ are in the first column,
and those of T (v2 ) are in the second column.
Expanding on the remark at the end of the example, the columns of [T ]γβ
are the coefficients of T (v1 ), T (v2 ), . . . , T (vn ) in the basis γ!
and therefore
[LA ]γβ = A.
The following lemma shows that the above example essentially describes
all linear transformations from IR n to IR m .
Proof. We simply expand out T (v) using equation (2.7), and using the lin-
earity of T to get
n
X n
X
i
T (v) = T ( ξ vi ) = ξ i T (vi ).
i=1 i=1
Pm
Now substituting for T (vi ) = Aai wa , we get
a=1
n m
! m n
!
X X X X
(2.8) T (v) = ξi Aai wa = Aai ξ i wa .
i=1 a=1 a=1 i=1
Since {wa }1≤a≤m is a basis the coefficients of wa in equation 2.8 must be the
same, and so
n
X
a
(2.9) η = Aai ξ i .
i=1
which agrees with T̂ on the basis β. We have therefore proved the following
lemma.
Lemma 2.1.7. Let β be a basis for the vector space V . Every linear trans-
formation T ∈ L(V, W ) uniquely determines a function T̂ : β → W . Con-
versely every function T̂ : β → W determines a unique linear transformation
T ∈ L(V, W ) which agrees with T̂ on the basis β.
ker(T ) = { v ∈ V | T (v) = 0W }
Proof. We need to show that ker(T ) satisfies conditions 1),2),3) from the
subspace Lemma 1.5.14. We begin by computing
This shows property 2) and 3) hold from Lemma 1.5.14, and so ker(T ) is a
subspace of V .
The proof that R(T ) is a subspace is left as an exercise.
w + 3x + 4y + 5z 0
The parametric solution to these equation is
v 1 0
w 1 1
x = t1 1 + t2 −2 .
y −1 0
z 0 1
Note that ker(Z) is a two-dimensional subspace with basis
1 0
1 1
βker(T ) = 1 , −2 .
−1 0
0 1
Proof. Exercise 3.
[T ]γβ = (Bja ) 1 ≤ a ≤ m, 1 ≤ j ≤ n
2.3. COMPOSITION, INVERSE AND ISOMORPHISM 35
and
[S]δγ = (Aαa ) 1 ≤ α ≤ p, 1 ≤ a ≤ m,
and
[S ◦ T ]δβ = (Cjα ) 1 ≤ α ≤ p, 1 ≤ j ≤ n
be the matrix representations of T (m × n),S (p × n) and S ◦ T (p × n).
Example 2.3.4. Let T and S be from equations (2.1) and (2.15), we then
check (2.17) using the standard basis for each space. That is we check [T ◦
S] = [T ][S] where these are the matrix representations in the standard basis.
We have
2 3
2 0 4 4
[T ] = 1 1 , [S] =
0 1 4 6
1 −1
36 CHAPTER 2. LINEAR TRANSFORMATIONS
( A | I ) → ( I | A−1 )
1. W is n-dimensional, and
2. T −1 : W → V is linear.
−1 −1
3. [T −1 ]βγ = [T ]γβ , where [T ]γβ is the inverse matrix of [T ]γβ .
Therefore,
−1 x −1 x x−y
T = [T ] = .
y y 2y − x
Double checking this answer we compute
−1 x x−y 2(x − y) + 2y − x x
T ◦T =T = =
y 2y − x x − y + 2y − x y
dim L(V, W ) = nm
Proof. Let β = {vi }{1 ≤ i ≤ n} and γ = {wa }1≤a≤m be basis for V and W
respectively. Define Φ : L(V, W ) → Mm×n (IR ) by
2.4 Exercises
1. A transformation T : IR 3 → IR 2 is defined by
x
x+y−z
(2.19) T( y ) =
.
2y + z
z
Tangent Vectors
This chapter will show why vector fields are written as differential operators
and then examine their behavior under a changes of variables.
This purely set theoretical discussion of tangent vectors does not reflect the
geometric meaning of tangency.
43
44 CHAPTER 3. TANGENT VECTORS
Example 3.1.1. For the helix σ(t) = (cos t, sin t, t) the tangent vector at
t = π4 ,is
!
dσ 1 1
(3.1) = −√ , √ , 1 .
dt t= π 2 2 ( √1 , √1 , π4 )
4 ( √1 , √1 , π4 ) 2 2
2 2
The tangent vector to α(t) at t = 0 is exactly the same as that in (3.1). Two
completely different curves can have the same tangent vector at a point.
(3.2) σ(t) = p + ta
i
(3.3) σ i (t) = xi0 + ea (t−1)
3.2 Derivations
The second geometric way to think about tangent vectors is by using the
directional derivative. Let (a)p be a tangent vector at the point p and let
σ : I → IR n be representative curve (so σ(t0 ) = p, σ̇(t0 ) = a). Recall that
the directional derivative at p of a smooth function f ∈ C ∞ (IR n ) along (a)p
and denoted Da f (p) is
d
Da f (p) = f ◦ σ
dt t=t0
n
X ∂f
i
dσ
(3.4) = i
i=1
∂x σ(t0 ) dt
t=t0
n
X ∂f
= ai ,
i
i=1
∂x p
where the chain-rule 1.4.2 was used in this computation. Using this for-
mula let’s make a few observations about the mathematical properties of the
directional derivative.
Lemma 3.2.1. Let (a)p ∈ Vp , then the directional derivative given in equa-
tion 3.4 has the following properties:
1. Da f (p) ∈ IR ,
where c ∈ IR .
46 CHAPTER 3. TANGENT VECTORS
Proof. These claims are easily verified. For example, to verify the second
property in equation 3.5,
d
Da (f g)(p) = (f ◦ σ g ◦ σ) |t=t0
dt
n
dσ i dσ i
X ∂f ∂g
= g(σ(t0 )) + f (σ(t0 )) i
i=1
∂xi σ(t0 ) dt t=t0 ∂x σ(t0 ) dt t=t0
= Da f (p)g(p) + f (p)Da g(p) by equation 3.4.
Lemma 3.2.3. With the operations 3.7 the set Der(C ∞ (p)) is a vector-space.
0p (f ) = 0.
Example 3.2.4. Let f ∈ C ∞ (p) and let Xp (f ) = ∂xi f |p , where i ∈ {1, . . . , n}.
Then Xp ∈ Der(C ∞ (IR n )). More generally, if (ξ i )1≤i≤n ∈ IR n then
Xp = (ξ 1 ∂1 + ξ 2 ∂2 + . . . + ξ n ∂n )|p
The function T takes the vector (a)p to the corresponding directional deriva-
tive. We need to check that T ((a)p ) is in fact a derivation. This is clear
though from property 4 in Lemma 3.2.1.
Therefore Xp (1U ) = 0. To prove 2, use linearity from equation 3.6 and part
1,
Xp (c) = Xp (c · 1) = cXp (1) = 0 f or all , c ∈ IR .
Proof. Let f ∈ C ∞ (p), and let U = Dom(f ) be open, and p = (x10 , . . . , xn0 ).
By Lemma 1.4.8 there exists functions gi ∈ C ∞ (p), 1 ≤ i ≤ n, defined on an
open ball Br (p) ⊂ U such that the function F : Br (p) → IR given by
n
X
(3.11) F (x) = f (p) + (xi − xi0 )gi (x), x ∈ Br (p),
i=1
where
∂f
gi (p) = ,
∂xi p
agrees with f (x) on Br (p). Since f and F agree on Br (p), Corollary 3.2.9
implies
Xp (f ) = Xp (F ).
Finally by using the properties for Xp of linearity, Leibniz rule and
Xp (f (p)) = 0 (Lemma 3.2.7) in equation 3.11 we have
n
!
X
Xp (f ) = Xp (F ) = Xp (xi − xi0 )gi (x)
i=1
n
X
(3.12) = Xp (xi − xi0 )gi (p) + (xi − xi0 )|x=p Xp (g i )
i=1
Xn
= Xp (xi )gi (p).
i=1
Property 2 in Lemma 1.4.8 gives gi (p) = (∂xi f )|p which in equation 3.12 gives
equation 3.9 and the theorem is proved.
Corollary 3.2.11. The set β = {∂xi |p }1≤i≤n forms a basis for Der(C ∞ (p)).
Corollary 3.2.11 leads to the modern definition of the tangent space.
Definition 3.2.12. Let p ∈ IR n . The tangent space at p denoted by Tp IR n
is the n-dimensional vector-space Der(C ∞ (p).
The linear transformation T : Vp → Tp IR n given in equation 3.8 is then an
isomorphism on account of Lemma 3.2.6 and the dimension theorem (2.2.7).
If σ : I → IR n is a curve with tangent vector (σ̇(t0 ))σ(t0 ) then the correspond-
ing derivation Xp = T ((σ̇(t0 ))σ(t0 ) is
n
dσ i
X ∂
(3.13) Xp = .
i=1
dt t=t0 ∂xi p=σ(t0 )
50 CHAPTER 3. TANGENT VECTORS
Example 3.2.13. Let p = (1, 1, 1) ∈ IR 3 and let (1, −2, 3)p ∈ Vp . The curve
is a representative curve for (1, −2, 3)p . The corresponding tangent vector
Xp ∈ Tp IR 3 is
Xp = (∂x − 2∂y + 3∂z )p .
and
Xp (xy) = = (y)|(1,2) Xp (y) + (x)|(1,2) Xp (y)
= 2Xp (x) + Xp (y)
Using Xp (x2 + y 2 ) = 3, Xp (xy) = 1 this gives the system of equations for
a = Xp (x) and b = Xp (y) to be
2 4 a 3
=
2 1 b 1
Therefore
1 2
Xp = ∂x + ∂y
6 3 p
(3.14) Xp (f 1 ) = c1 , Xp (f 2 ) = c2 , . . . , Xp (f n ) = cn .
3.3. VECTOR FIELDS 51
Jξ = c
We will write this as X = ni=1 ξ i (x)∂xi (dropping the |x ) since the point
P
at which ∂xi is evaluated can be inferred from the coefficients. The vector
field X is smooth or a C ∞ vector field if the coefficients ξ i (x) are smooth
functions. Vector fields will play a prominent role in the rest of the book.
X = xy 2 ∂x + xz∂y + ey+z ∂z
is smooth.
Vector fields have algebraic properties that are similar to tangent vectors.
Let U ⊂ IR n be an open set. A function X : C ∞ (U ) → C ∞ (U ) is called a
derivation of C ∞ (U ) if for all f, g ∈ C ∞ (U ) and a, b ∈ IR ,
Example 3.3.2. Consider ∂xi where i ∈ {1, . . . , n}, and C ∞ (IR n ). The
partial derivatives ∂xi satisfy properties one and two in equation 3.16, and
so ∂xi ∈ Der(C ∞ (IR n )). More generally, let
3.4 Exercises
1. Let (σ̇(t0 ))σ(t0 ) be the tangent vector to the curve at the indicated
point as computed in section 3.1. Give two other representative curves
for each resulting tangent vector (different than in equations 3.2 and
3.3), and give the corresponding derivation in the coordinate basis (see
example 3.2.13).
(a) x = t3 + 2 , y = t2 − 2 at t = 1,
(b) r = et , θ = t at t = 0,
(c) x = cos(t), y = sin(t) , z = 2t , at t = π/2.
