(Lahiri) Lecture Notes On Differential Geometry PDF
(Lahiri) Lecture Notes On Differential Geometry PDF
(Lahiri) Lecture Notes On Differential Geometry PDF
Chapter 1
Topology
1
2 Chapter 1. Topology
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 2
Manifolds
Now that we have the notions of open sets and continuity, we are
ready to define the fundamental object that will hold our attention
during this course.
• A manifold is a topological space which is locally like Rn . 2
That is, every point of a manifold has an open neighbourhood
with a one-to-one map onto some open set of Rn .
• More precisely, a topological space M is a smooth n-
dimensional manifold if the following are true:
i) We can cover the space with open sets Uα , i.e. every point of
M lies within some Uα .
ii) ∃ a map ϕα : Uα → Rn , where ϕα is one-to-one and onto
some open set of Rn . ϕα is continuous, ϕα −1 is continuous, i.e.
ϕα → Vα ∈ Rn is a homeomorphism for Vα .
(Uα , ϕα ) is called a chart (Uα is called the domain of the
chart). The collection of charts is called an atlas.
iii) In any intersection Uα ∩ Uβ , the maps ϕα ◦ ϕβ −1 , which are
called transition functions and take open sets of Rn to open
sets of Rn , i.e. ϕα ◦ ϕβ −1 : ϕβ (Uα ∩ Uβ ) → ϕα (Uα ∩ Uβ ), are
smooth maps. 2
• n is called the dimension of M. 2
We have defined smooth manifolds. A more general definition is
that of a C k manifold, in which the transition functions are C k , i.e. k
4
5
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
are real analytic, i.e. have a Taylor expansion at each point, which
converges. Smoothness of a manifold is useful because then we can
say unambiguously if a function on the manifold is smooth as we will
see below.
• A complex analytic manifold is defined similarly by replac-
ing Rn with Cn and assuming the transitions functions ϕα ◦ ϕβ −1 to
be holomorphic (complex analytic). 2
• Given a chart (Uα , ϕα ) for a neighbourhood of some point P, the
image (x1 , · · · , xn ) ∈ Rn of P is called the coordinates of P in the
chart (Uα , ϕα ). A chart is also called a local coordinate system.2
In this language, a manifold is a space on which a local coordi-
nate system can be defined, and coordinate transformations between
different local coordinate systems are smooth. Often we will suppress
U and write only ϕ for a chart around some point in a manifold. We
will always mean a smooth manifold when we mention a manifold.
Examples: Rn (with the usual topology) is a manifold. 2
The typical example of a manifold is the sphere. Consider the
sphere S n as a subset of Rn+1 :
1
The reason that it is not possible to cover the sphere with a single chart
is that the sphere is a compact space, and the image of a compact space un-
der a continuous map is compact. Since Rn is non-compact, there cannot be a
homeomorphism between S n and Rn .
6 Chapter 2. Manifolds
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
image of the point on the sphere which the line has passed through.
For points on the lower hemisphere, the line first passes through the
equatorial plane (image point) before reaching the sphere (source
point). Then using similarity of triangles we find (Exercise!) that
the coordinates on the equatorial plane Rn of the image of a point
on S n \{N } is given by
x2 xn+1
1 2 n+1
ϕN : x , x , · · · , x 7→ ,··· , . (2.2)
1 − x1 1 − x1
Similarly, the stereographic projection from the south pole is
ϕS : S n \{S} → Rn ,
x2 xn+1
1 2 n+1
x ,x ,··· ,x 7→ , · · · , . (2.3)
1 + x1 1 + x1
If we write
x2 xn+1
z= , · · · , , (2.4)
1 − x1 1 − x1
we find that
2 n+1 2
x2 1 − (x1 )2 1 + x1
2 x
|z| ≡ + · · · + = = (2.5)
1 − x1 1 − x1 (1 − x1 )2 1 − x1
The overlap between the two sets is the sphere without the poles.
Then the transition function between the two projections is
z
ϕS ◦ ϕN : Rn \{0} → Rn \{0}, z 7→ . (2.6)
|z|2
These are differentiable functions of z in Rn \{0}. This shows that
the sphere is an n-dimensional differentiable manifold. 2
• A Lie group is a group G which is also a smooth (real analytic
for the cases we will consider) manifold such that group composition
written as a map (x, y) 7→ xy −1 is smooth. 2
Another way of defining a Lie group is to start with an n-
parameter continuous group G which is a group that can be
parametrized by n (and only n) real continuous variables. n is called
not connected. 2
Rotations in three dimensions can be represented by 3 × 3 real
orthogonal matrices R satisfying RT R = I. Reflection is represented
by the matrix P = −I. The space of 3 × 3 real orthogonal matrices
is a connected manifold. 2
The space of all n× real non-singular matrices is called GL(n, R).
This is an n2 -dimensional Lie group and connected manifold. 2
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 3
Tangent vectors
f ◦ ϕβ −1 = (f ◦ ϕα −1 ) ◦ (ϕα ◦ ϕβ −1 ) (3.1)
9
10 Chapter 3. Tangent vectors
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
df
derivative. The rate of change of f along γ is written as .
dt
Suppose another curve another curve µ(s) meets γ(t) at some
point P , where s = s0 and t = t0 , such that
d d
(f ◦ γ) = (f ◦ µ) ∀f ∈ C ∞ (M) (3.2)
dt P ds P
That is, we are considering a situation where two curves are tangent
to each other in geometric and parametric sense. Let us introduce a
convenient notation. In any chart ϕ containing the point P, let us
write ϕ(P ) = (x1 , · · · , xn ). Let us write f ◦ γ = (f ◦ ϕ−1 ) ◦ (ϕ ◦ γ),
so that the maps are
f ◦ ϕ−1 : Rn → R, x 7→ f (x) or f (xi ) (3.3)
n i
ϕ◦γ : I →R , t 7→ {x (γ(t))}. (3.4)
The last are the coordinates of the curve in Rn .
Using the chain rule for differentiation, we find
d d ∂f dxi (γ(t))
(f ◦ γ) = f (x(γ(t))) = . (3.5)
dt dt ∂xi dt
Similarly, for the curve µ we find
d d ∂f dxi (µ(s))
(f ◦ µ) = f (x(µ(s))) = . (3.6)
ds ds ∂xi ds
Since f is arbitrary, we can say that two curves γ, µ have the