7. Let f1 = x + y + z, f2 = x2 + y 2 + z 2 , f3 = x3 + y 3 + z 3 be functions
on IR 3 .
3.4. EXERCISES 55
f = x2 yz, g = y2 − x + z2,
compute
10. Another common way to define tangent vectors uses the notion of
germs. On the set C ∞ (p) define two functions f, g ∈ C ∞ (p) to be equiv-
alent if there exists an open set V with p ∈ V , and V ⊂ Dom(f ), V ⊂
Dom(g) such that f (x) = g(x), f or allx ∈ V . Show that this is an
equivalence relation on C ∞ (p). The set of equivalence classes are called
germs of C ∞ functions at p. The tangent space Tp IR n is then defined
to be the vector-space of derivations of the germs of C ∞ functions at
p.
56 CHAPTER 3. TANGENT VECTORS
Chapter 4
57
58 CHAPTER 4. THE PUSH-FORWARD AND THE JACOBIAN
The map Φ∗,p we define below has the property that if (σ̇(t0 ))σ(t0 ) is the
tangent vector to the curve σ at p = σ(t0 ) with corresponding derivation Xp
as in equation 3.13, then Yq = Φ∗,p Xp is the derivation Yq corresponding to
the tangent vector Φ ◦ σ at the image point Φ(p) = Φ(σ(t0 ))! In example
4.1.1 this means with p = (1, 2) and Xp = (3∂x − 2∂y )(1,2) from 4.2, and
Yq = (−2∂u + 14∂v + 4∂w )(5,−3,2) from 4.3, that Yq = Φ∗,p Xp .
Let Xp ∈ Tp IR n and let σ : I → IR n be a smooth curve which represents
Xp by σ̇(t0 ) = Xp . By composing Φ : IR n → IR m with σ we get the image
curve Φ ◦ σ : I → IR m from which we prove
For the curve Φ(σ(t)), let Yq be the tangent vector at q = Φ(σ(t0 )) which is
n n n !
dΦi (σ(t)) ∂Φi dσ j )
X ∂ X X ∂
(4.6) Yq = i
=
i=1
dt
t=t0
∂y q i=1 j=1
∂xj p dt t=t0 ∂y i .
and show that the result is a derivation (see 3.2.2). Before giving the def-
inition we make the following simple observation. Let g ∈ C ∞ (q) which is
a function on the image space of Φ. The composition g ◦ Φ : IR n → IR ,
is a smooth function on an open subset of IR n which contains p, and so
Xp (g ◦ Φ) is well defined! Using this calculation, we are now ready to define
Φ∗,p (Xp ).
Theorem 4.2.1. Let Φ : IR n → IR m be a smooth function, let p ∈ IR n and
q = Φ(p). Given Xp ∈ Tp IR n define Φ∗,p (Xp ) : C ∞ (q) → IR by
Yq (g · h) = Xp (g ◦ Φ · h ◦ Φ)
= Xp (g ◦ Φ) · h ◦ Φ(p) + g ◦ Φ(p) · Xp (h ◦ Φ)
= Yq (g) · h(q) + g(q) · Yq (h).
then
m n a
!
X X
i ∂Φ ∂
(4.9) Φ∗,p Xp = ξ
a=1 i=1
∂xi p ∂y a q
Proof. By definition we have
n
!
X
(4.10) (Φ∗,p Xp )(g) = ξ i ∂xi |p (g ◦ Φ)
i=1
If we expand out the derivative term in 4.10 using the chain rule we get,
m
∂g ∂Φa
∂g ◦ Φ X
= .
∂xi p a=1 ∂y a q ∂xi p
where in the second line we have switched the order of the summations. By
taking the coefficients of ∂ya |q equation 4.9 now follows directly from 4.11.
∂Φa
(4.12) [Φ∗,p ] =
∂xi p
This entries in the matrix Jia are easily determined using equation 4.9 which
gives
m
∂Φa
X
(4.14) Φ∗,p (∂xi |p ) = i
∂ya |q ,
a=1
∂x p
therefore
∂Φa
(4.15) [Φ∗,p ] = .
∂xi p
Letp = (1, 2) and Xp = (3∂x − 2∂y )|p . Compute Φ∗,p Xp first using derivation
approach 4.7, and then the Jacobian (4.15.
With q = Φ(p) = (5, −3, 2) we now find the coefficients of Yq = (a∂u +
b∂v +c∂w )|q = Φ∗,p Xp by using the definition 4.7 in the form of equation 4.16,
This gives
We now compute Φ∗,p Xp using the Jacobian matrix at (1, 2), which is
2x 2y 2 4
∂Φa
= 2x −2y = 2 −4 ,
∂xi (1,2) y x (1,2) 2 1
This gives the same coefficients for Φ∗,p Xp in the coordinate basis {∂u |q , ∂v |q , ∂w |q }
as in equation (4.18).
Remark 4.2.7. Note that the definition of Φ∗,p in 4.7 was given independent
of coordinates! Then the coordinate dependent formulas 4.16 and 4.12 for
Φ∗,p were derived from its definition. This is what we strive for when giving
definitions in differential geometry.
4.3. THE CHAIN RULE, IMMERSIONS, SUBMERSIONS, AND DIFFEOMORPHISMS63
The set T IR n consists of all possible tangent vectors at every possible point
and an element of T IR n is just a tangent vector Xp at a particular point
p ∈ IR n . We now define the map Φ∗ : T IR n → T IR m by the point-wise
formula
This formula agrees with the point-wise given formula in equation 3.13.
(Ψ ◦ Φ)∗,p Xp (g) = Xp (g ◦ Ψ ◦ Φ) .
64 CHAPTER 4. THE PUSH-FORWARD AND THE JACOBIAN
We have shown (Ψ ◦ Φ)∗ Xp (g) = ((Ψ∗ ) ◦ (Φ∗ )Xp ) (g) for all g ∈ C ∞ (r).
Therefore (Ψ ◦ Φ)∗ Xp = ((Ψ∗ ) ◦ (Φ∗ )) Xp for all Xp ∈ T IR n and the theorem
is proved.
Recall in section 3.1 that the rank of a linear transformation T is the
dimension of R(T ), the range space of T .
Let Φ : U → V , where U ⊂ IR n and V ⊂ IR m be a C ∞ function.
Definition 4.3.3. The function Φ is an immersion at p ∈ U if the rank
of Φ∗,p : Tp U → TΦ(p) V is n. The function Φ is an immersion if it is an
immersion at each point in p ∈ U .
By fact that dim Tp U is n-dimensional, the dimension Theorem 2.2.7
shows that the map Φ∗ is an immersion if it is injective at each point p ∈ U .
This also implies by exercise 2 in the section 3 that the kernel Φ∗,p at each
point p ∈ U consists of only the zero tangent vector at p.
Example 4.3.4. Let Φ : IR 2 IR 3 be given by
According to Lemma 2.2.4 (see also exercise 2), Φ∗ is injective if and only if
ker[Φ∗ ] is the zero vector in IR 2 . This fails to happen only when x = y = 0.
Therefore Φ is an immersion at every point except the origin in IR 2 .
Definition 4.3.5. The function Φ is a submersion at p ∈ U if the rank
of Φ∗,p : Tp U → TΦ(p) V is m. The function Φ is an submersion if it is a
submersion at each point in p ∈ U .
4.3. THE CHAIN RULE, IMMERSIONS, SUBMERSIONS, AND DIFFEOMORPHISMS65
f (t) = t3
then
1
f −1 (t) = t 3 .
y i = Φi (x1 , . . . , xn ), 1 ≤ i ≤ n.
If p ∈ U and it has x-coordinates (x10 , . . . , xn0 ), then y0i = Φi (x10 , . . . , xn0 ) are
the y-coordinates of Φ(p).
Since Φ is a diffeomorphism, then Φ−1 : V → U is a diffeomorphism.
is a change of coordinates.
is a change of coordinates.
If we do this for every point p ∈ U , and use the fact that Φ is a diffeomorphism
(so one-to-one and onto) we find that Φ∗,p applied to X defines a tangent
vector at each point q ∈ V , and therefore a vector-field. To see how this
is defined, let q ∈ V , and let p ∈ U be the unique point in U such that
p = Φ−1 (q). We then define the vector-field Y point-wise on V by
Yq = Φ∗,p (Xp ).
4.4. CHANGE OF VARIABLES 67
Let’s give a coordinate version of this formula. Suppose (xi )1≤i≤n are coor-
dinates on U , and (y i )1≤i≤n are coordinates on V . We have a vector-field
n
X ∂
X= ξ i (x) ,
i=1
∂xi
Let p ∈ U , and q = Φ(p) the formula we have is Yq = (Φ∗,p Xp )p=Φ−1 (q) and
by equation 4.9 is
n n !
∂Φj
X X
i ∂
(4.25) Yq = ξ (p) .
∂xi p −1 ∂y j q
j=1 i=1 p=Φ (q)
or by equation 4.16
The formulas 4.26 and 4.27 for Y is called the change of variables formula
for a vector-field. The vector-field Y is also called the push-forward by Φ of
the vector-field X and is written Y = Φ∗ X.
Φ∗ X = a∂x + b∂y
68 CHAPTER 4. THE PUSH-FORWARD AND THE JACOBIAN
Similarly we find
Similarly we have
n −1 i
X ∂(Φ ) ∂
(4.32) Φ−1
∗ (∂y j ) = j
i
.
i=1
∂y
Φ(x) ∂x
4.4. CHANGE OF VARIABLES 69
Similarly we get
Φ∗ ∂θ = −y∂x + x∂y ,
p
and using ρ sin φ = x2 + y 2 we get
z p
Φ∗ ∂φ = p (x∂x + y∂y ) − x2 + y 2 ∂z .
x2 + y 2
X = Aρ ∂ρ + Aθ ∂θ + Aφ ∂φ
Aρ
Φ∗ (Aρ ∂ρ + Aθ ∂θ + Aφ ∂φ ) = p (x∂x + y∂y + z∂z ) +
x2 + y 2 + z 2
z p
+ Aθ (−y∂x + x∂y ) + Aφ ( p (x∂x + y∂y ) − x2 + y 2 ∂z ) =
x2 + y 2
!
xAρ xzAφ
p − yAθ + p ∂x +
x2 + y 2 + z 2 x2 + y 2
!
yAρ yzAφ
p + xAθ + p ∂y +
x2 + y 2 + z 2 x2 + y 2
!
zAρ p
p − x2 + y 2 Aφ ∂z .
2
x +y +z2 2
xAρ xzAφ
Ax = p − yAθ + p
x2 + y 2 + z 2 x2 + y 2
yAρ yzAφ
Ay = p + xAθ + p
2
x +y +z2 2 x2 + y 2
zAρ p
Az = p − x2 + y 2 Aφ ,
x2 + y 2 + z 2
y2 y
Φ(x, y, z) = (x − , z, ).
2z z
where V = { (u, v, w) ∈ IR 3 | v 6= 0 }. We write the vector-field
X = y∂x + z∂y .
y2
Φ∗ (X)(u) = (y∂x + z∂y )(x − )=0
2z
Φ∗ (X)(v) = (y∂x + z∂y )(z) = 0
y
Φ∗ (X)(u) = (y∂x + z∂y )( ) = 1.
z
Therefore
Φ∗ X = ∂w .
4.5 Exercises
1. For each of the following maps and derivations compute Φ∗,p (X) using
the definition of Φ∗,p in terms of derivations defined in Theorem 4.2.1
(see Example 4.2.6).
3. Let
Φ(x, y) = (u = x3 + 3xy 2 , v = 3x2 y + y 3 ),
and let p = (1, −1). Compute the matrix representation of Φ∗,p :
Tp IR 2 → TΦ(p) IR 2 with respect to the basis {X1 , X2 } for the tangent
space for Tp R2 and {Y1 , Y2 } for TΦ(p) R2 .
(a) X1 = ∂x |p , X2 = ∂y |p , Y1 = ∂u |q , Y2 = ∂v |q .
(b) X1 = (∂x − 2∂y )|p , X2 = (∂x + ∂y )|p , Y1 = ∂u |q , Y2 = ∂v |q .
(c) X1 = (∂x − 2∂y )|p , X2 = (∂x + ∂y )|p , Y1 = (2∂u + 3∂v )|q , Y2 =
(∂u + 2∂v )|q .
∂ ∂
4. (a) Write the tangent vector Xp = − p = (1, 1), in polar
∂x p ∂y p
coordinates using the coordinate basis.
∂ ∂
(b) Write the tangent vector Xp = −2 +3 , p = (r = 1, θ =
∂r ∂θ
p p
π/3) given in polar coordinates, in Cartesian coordinates using
the coordinate basis.
(c) Let
1 2 2
Φ(u, v) = uv, (v − u )
2
4.5. EXERCISES 73
be the change
to parabolic
coordinates. Write the tangent vector
∂
∂
Xp = − , p = (u = 1, v = 1) given in parabolic
∂u p
∂v p
coordinates, in Cartesian coordinates using the coordinate basis.
(d) Write the tangent vector Xp in part c) in polar coordinates using
the coordinate basis.
5. Check the chain rule (Ψ ◦ Φ)∗,p = Ψ∗,p ◦ Φ∗,p for each of the following
maps at the point indicated. You may calculate in the coordinate basis.
6. Find the inverses of the following maps Φ and check that (Φ∗,p )(−1) =
(Φ(−1) )∗,p at the indicated point.
7. Find the points, if any, in the domain of the following maps about
which the map fails to be a local diffeomorphism.
(a) Φ(u, v) = (u + v, u − v, u2 + v 2 )
(b) Φ(u, v, w) = (u+v+w, u2 +v 2 +w2 , u3 +v 3 +w3 , u4 +v 4 +w4 ), w>
v > u.
(a) Φ(x, y, z) = (x + y − z, x − y)
74 CHAPTER 4. THE PUSH-FORWARD AND THE JACOBIAN
ordinates.
75
76CHAPTER 5. DIFFERENTIAL ONE-FORMS AND METRIC TENSORS
Then
n
X n
X
i
τ (vj ) = α(vi )α (vj ) = α(vi )δji = α(vj ).
i=1 i=1
Given the basis β the basis {αi }1≤i≤n for V ∗ in 5.3 is called the dual basis.
78CHAPTER 5. DIFFERENTIAL ONE-FORMS AND METRIC TENSORS
(5.8) Xp (f ) ∈ IR .
If Yp ∈ Tp IR n then
Example 5.1.8. On IR 3 ,
α = ydx − xdy + zydz
is a one-form field. At the point p = (1, 2, 3),
αp = 2dxp − dyp + 6dzp
5.2. BILINEAR FORMS AND INNER PRODUCTS 81
These equations imply that B is linear as a function of each of its two ar-
guments. A function B which satisfies these conditions is called a bilinear
form.
B(ei , ej ) = Aij
Then
n
X
(5.25) B(v, w) = Bij ai bj = [v]T (Bij )[w],
i,j=1
where [v] = [ai ], [w] = [bi ] are the column vectors of the components of v and
w in (5.24).
Proof. We simply expand out using (5.24), and the bi-linearity condition,
n n
!
X X
B(v, w) = B ai vi , bj vj
i=1 j=1
n n
! n X
n
X X X
= ai B vi , bj vj = ai bj B (vi , wj ) .
i=1 j=1 i=1 j=1
Example 5.2.6. Let V = IR n and let Qij be any symmetric matrix. Define
n
X
B(x, y) = Qij xi y j = xT Qy.
i,j=1
2. γ(v, v) ≥ 0,
n
X
||v|| = (ai )2 .
i=1
Let β = {vi }1≤i≤n be a basis for V and let γ is an inner product. The
n × n matrix [gij ] with entries
(5.28) gij = γ(vi , vj ) , 1 ≤ i, j ≤ n,
is the matrix representation 5.20 of the bilinear form γ in the basis β. Prop-
erties 1 in 5.2.8 implies the matrix [gij ] is symmetric.
Property 2 and 3 for γ in definition 5.2.8 can be related to properties of
its matrix representation [gij ]. First recall that A real symmetric matrix is
always diagonalizable over the IR (Theorem 6.20 of [6]). This fact gives a
test for when a symmetric bilinear form γ is positive definite in terms of the
eigenvalues of a matrix representation of γ.
Theorem 5.2.11. Let γ be a bilinear form on V and let gij = γ(vi , vj ) be the
coefficients of the matrix representation of γ in the basis β = {vi }1≤i≤n . The
bilinear form γ is an inner product if and only if its matrix representation
[gij ] is a symmetric matrix with strictly positive eigenvalues.
The property that [gij ] is symmetric with positive eigenvalues does not
depend on the choice of basis.
Example 5.2.12. Consider again the bilinear form β : IR 3 × IR 3 → IR from
example 5.2.2,
1 1
x y
B 2 2
x , y = 2x1 y 1 − x1 y 2 + x1 y 3 − x2 y 1 − x2 y 2 + x3 y 1 + x3 y 3 .
x3 y3
In the basis in equation 5.21 the matrix representation is
1 0 0
[Bij ] = 0 −1 0 .
0 0 2
86CHAPTER 5. DIFFERENTIAL ONE-FORMS AND METRIC TENSORS
= Bkl .
Here we have used αi (vk ) = δki , αj (vl ) = δlj . Therefore by corollary 5.2.5
B = C, and so ∆ is a spanning set. The proof that ∆ is a linearly independent
set is left for the exercises.
Note that formula 5.30 for bilinear forms is the analogue to formula 5.7
for linear function.
Example 5.3.3. Let B : IR 3 × IR 3 → IR be the bilinear form from example
5.2.2. Let {αi }1≤i≤3 be the dual basis of the basis in equation 5.21. Then by
Theorem 5.3.2 and equation 5.22 we have
(5.31) B = α1 ⊗ α1 − α2 ⊗ α2 + 2α3 ⊗ α3 .
Let β = {vi }1≤i≤n be a basis for V and ∆ = {αi }1≤i≤n be a basis for the
dual space V ∗ and let B be symmetric bilinear form on V . Theorem 5.3.2
allows us to write the inner product B as in equation 5.30 we have
X
(5.32) B= Bij αi ⊗ αj ,
1≤i,j≤n
88CHAPTER 5. DIFFERENTIAL ONE-FORMS AND METRIC TENSORS
where Bij = B(vi , vj ). In equation 5.32 we note that by using the symmetry
Bij = Bji we can write
Bij αi ⊗ αj + Bji αj ⊗ αi = Bij (αi ⊗ αj + αj ⊗ αi ).
Therefore we define
1
αi αj = (αi ⊗ αj + αj ⊗ αi ),
2
so that equation 5.32 can be written
X
(5.33) B= Bij αi αj .
1≤i,j≤n
This notation will be used frequently in the next section. Equation 5.31 of
example we can be rewritten as
B = α1 ⊗ α1 − α2 ⊗ α2 + 2α3 ⊗ α3 = (α1 )2 − (α2 )2 + 2(α3 )3 .
Example 5.3.4. Let V be a vector-space and β = {vi }1≤i≤n a basis and
β ∗ = {αj }1≤j≤n the dual basis. Let
n
X n
X
i i
γ= ci α ⊗ α = ci (αi )2 ci ∈ IR , ci > 0.
i=1 i=1
If we use the dual basis {dxi |p }1≤i≤n , then equation 5.30 in Theorem 5.3.2
says
X
γ(p) = gij dxi |p ⊗ dxj |p
1≤i,j≤n
(5.34) X
= gij dxi |p dxj |p by equation 5.33.
1≤i,j≤n
γp : Tp IR n × Tp IR n → IR
A smooth metric tensor is one where the functions gij (x) are C ∞ functions
on IR n . Using {dxi |p }1≤i≤n as the dual basis for ∂xi |p , applying the formula
5.34 pointwise for γ using equation 5.35 yields
X
(5.36) γ= gij (x) dxi dxj .
1≤i,j≤n
Note that at each fixed point p ∈ IR n , the matrix [gij (p)] will be symmetric
and positive definite (by Theorem 5.2.11) because γp is an inner product.
Conversely we have the following.
90CHAPTER 5. DIFFERENTIAL ONE-FORMS AND METRIC TENSORS
is a metric tensor on IR n .
This theorem then states that every metric tensor is of the form 5.37. The
functions gij (x) are be called the components of γ in the coordinate basis,
or the components of γ. Pn
Given a metric tensor γ = g (x)dxi dxj , a point p ∈ IR n and
Pn i Pn i,j=1i ij
vectors Xp = i=1 ξ ∂xi |p , Yp = i=1 η ∂xi |p , Theorem 5.2.4 gives,
n
X
T
(5.38) γp (Xp , Yp ) = [Xp ] [gij (p)][Yp ] = gij (p)ξ i η j
i,j=1
If X = z∂x + y∂y + x∂z and Y = y∂x − x∂y then equation 5.39 gives
yz y y
γ(X, Y ) = − + .
x2 x z
Example 5.4.4. Let γ be a metric tensor on IR 2 . By equation 5.36 there
exists functions E, F, G ∈ C ∞ (IR 2 ) such that
γ = E dx ⊗ dx + F (dx ⊗ dy + dy ⊗ dx) + G dy ⊗ dy
(5.41)
= E(dx)2 + 2F dxdy + G(dy)2 = Edx2 + 2F dxdy + Gdy 2 .
1 dx2 + dy 2
γ= (dx ⊗ dx + dy ⊗ dy) = .
y2 y2
The real number γp (Xp , Yp ) can be computed using formula 5.38 or by ex-
panding,
1 1 1 5
γ(Xp , Yp ) = (dx(Xp )dx(Xp ) + dy(Xp )dy(Xp )) = (2)(−1)+ (−3)(1) = −
4 4 4 4
If the point were p = (2, 5) with Xp , Yp from equation 5.45 we would have
1 1 1
γ(Xp , Yp ) = (2)(−1) + (−3)(1) = − .
25 25 5
Notice how the computation depends on the point p ∈ U .
This metric tensor is not a Riemannian metric tensor. It does not satisfy the
positive definite criteria as a symmetric bilinear form on the tangent space
at each point. It is however non-degenerate.
The Einstein equations in general relativity are second order differential
equations for the coefficients of a metric [gij ], 1 ≤ i, j, ≤ 4. The differen-
tial equations couple the matter and energy in space and time together with
the second derivatives of the coefficients of the metric. The idea is that the
distribution of energy determines the metric tensor [gij ]. This then deter-
mines how things are measured in the universe. The Scwartzschild metric
represents the geometry outside of a fixed spherically symmetric body.
Remark 5.4.9. You need to be careful about the term metric as it is used
here (as in metric tensor). It is not the same notion as the term metric in
topology! Often in differential geometry the term metric is used, instead of
the full name metric tensor, which further confuses the issue. There is a
relationship between these two concepts - see remark 5.4.13 below.
5.4.1 Arc-length
Pn
Let γ = i,j=1 gij (x)dxi dxj be a metric tensor on IR n , and σ : [a, b] → IR n
a continuous curve on the interval [a, b], and smooth on (a, b). Denote by
σ(t) = (x1 (t), x2 (t), . . . , xn (t)) the components of the curve. At a fixed value
of t, σ̇ is the tangent vector (see equation 3.13), and it’s length with respect
to γ is compute using 5.38 to be
v
u n
uX
p dxi dxj
γ(σ̇, σ̇) = t gij (σ(t)) .
i,j=1
dt dt
which is the arc-length of σ with respect to the metric γ. Note that the
components gij (x) of the metric tensor are evaluated along the curve σ.
94CHAPTER 5. DIFFERENTIAL ONE-FORMS AND METRIC TENSORS
as expected.
Example 5.4.11. Compute the arc-length of the line
σ(t) = (t + 1, 2t + 1) 0 ≤ t ≤ 1
This is not the same as the arc-length using the Euclidean metric tensor
which is found to be Z 1
√ √
1 + 4dt = 5.
0
γ(ui , uj ) = δij , 1 ≤ i, j ≤ n.
w 1 = v1 ,
j−1
X γ(vj , wk )
w j = vj − wk , 2 ≤ j ≤ n.
k=1
γ(wk , vw )
Theorem 5.4.14. The set of vector β̃ form a basis for V and are orthogonal
to each other
γ(wi , wj ) = 0 i 6= j.
The proof of this is standard see [6]. From the set β 0 the final step in the
Gram-Schmidt process is to let
1
ui = p wi 1≤i≤n
γ(wi , wi )
We now show how to construct orthonormal frame fields using the Gram-
Schmidt process. Let γ be a Riemannian metric tensor on IR n . Construct
the following vector fields from the frame field β 0 = {Xi }1≤i≤n ,
Y1 = X1 ,
j−1
(5.48) X γ(Xj , Yk )
Yj = Xj − Yk , 2 ≤ j ≤ n.
k=1
γ(Yk , Yk )
As in Theorem 5.4.14
Theorem 5.4.15. The vector-fields {Yi }1≤i≤n are smooth and linearly inde-
pendent at each point in IR n , and are mutually orthogonal at each point
γ(Yi , Yj ) = 0, i 6= j.
Again as in the final step of the Gram-Schmidt process let
1
Zi = p Yi 1 ≤ i ≤ n,
γ(Yi , Yi )
and the set β = {Zi }1≤i≤n is a set of vector-fields that form an orthonormal
basis with respect to γp for Tp IR n for each point p ∈ IR n , and so form an
orthonormal frame field.
Example 5.4.16. Let γ be the metric tensor on U = {(x, y, z) ∈ IR 3 | xz 6=
0} given by
2
1 2 1 2 y y 1
γ = 2 dx + 2 dy − 2 2 dydz + + dz 2 .
x x zx x2 z 2 z 2
We find an orthornomal frame field starting with the coordinate frame ∂x , ∂y , ∂z
and then using equation 5.48. The first vector-field is Y1 = ∂x , the second is
γ(∂y , Y1 )
Y2 = ∂y − Y1 = ∂ y ,
γ(Y1 , Y1 )
and the third is
γ(∂z , Y1 ) γ(∂z , Y2 )
Y3 = ∂z − Y1 − Y2
γ(Y1 , Y1 ) γ(Y2 , Y2 )
y
= ∂z − 0∂x + ∂y .
z
Finally the resulting orthonormal frame field is
(5.49) Z1 = x∂x , Z2 = x∂y , Z3 = z∂z + y∂y .
5.5. RAISING AND LOWERING INDICES AND THE GRADIENT 97
The form αv = TB (v) in equation 5.55 is called the dual of v with respect to
B, and this process is also sometimes called “lowering the index” of v with
B.
Another way to compute TB (v) is given by the following corollary (see
Theorem 5.3.2).
Corollary 5.5.2. Let
n
X
B= gij αi ⊗ αj
i,j=1
Let βP= {vi }1≤i≤n be a basis for V , β ∗ = {αi }1≤i≤n the dual basis, and
let α = ni=1 bi αi . We now find v = TB−1 (α) in the basis β. By Proposition
∗ ∗
5.5 [TB ]ββ = [gij ]. Using Proposition 2.3.7, [TB−1 ]ββ ∗ = ([TB ]ββ )−1 , and so let
∗
[g ij ] denote the inverse matrix of [gij ] = [TB ]ββ . Utilizing Lemma 2.1.6 the
coefficients aj of the image vector TB (α) in terms of the coefficients bi of α
are given by
n
X
j
(5.59) a = g ij bi ,
i=1
and
n n
!
X X
(5.60) v = TB−1 (α) = g ij bi vj .
j=1 i=1
The vector v in formula 5.59 is called the dual of α with respect to B. The
process is also sometimes called “raising the index” with B.
Example 5.5.5. We continue with the example 5.5.3 and we compute the
dual of α(xe1 + ye2 + ze3 ) = 3x − y + 2z with respect to B (raise the index).
From equation 5.57, we find the matrix [g ij ] = [gij ]−1 is
1 −1 −1
1
[g ij ] = −1 −3 1 .
4
1 −1 3
Therefore
1
[TB−1 (α)] = [1, 1, 5]
2
and
1 1 5
TB−1 (α) = e1 + e2 + e3 .
2 2 2
5.5. RAISING AND LOWERING INDICES AND THE GRADIENT 101
then formula 5.55 or equation 5.56 applied point-wise for lowering the index
gives
(5.62) ! !
Xn Xn Xn X n
αX = Tγ (X) = gij (x)dxi (X) dxj = gij (x)ξ i (x) dxj .
j=1 i=1 j=1 i=1
Example 5.5.7. Continuing from Example 5.5.6 with the metric tensor γ =
(1 + ex )−1 (dx2 + dy 2 ) on IR 2 , if f ∈ C ∞ (IR 2 ) then df = fx dx + fy dy and its
dual with respect to γ is
grad(f ) = (1 + ex )fx ∂x + (1 + ex )fy ∂y .
5.5. RAISING AND LOWERING INDICES AND THE GRADIENT 103
Example 5.5.9. If
n
X
γE = (dxi )2
i=1
In this case the coefficients of the gradient of a function are just the partial
derivatives of the function. This is what occurs in a standard multi-variable
calculus course.
Up (f ).
The result follows by noting the maximum rate ||gradp (f )||γ of |Up (f )| is
obtained when Up = ||gradp (f )||−1
γ gradp (f ).
This next bit is not necessary for doing the assignment but is another
way to define the raising the index procedure. Since Tγ is an invertible
linear transformation it can be used to define a non-degenerate bilinear form
γ ∗ : V ∗ × V ∗ → IR , defined by
γ ∗ (αi , αj ) = g ij
Tγ ∗ (α)(τ ) = γ ∗ (α, τ ).
αi (vj ) = δji , 1 ≤ i, j ≤ n.
σ i = Tγ (vi ), 1 ≤ i ≤ n.
Proof. The fact that β̃ ∗ is a basis will be left as an exercise. Suppose that β
is an orthornormal basis, then
Theorem 5.6.1 has an analogue for frames fields. Let {Xi }1≤i≤n be a
frame field on IR n . The algebraic dual equations
Corollary 5.6.2. The one-form fields {σ i }1≤i≤n from equation 5.67 define a
basis for Tp∗ IR n for each point p ∈ IR n . The fields satisfy αi = σ i , 1 ≤ i ≤ n,
if and only if Xi is an orthonormal frame field.
5.7 Exercises
1. Let V = IR 3 and T ∈ V ∗ . Prove there exists a, b, c ∈ IR such that
x
T y = ax + by + cz.
z
2. Let Xp , Yp ∈ Tp IR 2 , with p = (1 − 2) be
4. Show that B(V ) the space of bilinear functions on V with addition and
scalar multiplication defined by equation 5.19 is a vector-space. (Do
not assume that V is finite dimensional.)
B(v, w) = 0 f or all w ∈ V
then v = 0.
η(x, y) = x1 y 2 + x2 y 1 ,
where x = (x1 , x2 ), y = (y 1 , y 2 ).
11. Let γ be a symmetric bilinear forms and β = {vi }1≤i≤n a basis for V ,
and β ∗ = {αj }1≤j≤n a basis for V ∗ .
(a) compute the dual of the vector-field z∂x + y∂y + x∂z with respect
to γ (lower the index).
(b) Compute the dual of the differential form ydx + zdy − (1 + x2 )dz
with respect to the metric γ (raise the index).
(c) Find an orthonormal frame field and its dual. (Hint: See Corollary
5.6.2)
(a) Compute
√ the arc-length of a “straight line” between the points
(0, 2) and (1, 1).
(b) Compute
√ the arc-length of a circle passing through the points
(0, 2) and (1, 1) which has its center on the x-axis. Compare to
part (a). Hint: You will need to find the circle.
(c) Find an orthonormal frame field and its dual.
(d) Find the gradient of f ∈ C ∞ (U ).
111
112 CHAPTER 6. THE PULLBACK AND ISOMETRIES
Proof. The first part of this lemma is proved by showing that T t (τ ) is a linear
function of v in equation 6.1. Suppose v1 , v2 ∈ V , and c ∈ IR , then
T t (τ )(cv1 + v2 ) = τ (T (cv1 + v2 ))
= τ (cT (v1 ) + T (v2 ))
(6.2)
= cτ (T (v1 )) + τ (T (v2 ))
= cT t (τ )(v1 ) + T t (τ )(v2 ).
Now let β ∗ = {αi }1≤i≤n be the basis of V ∗ which is the dual basis of β
for V , and γ = {τ a }1≤a≤n the basis of W ∗ which is the dual basis of γ for
W ∗ as defined in equation 5.3. The matrix representation of function the∗
T t : W ∗ → V ∗ will now be computed in the basis γ ∗ , β ∗ . Let B = [T t ]βγ ∗
which is an n × m matrix, which is determined by
n
X
t a
(6.4) T (τ ) = Bia αi .
i=1
In other words the coefficients of the image [T t (τ )]β ∗ are the row vector we
get by multiplying A on the left by the row vector [τ ]γ ∗ = [c1 , . . . , cm ],
[T t (τ )]β ∗ = [τ ]γ ∗ A.
Let β = {∂xi |p }1≤i≤n be the coordinate basis for Tp IR n , and γ = {∂ya |q }1≤a≤m
the coordinate basis for Tq IR n , and the corresponding dual basis are {dxi |p }1≤i≤n ,
and {dy a |q }1≤a≤m . The matrix representation of Φ∗,p in the coordinate basis
is
m
∂Φa
X
Φ∗,p (∂xi |p ) = ∂ya |q ,
a=1
∂xi p
and so equation 6.6 gives
n
∂Φa
X
(6.9) Φ∗q (dy a |q ) = i
dxi |p .
i=1
∂x p
Note the difference in the summation index in these last two equations.
An important observation from equation 6.9 needs to be made. Equation
5.16 is
n
∂Φa
X
a
(dΦ )p = dxi |p ,
∂xi i=1 p
(6.11) Φ∗ g = g ◦ Φ,
and Φ∗ g ∈ C ∞ (IR n ).
and let τq = (2du − 3dv + dw)q where q = Φ(1, 2) = (3, −3, 2). We compute
Φ∗q τq by first using 6.10,
Φ∗q (2du − 3dv + dw)|q = 2Φ∗q (du|q ) − 3Φ∗q (dv|q ) + Φ∗q (dw|q )
= 2(dx + dy) − 3(2dx − 4dy) + 2dx + dy
= (−2dx + 15dy)(1,2) .
Finally we give a fundamental theorem relating the pull back and the
exterior derivative.
Theorem 6.1.5. Let Φ : IR n → IR m be a smooth function. Then
where g ∈ C ∞ (IR m ).
This is states that Φ∗ and d commute. A proof of this is easily given by
writing both sides out in coordinates, but here is an alternative one which is
the standard proof in differential geometry.
Proof. To prove equation 6.37 holds, let p ∈ IR n and Xp be in Tp IR n . We
then show
Next we expand the right hand side of equation 6.21 using 6.8, 5.10 , and 4.7
to get
The equality of equations 6.22 and 6.23 proves equation 6.37 holds.
(T t (B)) = AT (B)A.
Now let β ∗ = {αi }1≤i≤n be the basis of V ∗ which is the dual basis of β
for V , and γ = {τ a }1≤a≤n the basis of W ∗ which is the dual basis of γ for
W ∗ . Using equation 6.25 and the tensor product basis as in equation 5.30
we have
X
(6.27) B= Bab τ a ⊗ τ b .
1≤a,b≤m
While using the coefficients in equation 6.26 and the tensor product basis as
in equation 5.30 we have
!
X X
(6.28) T t (B) = Aai Abj Bab αi ⊗ αj .
1≤i,j≤n 1≤a,b≤m
By using the formula for T t (τ a ), T t (τ b ) from equation 6.6, this last formula
can also be written as
!
X
T t (B) = T t Bab τ a ⊗ τ b
(6.29) 1≤a,b≤m
X
= Bab T t (τ a ) ⊗ T t (τ b ) .
1≤a,b≤m
Therefore writing
X
B= Bab dy a |q ⊗ dy b |q , Bab ∈ IR ,
1≤a,b≤m
6.31 we find
Therefore
Φ∗ (dx2 + dy 2 + dz 2 ) = dx2 + dy 2 + (fx dx + fy dy)2
= (1 + fx2 )dx2 + 2fx fy dxdy + (1 + fy2 )dy 2 .
Example 6.2.5. Let U ⊂ IR 3 and γ be the metric tensor in 5.4.3, and let
Φ : IR 3 → U be
Φ(u, v, w) = (x = eu , y = veu , z = ew ) .
Proof. We only need to prove that at each point p ∈ IR m that (Φ∗ γ)p is an
inner product. XXXX
6.3 Isometries
Let γ be a fixed metric tensor on IR n .
6.3. ISOMETRIES 123
(6.38) Φ∗ γ = γ.
Let’s write this out more carefully using the definition of pullback in
equation 6.33. We have Φ is an isometry of γ if and only if
1. Φ is an isometry.
n
X n
X
i i
γ(Φ∗,p Xp , Φ∗,p Xp ) = γ(Φ∗,p X ∂x , Φ∗,p X j ∂xj )
i=1 j=1
n
X
= X i X j γ(Φ∗,p ∂xi |p , Φ∗,p ∂xj |p )
i,j=1
X n
= X i X j γ(∂xi |p , ∂xj |p ) by1.
i,j=1
= γ(X, X).
Therefore 2 implies 3.
For 3 to imply 1), let Xp , Yp ∈ Tp IR n , then by hypothesis
Expanding this equation using bilinearity the left hand side of this is
Again using the hypothesis 3, γ(Φ∗,p Xp , Φ∗,p Xp ) = γ(Xp , Xp ), and γ(Φ∗,p Y, Φ∗,p Y ) =
γ(Y, Y ) in equation 6.44 we are left with
The two by two matrix is the Jacobian matrix for Φ at the point p (in this
case the point p doesn’t show up in evaluating the Jacobian). We see the
the coefficients of the image vector, are just the rotated form of the ones we
started with in Xp .
XXXXXXXXXXXXX DRAW PICTURE XXXXXXXXXXXXXX
6.3. ISOMETRIES 125
where t ∈ IR . We compute
Ψ∗t γ E = dx2 + dy 2 .
126 CHAPTER 6. THE PULLBACK AND ISOMETRIES
This will not equal γ unless t = 0. In which case Ψt is just the identity
transformation.
Given a metric tensor γ the the isometries have a simple algebraic struc-
ture.
Theorem 6.3.8. Let γ be a metric tensor in IR n . The set of isometries of
γ form a group with composition of functions as the group operations. This
group is called the isometry group of the metric.
Proof. Let Φ and Ψ be isometries of the metric γ and Xp , Yp ∈ Tp IR n , then
Lastly
Pn wei write out in coordinates the isometry condition. Suppose that
γ = i,j=1 g j(x)dx dx is a metric tensor in IR n and that Φ : IR n → IR n
i j
Now using equation 6.54 but replacing i with m and then differentiating with
respect to xi we get equation 6.55 with i and m switched,
n
∂ 2 Φk ∂Φl ∂Φk ∂ 2 Φl
X
(6.56) m ∂xi ∂xj
+ m j i δkl = 0.
k,l=1
∂x ∂x ∂x ∂x
Φ(x) = Ax + b
The method we used to solve these equations is to take the original system
of equations and differentiate them to make a larger system for which all
possible second order partial derivatives are prescribed. This holds in general
for the isometry equations for a diffeomorphism and the equations are what
is known as a system of partial differential equations of finite type.
There another way to find isometries without appealing to the equations
(6.53) for the isometries. This involves finding what are known as “Killing
vectors” and their corresponding flows, which we discuss in the next two
chapters. The equations for a Killing vector are linear and often easier to
solve than the non-linear equations for the isometries. “Killing vectors” are
named after the mathematician Wilhelm Killing (1847-1923).
6.4. EXERCISES 131
6.4 Exercises
1. If T : V → W is a linear transformation, show that T t : W ∗ → V ∗ is
also a linear transformation.
2. Let Φ : IR 3 → IR 2 be
Φ(x, y, z) = (u = x + y + z, v = xy + xz)
and compute
(a) Φt 2du|(4,3) − dv|(4,3) (1,2,1)
, and
(b) Φ∗ (vdu + dv).
and compute
ψt (x, y) → (et x, et y)
are isometries.
132 CHAPTER 6. THE PULLBACK AND ISOMETRIES
Φ(a,b,c) (x, y, z) = (x + a, y + az + b, z + c)
7. For which a, b, c ∈ IR is
Φ(a,b,c) (x, y, z) = (x + a, y + cx + b, z + c)
(6.59) Φ(x) = Ax + b,
Hypersurfaces
S = { p ∈ IR n+1 | F (p) = c}
133
134 CHAPTER 7. HYPERSURFACES
and note that at any point p ∈ Srn not all x1 , . . . , xn+1 can be zero at the
same time (because r2 > 0).
Let S ⊂ IR n+1 be a regular level surface F = c. The tangent space
Tp S, p ∈ S is
Tp S = { Xp ∈ Tp IR n+1 | Xx ∈ ker F∗ |p }.
Note that since F∗ has rank 1, that dim Tp S = n by the dimension theorem
2.2.7.
Lemma 7.1.2. If Xp ∈ Tp IR n+1 then Xp ∈ Tp S if and only if
Xp (F ) = 0
Proof. Let ι ∈ C ∞ (IR ) be the identity function. We compute F∗ Xp (ι) =
Xp (ι ◦ F ) = Xp (F ). This vanishes if and only if Xp (F ) = 0.
Example 7.1.3. Let r > 0 and Sr2 = { (x, y, z) ∈ IR 3 | x2 + y2 + z 2 = r2 },
which is the 2-sphere in IR 3 of radius r. Let p = √16 , √16 , √13 ∈ S 2 (where
r = 1), and let
Xp = ∂x − ∂y
then
Xp (x2 + y 2 + z 2 ) = (2x − 2y)|p = 0.
Therefore Xp ∈ Tp S 2 .
Let’s compute the tangent space Tp Sr2 by finding a basis. In the coordinate
basis we have by equation 7.1,
[F∗ ] = (2x, 2y, 2z)
where x, y, z satisfy x2 + y 2 + z 2 = r2 . In order to compute the kernel note
the following, if z 6= ±r, then
ker[F∗ ] = span{(−y, x, 0), (0, −z, y)}.
Rewriting this in the standard basis we have
Tp Sc2 = span{−y∂x + x∂y , −z∂y + y∂z } p 6= (0, 0, ±r).
At the point p = (0, 0, ±r) we have
Tp Sc2 = span{∂x , ∂y } p = (0, 0, ±r).
7.2. PATCHES AND COVERS 135
F ◦ σ = c.
(F ◦ σ)∗ ∂t = F∗ σ∗ ∂t = c∗ ∂t = 0.
1. U ⊂ IR n is open,
2. ψ : U → IR n+1 is a smooth injective function,
3. ψ(U ) ⊂ S,
4. and ψ∗ is injective at every point in U (so ψ is an immersion)
The conditions for a patch are easily checked. Every point (x0 , y0 , z0 ) ∈ S is
contained in the given patch.
Suppose (U, ψ) is a patch on a regular level surface S. Then ψ : U → S is
a one-to-one immersion. The differential ψ∗ is injective by definition and so
ψ∗ (Tx U ) ⊂ Tψ(x) IR n+1 is an n-dimensional subspace. However we find even
more is true.
Lemma 7.2.4. The map ψ∗ : Tp U → Tq S, where q = ψ(p) is an isomor-
phism.
Proof. We have ψ∗ (Tp U ) and Tq S are n-dimensional subspaces of Tq IR n+1 .
If ψ∗ (Tp U ) ⊂ Tp S then they are isomorphic. Let Xp ∈ Tp U , we only need to
check that ψ∗ Xp ∈ ker F∗ . We compute
F∗ ψ∗ Xp = Xp (F ◦ ψ).
ψ∗ ∂u |(u,v) = (− sin u sin v∂x + cos u sin v∂y )(cos u sin v,sin u sin v,cos v) .
is a surface patch on Sr2 (the upper hemi-sphere). Likewise the pair (D, ψz− )
is a surface patch on Sr2 (the bottom hemi-sphere) where
√
ψz− (u, v) = (x = u, y = v, z = − r2 − u2 − v 2 ).
Continuing in the way we construct four more patches all using D and the
functions √
ψx± (u, v) = (x = ± r2 − u2 − v 2 , y = u, z = v)
√
ψy± (u, v) = (x = u, y = ± r2 − u2 − v 2 , z = v).
The collection
C = {(D, ψz± ), (D, ψx± ), (D, ψy± ) }
is a cover of Sr2 by coordinate patches.
The fact is regular level surfaces always admit a cover. This follows from
the next theorem which we won’t prove.
138 CHAPTER 7. HYPERSURFACES
We can view the disk u2 + v 2 < 1 as lying in the xy-plane and the image
under ψ as the upper part of S 2 . Let σ(t) = (x(t), y(t), z(t)) be a curve
on the upper half of the sphere. The curve σ is the image of the curve
τ (t) = (u(t), v(t)) ⊂ U which is the projection of σ into the xy-plane. In
particular from equation 7.3 we have u(t) = x(t), v(t) = y(t) and therefore
have
p
(7.4) σ(t) = ψ ◦ τ (t) = (u(t), v(t), 1 − u(t)2 − v(t)2 )
Given the curve (7.4) on the surface of the sphere we can compute its arc-
length as a curve in IR 3 using the Euclidean metric γEu (see equation ??),
s
Z b 2 2 2
dx dy dz
LEu (σ) = + + dt
a dt dt dt
v !2
Z bu
u du 2 dv 2 1 du dv
= t + + p (u + v ) dt
a dt dt 1 − u(t)2 − v(t)2 dt dt
Z br
1
= ((1 − v 2 )(u̇)2 + 2uv u̇v̇ + (1 − u2 )(v̇)2 ) dt
a 1 − u2 − v 2
Note that this is the same arc-length we would have computed for the curve
τ (t) = (u(t), v(t)) using the metric tensor
1 2 2 2
(7.5) γ̂U = (1 − v )du + 2uvdudv + (1 − u )dv
1 − u2 − v 2
defined on the set U !
Let’s look at this problem in general and see where the metric tensor 7.5
comes from.
Suppose S ⊂ IR n+1 is a regular level surface. Let γ be a metric tensor
on IR n+1 . The metric γ induces a metric tensor on S, we denote by γS as
follows. Let q ∈ S, and X, Y ∈ Tq S, then
(7.8) γ̂U = ψ ∗ γ.
Therefore
ψ ∗ γEu = (ψ ∗ dx)2 + (ψ ∗ dy)2 + (ψ ∗ dz)2 + (ψ ∗ dw)2
2t2 + 1 2
= 2 dt + (t2 + 1)(sin2 v du2 + dv 2 ).
t +1
Remark 7.5.4. If γ is not positive definite then the signature of γS will depend
on S.
(7.9) φ−1 ◦ ψ : U0 → V0 ,
where φ−1 : W → V0 . While the function φ−1 ◦ ψ exists, its explicit determi-
nation is not always easy.
The functions ψ and φ provide two different coordinate systems for the
points of S which lie in W , and 7.9 are the change of coordinate functions
for the points in W .
7.5. METRIC TENSORS ON SURFACES 143
Example 7.5.5. Let (U, ψ) be the chart on S 2 from equation (7.3) (upper
half of sphere), and let (V, φ) be the chart
√
φ(s, t) = (x = s, y = 1 − s2 − t2 , z = t)
where V is the interior of the unit disk V = {(s, t) | s2 + t2 < 1 }. The set
W is then set
Drawpicture,
and
ψ(θ, φ) = (cos θ sin φ, sin θ sin φ, cos φ), 0 < θ < 2π, 0 < φ < π,
The overlap on S 2 consists of points in the top half of S 2 minus those with
y = 0, x ≥ 0, and
π
U0 = {(θ, φ) | 0 < θ < 2π, 0 < φ < , V0 = {(u, v) | u2 +v 2 < 1}−{(u, v) | v = 0, 0 < u < 1}.
2
The function λ : U0 → V0 is again easily determined using the projection
map π(x, y, z) = (x, y) to give
We now find the relationship between γ̂U and γ̂V . Assume that the change
of coordinate functions λ : U0 → V0
(7.10) λ = φ−1 ◦ ψ.
144 CHAPTER 7. HYPERSURFACES
ψ = φ ◦ λ.
We claim that
λ∗ du = λ∗ du = − sin θ sin φdθ+cos θ cos φdφ, , λ∗ dv = cos θ sin φdθ+sin θ cos φdφ,
1 − sin2 θ sin2 φ
λ∗ γ̂V0 = 2
(− sin θ sin φdθ + cos θ cos φdφ)2
cos φ
2 cos θ sin θ sin2 φ
+ (− sin θ sin φdθ + cos θ cos φdφ) (cos θ sin φdθ + sin θ cos φdφ)
cos2 φ
1 − cos2 θ sin2 φ
+ 2
(cos θ sin φdθ + sin θ cos φdφ)2 ,
cos φ
which expands to (noting that the term with dθdφ is zero after expanding),
(1 − sin2 θ sin2 φ) sin2 θ sin2 φ − 2 cos2 θ sin2 θ sin4 φ + (1 − cos2 θ sin2 φ) cos2 θ sin2 φ 2
dθ
cos2 φ
+ (1 − sin2 θ sin2 φ) cos2 θφ + 2 cos2 θ sin2 θ sin2 φ + (1 − cos2 θ sin2 φ) sin2 θ dφ2
finally giving
λ∗ γ̂V0 = sin2 φdθ2 + dφ2 = γ̂U0 .
7.5. METRIC TENSORS ON SURFACES 145
Example 7.5.8. Let’s check formula (7.11) using the two patches on S 2
from example (7.5.5) above. Here we have
1
(1 − v 2 )du2 + 2uvdudv + (1 − u2 )dv 2
γ̂U = 2 2
1−u −v
1 2 2 2 2
γV = (1 − t )ds + 2stdsdt + (1 − s )dt
1 − s2 − t2
√
using λ(u, v) = (u, 1 − u2 − v 2 ), then
u v
λ∗ ds = du, λ∗ dt = − √ du − √ dv.
2
1−u −v 2 1 − u2 − v 2
Therefore
1 − u2
∗ 1 2 2 2 2
λ γ̂V = 2 (u + v )du − 2udu(udu + vdv) + (udu + vdv)
v 1 − u2 − v 2
(1 − u2 )2
1 2 2 2
= 2 (u + v ) − 2u + 2 2
du2
v 1−u −v
2
1 − u2
1 2uv(1 − u )
+ 2 − 2uv dudv + dv 2
v 1 − u2 − v 2 1 − u2 − v 2
1 2 2 2 2
= (1 − v )du + 2uvdudv + (1 − u )dv
1 − u2 − v 2
146 CHAPTER 7. HYPERSURFACES
7.6 Exercises
1. Find a basis for the tangent spaces (as subspaces of the tangent space
to the ambient IR n+1 ) to the following hypersurfaces at the indicated
points in two different ways. First computing the kernel of F∗ , and
then by constructing a sufficient number of curves on the level surface
which pass through the given point. Check the answers are the same.
(7.12) S = { (x, y, z) ∈ IR 3 | x2 + y 2 − z 2 = 1}
is a smooth patch on S.
(d) Find p = (u0 , v0 ) ∈ U and Yp ∈ Tp U such that ψ(p) = (1, 1, 1) ∈ S
and ψ∗ Yp = (∂x − ∂y )q .
(e) Find a cover of S by patches of which one is ψ from equation 7.13.
3. Find a cover of Sr2 containing two patches, where one patch is in ex-
ample 7.2.2.
Σ = { (u, v, w) | w2 − u2 − v 2 = −16},
and let Φ : IR 3 → IR 3 be
5. Let ψ1 , ψ2 : IR 2 → IR 3 , where
u2 + v 2 − 1
1 2u 2v
ψ (u, v) = , ,
1 + u2 + v 2 1 + u2 + v 2 1 + u2 + v 2
1 − u2 − v 2
2 2u 2v
ψ (u, v) = , , .
1 + u2 + v 2 1 + u2 + v 2 1 + u2 + v 2
6. Let L± be the line in IR 3 with end point (0, 0, ±1) and passing through
the point (u, v, 0) ∈ IR 3 . Compute the (x, y, z) coordinates of the point
on S 2 where L± intersects S 2 in terms of u and v. Relate this to the
previous problem.
(a) Compute the coordinate form of the metric γS 2 on the two stere-
ographic charts ψ 1,2 from question 5.
(b) Compute the change of coordinate functions and identify the do-
main on S 2 where this is valid. (Draw a picture).
(c) Check the overlap formula (7.11) and identify the domain where
this is valid.
Chapter 8
8.1 Flows
In equation (6.47) of example 6.3.5 the family of diffeomorphisms Ψt : IR 2 →
IR 2 are given
Ψt (x, y) = (x + t, y)
which are parameterized by t ∈ IR . We can rewrite the function Ψt as
Ψ : IR × IR 2 → IR 2 by
Another term often used for a flow is a one parameter group of transfor-
mations. The reason for this terminology is given in chapter 10.
149
150CHAPTER 8. FLOWS, INVARIANTS AND THE STRAIGHTENING LEMMA
These are smooth functions (since Ψ is), and so X is a smooth vector field.
The vector field X is called the infinitesimal generator of the flow Ψ.
8.1. FLOWS 151
Example 8.1.3. Let Ψ be the flow equation (6.47). By equation 8.3 the
corresponding vector field is just
X = ∂x .
Example 8.1.4. Let Ψ be the flow from equation 8.1, then the corresponding
vector field X = X 1 ∂x + X 2 ∂y are computed by equation 8.3 to be
1 ∂
X = (x cos t − y sin t) = −y
∂t t=0
2 ∂
X = (x sin t + y cos t) = x.
∂t t=0
X = −y∂x + x∂y .
X = y∂x + z∂y .
Proof. Let p1 = ψp (t1 ) be a point along the curve, and let s = t − t1 . Define
the curve σ(s) = ψp (s + t1 ), and note
dσ(s) dψp (t)
(8.5) = .
ds s=0 dt t=t1
152CHAPTER 8. FLOWS, INVARIANTS AND THE STRAIGHTENING LEMMA
X = y∂x + z∂y .
Ψ(t, x0 ) = x(t)
8.2 Invariants
We begin this section with some ideas that we will study in more detail in
the rest of the book. Let Φ : IR n → IR n be a diffeomorphism. A subset
S ⊂ IR n is a Φ invariant subset if Φ(S) ⊂ S, and a function f ∈ C ∞ (IR n )
is a Φ invariant function if Φ∗ f = f . Given these definitions we now define
what a flow invariant set and a flow invariant function are.
Let Ψ : IR × IR n → IR n be a flow. If we fix t = t0 in the function Ψ, let
Ψt0 : IR n → IR n be the smooth function
Ψ−t0 ◦Ψt0 (x) = Ψ(−t0 , Ψ(t0 , x)) = Ψ(−t0 +t0 , x) = Ψ(0, x) = x = Ψt0 ◦Ψ−t0 (x).
S = { (x, y) ∈ IR 2 | a ≤ x2 + y 2 ≤ b}
156CHAPTER 8. FLOWS, INVARIANTS AND THE STRAIGHTENING LEMMA
The sum of the squares of these entries is x20 + y02 , and so Ψ(t, p) ∈ S.
2
Example 8.2.3. Let Ψ(t, (x, y, x)) = (x + ty + t2 z, y + tz, z) be the flow from
example 8.1.5. The set
S = { (x, y, 0) ∈ IR 3 }
Example 8.2.6. For the flow Ψ in example 8.1.2, the function f (x, y) =
x2 + y 2 is invariant. We check this by equation 8.11,
f (Ψ(t, (x, y)) = (x cos t − y sin t)2 + (x sin t + y cos t)2 = x2 + y 2 = f (x, y).
Example 8.2.7. For the flow in example 8.1.5 above, the function f (x, y, z) =
z is an invariant. Let S = {(x, y, z) ∈ IR 3 | z 6= 0}, which is a Ψ invariant
2
set. The function g(x, y, z) = x − y2z defined on S is Ψ invariant. We check
that g is an invariant using equation 8.11 at points in S,
t2 (y + tz)2 y2
g(Ψ(t, (x, y, z))) = x + ty + z− =x− = g(x, y, z)
2 z 2z
8.2. INVARIANTS 157
(8.12) X(f ) = 0
Example 8.2.10. For X = y∂x + z∂y the function f (x, y, z) = z, and the
2
function g(x, y, x) = x − y2z defined on z 6= 0 are X-invariant.
X = px ∂x + py ∂y − x∂px − y∂py .
This is the Hamiltonian vector field for the Hamiltonian H = p2x +p2y +x2 +y 2 .
The functions
f1 = ypx − xpy , p2x + x2 , p2y + y 2
are all X-invariants. The first invariant is the angular momentum.
The invariants in examples 8.1.2 and 8.2.9, and example 8.2.7 and 8.2.10
suggest a relationship between invariants of a flow and invariants of their
corresponding infinitesimal generator.
Expanding this equation and using equation 8.3 the left hand side becomes
n
∂Ψi
X ∂f
= X(f ) = 0
i=1
∂t t=0 ∂xi Ψ(0,x)
158CHAPTER 8. FLOWS, INVARIANTS AND THE STRAIGHTENING LEMMA
LX f = 0
8.3 Invariants I
In Theorem 8.2.12 it was shown that finding invariants of a flow Ψ and
its infinitesimal generator X are the same problems. We will describe two
different ways to find these invariants, the first is based on the equation 8.12
X(f ) = 0, and the second is based on the equation 8.11 f (Ψ(t, x)) = f (x).
Suppose that
Xn
X= X i (x)∂xi
i=1
8.3. INVARIANTS I 159
The partial differential equation (8.14) is linear, and so the constants are
solutions. These turn out to not be useful solutions.
Therefore the first method to finding invariants is to solve the partial
differential equation (8.14). This theory is well developed, and in this section
will just demonstrate it through a few examples. The method developed in
the next section for finding invariants of flows solves (in a sense defined below)
the partial differential equations 8.14.
To see how this works, suppose that X = a(x, y)∂x + b(x, y)∂y , and that
we are searching for an invariant f (x, y). Consider the ordinary differential
equation
dy b(x, y)
(8.15) =
dx a(x, y)
which is only defined where a(x, y) 6= 0. If
(8.16) g(x, y) = c
is the general solution to equation 8.15, then g(x, y) is an invariant of X.
Let’s check this in an example.
Example 8.3.1. Let X = x∂x + y∂y . The ode (8.15) is
dy y
= .
dx x
The general solution is y = kx, k ∈ IR a constant. The function xy is easily
checked to be X invariant. This can easily be generalized for special vector
fields in higher dimension. For example
X = x∂x + y∂y + z∂z ,
the functions xy −1 and xz −1 are invariant. This works because of missing
variables in the coefficients of X. If
X = (x + y + z)∂x + (x + y)∂y + (x + z)∂z
then this would not work.
160CHAPTER 8. FLOWS, INVARIANTS AND THE STRAIGHTENING LEMMA
0 0 1
and so the two functions are functionally independent everywhere they are
defined. By Theorem 8.3.2 the general solution to the partial differential
equation
∂w ∂w
y +z =0
∂x ∂y
2
is w = F z, x − y2z where F : IR 2 → IR is arbitrary.
Y 1 = Y (y 1 ) = X(y 1 ) = X(f 1 ) = 0,
Y 2 = X(y 2 ) = X(f 2 ) = 0,
..
.,
n−1
Y = X(y n−1 ) = X(f n−1 ) = 0.
X = y∂x + z∂y .
2
The two invariants are u = x − y2z , v = z, and these are functionally indepen-
dent away from (y, z) = (0, 0). Take for the third function w = y, so that
u, v, w are a functionally independent set. We now find the vector field in X
these coordinates. All we need is
X(w) = z.
X = v∂w ,
y
as expected from Lemma 8.3.4. Note if we take as our third coordinate w = z
we get
X(w) = 1
and
X = ∂w .
8.4 Invariants II
We now look at a method for finding the invariants of a flow Ψ : IR × IR n →
IR n .
We begin with a global description of the solution. Let ι : IR n−1 → IR n
be a smooth function and define the function Φ : IR × IR n−1 → IR n , by
Proof. Suppose that p = ι(u0 ) and that ψp (t1 ) = ι(u1 ). This implies
Proof. To prove this we just write out the definitions using equation 8.17,
This lemma will be critical for what we have to say next. Consider Φ−1 :
IR n → IR × IR n−1 so that we can write the components
By taking the individual components in this equation, and using the fact
that p was arbitrary proves the lemma.
Therefore the function ρ(p) gives the t-value of the flow required to get from
the point u on the cross-section K = ι(IR n−1 ) to the point p. It is also worth
noting then that −ρ(p) is the t-value for p that flows p back to u ∈ K. This
interpretation helps in determining what to chose for K, and also what Φ−1
is. In other words, we need to choose K so that we can flow back to K from
any point in IR n , this determines ρ which in turn determines Φ−1 and hence
the invariants.
y y2
(8.21) Ψ(− , (x, y, z)) = (x − , 0, z)
z 2z
The values of our invariants (u, v) are read off equation 8.21 to be,
y 1 y2
u=x− y+ z, v = z.
z 2 z2
Example 8.4.8. The rotation in the plane flow is
d
Xp (ρ) = ρ(ψp (t))|t=0
dt
which simplifies on account of (8.20) to give
Xp (y n ) = X(ρ)p = 1.
8.5 Exercises
1. Check that each of the following mappings define a flow.
2. Plot the curves Ψp (t) for a few initial points p in the flows in question
1.
(a) X = ∂x − ∂y + 2∂z .
(b) X = x∂x − y∂y + 2z∂z .
(c) X = x∂y + y∂z .
(d) X = x∂y + y∂x .
(e) X = ex ∂x ( on R).
(f) X = (2x − y)∂x + (3x − 2y)∂y
5. For each of the global flows in problem 4, find an open set together with
a complete set of invariants on this set. Check that your invariants are
indeed constant along the flow (constants of motion).
6. For each global flow in problem 4, straighten out the vectors fields on
an open set about a point of your choosing.
7. Find a complete set of invariants for each of the following vector fields
(Remember that you do not necessarily need to find the flow). Identify
an open set where the invariants are functionally independent.
8.5. EXERCISES 169
(a) X = y 3 ∂x − x3 ∂y .
(b) X = yz∂x + xz∂y + xy∂z .
(c) X = (x2 + y 2 )∂x + 2xy∂y
(d) X = (y − z)∂x + (z − x)∂y + (x − y)∂z ,
(e) X = x2 ∂x + y 2 ∂y + (x + y)z∂z
9. Show that if f and g are invariants for a flow Ψ, then h(x) = F (f (x), g(x))
is an invariant, where F is any smooth function of two variables. Re-
peat this problem if f, g are invariants of a vector field X.
10. Let X = ni=1 nj=1 Aij xj ∂i be a linear vector field on Rn . Find matrix
P P
equations
Pn which are necessary and suffice for aPlinear function f (x) =
i n Pn i j
i=1 bi x and a quadratic function g(x) = i=1 j=1 bij x x (here
bij = bji ) to be invariants of X.
171
172 CHAPTER 9. THE LIE BRACKET AND KILLING VECTORS
X = ∂x + ∂y , Y = x∂x + y∂y , Z = x2 ∂x + y 2 ∂y
(9.7) LX Y = [X, Y ].
Proof. In order to check this formula we fix a point p ∈ IR n and show equa-
tion 9.7 holds at p. Equation (9.6) written in components in the coordinate
basis {∂xi }1≤i≤n ,
(9.8) !
n i
1 X ∂Ψ (−t, p)
(LX Y )ip = lim j
Y j (Ψ(t, p)) − Y i (Ψ(t, p)) + Y i (Ψ(t, p)) − Y i (p) .
t→0 t ∂x
j=1
where we have switched the order of differentiation in the term with the
second derivatives. Comparing this with the components of [X, Y ] in the
coordinate basis in equation 9.3 proves [X, Y ]ip = (LX Y )ip . Since p is arbitrary
the theorem is proved.
Let Φ : IR n → IR n be a diffeomorphism. A vector field Y on IR n is
Φ-invariant if
Φ∗ Y = Y.
A vector-field Y is said to be invariant under the flow Ψ if for all diffeomor-
phisms Ψt , t ∈ IR ,
(9.10) (Ψt )∗ Y = Y .
LX Y = 0.
if and only if
[X, Y ] = 0.
176 CHAPTER 9. THE LIE BRACKET AND KILLING VECTORS
Another way to write equation 9.11 is that for all t, s ∈ IR the diffeomor-
phisms commute
Ψt ◦ Φs = Φs ◦ Ψt .
Example 9.1.11. Let X = x∂x + y∂y , and Y = −y∂x + x∂y . We’ll show
that [X, Y ] = 0, the flows commute, and Y is invariant under the flow of X.
First using equation 9.3 we have
We easily compute the two compositions in 9.11 using the flows in 9.12
Φ(s, Ψ(t, (x, y)) = (xet cos s − yet sin s, xet sin s + yey cos s)
= Ψ(t, Ψ(s, (x, y)).
Finally let’s check that Y is invariant under the flow of Ψ. We find the
components from push-forward formula in equation ?? to be
n
∂Ψi (t, x)
i
X
j −yet
((Ψt )∗ Y ) = Y (x) = = Y i (Ψ(t, x),
∂xj xet
j=1
and so Y is Ψ invariant.
Lemma 9.1.12. Suppose that the vector fields X and Y are linearly inde-
pendent at the point p and that [X, Y ] = 0. Then there exists an open set
U ⊂ IR n with p ∈ U , and coordinates (y 1 , . . . , y n−2 , r, s) on U such that
X = ∂s , Y = ∂r .
9.1. LIE BRACKET 177
Proof. We show two proofs here. The first is based on the Straightening
Lemma 8.4.9. Let (ui , v)1≤i≤n−1 be the straightening coordinates for X (on
an open set V with p ∈ V ), so that
(9.13) X = ∂v .
which does not vanish at p, by the linearly independence condition. Now use
the Straightening Lemma 8.4.9 on Z, to find an open set U with p ∈ U , and
coordinates on U , (y i = f i (u), r = f (u), v = v)1≤i≤n−2 , such that
Z = ∂r .
With the coordinates (y i , r, v) the form of the vector field X in 9.13 does not
change, and from equation 9.14
Y = Y 0 (r, y i )∂v + ∂r .
Proof. The second proof is similar to the sequence of steps used in proof 2
of the Straightening Lemma 8.4.9. Suppose Ψ is the flow for X and Φ the
flow for Y , and define H : IR 2 × IR n → IR n by
Expanding out the left hand side of this in components using equation 6.35
to compute Ψ∗t γΨ(t,x) we have
n
!
X ∂Ψi (t, x) ∂Ψj (t, x)
Ψ∗t γΨ(t,x)
kl
= gij (Ψ(t, x)) .
i,j=1
∂xk ∂xl
180 CHAPTER 9. THE LIE BRACKET AND KILLING VECTORS
n n
X ∂gij ∂Ψm ∂Ψi ∂Ψj X ∂gij m i j ∂gkl (x) m
(9.20) = X δ k δ l = X (x)
i,j,m=1
∂xm ∂t ∂xk ∂xl i,j,m=1
∂xm ∂xm
where we have substitute from equations 8.3, and 9.9. Next we evaluate at
t = 0 in the second derivative term to get,
where again we have used equation 8.3 for the infinitesimal generator. Finally
substituting equations 9.20 and 9.21 into equation 9.19 gives 9.17.
2 2 2 ∂X 1
(9.23) − X + =0
y3 y 2 ∂x
2 2 2 ∂X 2
(9.24) − X + = 0,
y3 y 2 ∂y
∂X 2 ∂X 1
(9.25) + = 0.
∂x ∂y
This example highlights one aspect of the Killing equations 9.17. Given a
2
metric tensor γ on IR n , equation 9.17 can be considered a system of n (n+1)2
partial differential equations for the n unknown functions X i (x) of n variables
xi . These partial differential equations are linear in the unknowns X i , and
are very over determined (more equations than unknown functions). Solving
these determines all the infinitesimal generators whose flow is by isometries.
Example 9.2.3. Continuing from example 9.2.2 above, we found the par-
tial differential equations for a Killing vector X = X 1 ∂x + X 2 ∂y to be
182 CHAPTER 9. THE LIE BRACKET AND KILLING VECTORS
(9.23,9.25,9.24),
2 2 2 ∂X 1
− X + = 0,
y3 y 2 ∂x
2 2 ∂X 2
(9.26) − 3 X2 + 2 = 0,
y y ∂y
∂X 2 ∂X 1
+ =0.
∂x ∂y
We first integrate in the second equation in 9.26 to get
X2 = F (x)y,
Proof. We observe
1 ∂ ∗
LX γ = lim Ψ∗t γΨ(t,p) − γp =
Ψt γΨ(t,p) ,
t→0 t ∂t t=0
Theorem 9.2.6. Let X and Y be Killing vector fields for the metric γ, then
[X, Y ] is also a Killing vector field.
Switching the role of X and Y in this equation, and taking the difference
184 CHAPTER 9. THE LIE BRACKET AND KILLING VECTORS
gives
(9.29)
n
∂ 2 gkl m m
X
m j m j ∂gkl j ∂X j ∂Y
0= (X Y − Y X ) + m Y −X
j,m=1
∂xm ∂xj ∂x ∂xj ∂xj
m m 2 m 2 m
∂gkm j ∂X j ∂Y j ∂ X j ∂ Y
+ j j Y −X + gkm Y −X
Y ∂x ∂xl ∂xl ∂xl ∂xj ∂xl ∂xj
m m 2 m 2 m
∂gml j ∂X j ∂Y j ∂ X j ∂ Y
+ Y −X +gml Y −X .
∂xj ∂xk ∂xk ∂xk ∂xj ∂xk ∂xj
Now by the equality of mixed partial derivatives, the first term in equation
9.29 satisfies n
X ∂ 2 gkl
m j
(X m Y j − Y m X j ) = 0
j,m=1
∂x ∂x
and equation 9.29 is the Killing equation for the vector-field [X, Y ].
The equations LX γ = 0 are linear equations for the coefficients of X.
Therefore any linear combination of Killing vector fields is a Killing vector
field. Combining this with Theorem 9.2.6 above and we have the following.
Corollary 9.2.7. Let Γ be the vector space of Killing vectors, then Γ with
the vector-field bracket [ , ] is a Lie algebra.
Remark 9.2.8. An upper bound on the dimension of the Lie algebra of Killing
vectors is known to be [?]
n(n + 1)
dim Γ ≤
2
9.3. EXERCISES 185
9.3 Exercises
1. Which set of sets of vector-fields form a Lie algebra as a subalgebra of
all vector fields on the givenIR n .
(a) [X, Y ] = 0.
(b) The flows Ψ and Ψ are commuting flows.
186 CHAPTER 9. THE LIE BRACKET AND KILLING VECTORS
(c) X is Φ invariant.
(d) Y is Ψ invariant.
(e) Find coordinates which simultaneously straighten the vector fields
X and Y . (Lemma 9.1.12)
187
188CHAPTER 10. GROUP ACTIONS AND MULTI-PARAMETER GROUPS
(a, b) ∗ (a0 , b0 ) = (a + a0 , b + b0 ).
This group action is transitive. Why? Let (x1 , y1 ) and (x2 , y2 ) be two points
in IR 2 . We can find a group element (a, b) such that
and this is (a = x2 − x1 , b = y2 − y1 ).
and ι is smooth.
An action of the group G2 on IR 2 is
This action is not transitive. For example (0, 1) and (1, 0) are not in the
same orbit. It is not difficult to check that there are only two orbits in this
example.
Example 10.2.2. Consider the group and actions in example 13.5. We have
e = (1, 0), and let
e1 = ∂a , e2 = ∂b
be a basis for Te G. The two curves
satisfy
σ1 (1) = e,σ̇1 (1) = e1 ,
σ2 (0) = e,σ̇2 (0) = e2 .
Using these curves we can find the corresponding infinitesimal generators.
For the action of G3 on IR we find the infinitesimal generator for e1 is
d
X1 (x) = (xt)|t=1 = (x)x , X1 = x∂x .
dt
The infinitesimal generator corresponding to e2
d
X2 = (x + t)|t=0 = 1x , X2 = ∂x .
dt
For the action of G3 on IR 2 we find the infinitesimal generator corre-
sponding to e1 is
X1 = x∂x ,
10.2. INFINITESIMAL GENERATORS 193
Γ = span{x∂x , ∂x },
is a two dimensional vector space. To check this forms a Lie algebra we need
to only check that [X1 , X2 ] ∈ Γ. We find
[X1 , X2 ] = −∂x ∈ Γ.
µ((a, b), (t, 0)) = (at, b), µ((a, b), (1, t)) = (a, b + at).
d d
U(a,b) = µ((a, b), (t, 0)|t=1 = (ta, b)|t=1 = (a, 0), U = a∂a ,
dt dt
d d
V(a,b) = µ((a, b), σ2 (t))|t=0 = (a, b + ta)|t=0 = (0, 1), V = a∂b .
dt dt
Note that
[U, V ] = V
and so the left invariant vector fields g = span{X, Y } are a two dimensional
Lie algebra.
If σ i are curves which give rise to the right invariant vector-fields Xi , and
the left invariant vector-fields Yi then
n
X n
X
[Xi , Xj ] = Cijk Xk , [Yi , Yj ] = − Cijk Yk .
k=1 k=1
[Xi , Yj ] = 0!
Proof. We
= 1.
The right hand side of equation 10.5 is
n
X
γp (Xi (p), Xi (p)) = αpj (Xi (p))αap
j
(Xi (p)) = 1.
j=1
the identity if e = (1, 0, 1). The left invariant vector fields for the basis
{∂a , ∂b , ∂c of Te G, the left invariant vector-fields are computed to be
Using the one-forms in equation 10.8 the metric tensor in equation 10.4 is
2 2 2
1 1 y 1
γ= dx + (dy − dz) + dz
x x z z
(10.9) 2
1 2 2 y y 1
= 2 dx + dy − 2 2 dydz + + dz 2
x zx x2 z 2 z 2
10.5 Exercises
1. Suppose that each of the following transformations defines a group
action. Find the maximum domain for the parameters and group mul-
tiplication law ∗. Determine the identity in the group and give the
inverse of each element.
2. Let µ : G × IR 3 → IR 3 be given by
(a) Question 1 a.
200CHAPTER 10. GROUP ACTIONS AND MULTI-PARAMETER GROUPS
(b) Question 1 d.
(c) The group G = SL(2) acting on the upper half plan Im z > 0 by
az + b
z→ , where
cz + d
a b
SL(2) = { ∈ M2×2 (IR ) | ad − bc = 1}.
c d
4. Find the isotropy subgroup for each of the following actions at the given
point.
6. Let
ds2 = (u + v)−2 (du2 + dv 2 )
be a metric tensor on U = {(u, v) ⊂ IR 2 | u + v 6= 0}, and let a ∈
IR ∗ , b ∈ IR , and let Φ(a,b) : IR 2 → IR 2 be
ρ(A, x) = Ax A ∈ GL(n, IR ), x ∈ IR n .
is a group action.
(b) Let M = Mn×n (IR ) be the set of all n × n real matrices. Is the
function ρ : G × M → M given by
ρ(A, X) = AXA−1 , A ∈ G, X ∈ M
2
an action of GL(n, IR ) on M = Rn ?
Computations in Maple
203
Load the differential geometry and tensor packages.
> DGsetup([x,y,z],M):
Input the metric (named g here) from example 6.3.7 and check that the
transformations in 6.42 are isometries. The expression &t stands for the
tensor
> A:=matrix(3,3):
> for i from 1 to 3 do for j from 1 to 3 do
A[i,j]:=(ContractIndices(ContractIndices(g, Z||i,
[[1,1]]),Z||j,[[1,1]])):od:od:
> op(A);
1 0 0
0 1 0
0 0 1
Algebraic Notions
12.1 Groups
A binary operation on a set S is a function b : S × S → S. So it takes two
things in S (binary) and it’s output is some other element in S.
If ∗ satisfies
G4’) a ∗ b = b ∗ a
then G is an Abelian group.
Example 12.1.2. Let G = IR and take for the group operation ordinary
addition of real numbers. That is a ∗ b = a + b. Then IR with + is an Abelian
group. The identity element is the number 0.
207
Example 12.1.4. Let G = GL(n, IR ) denote the set of n × n invertible
matrices. For the binary operation ∗ we take ordinary matrix multiplication.
That is A∗B = AB. If n = 1 this is example 9.3, but when n > 1 this group is
not Abelian because matrix multiplication is not commutative. The identity
element is the identity matrix.
We can ask if this is a group on it’s own with the group operation being
matrix multiplication. However it is not obvious that matrix multiplication
is a binary operation on SL(n, IR ). That is, if A, B ∈ SL(n, IR ) is AB ∈
SL(n, IR )? This is easily checked
1
det A−1 = = 1.
det A
The simple test for a subgroup is the following.
2. if a ∈ H then a−1 ∈ H.
If in addition · satisfies,
R7’) a · b = b · a, f or alla, b ∈ R
then R is a commutative ring. If there exits 1R ∈ R satisfying
R8’) 1R · a = a · 1R = a
then R is a ring R with multiplicative identity.
and multiplication
(f · g)(x) = f (x)g(x),
where f, g ∈ C ∞ (U )
211
Tangent vector, 43
Tensor product, 82
Trace, 10
Unit n-sphere, 2
Vector space, 10
Bibliography
213
[12] W.R. Wade, An Introduction to Analysis, Third Edition, Prentice Hall,
New York 2003.
215
Tangent vector, 43
Tensor product, 82
Trace, 10
Unit n-sphere, 2
Vector space, 10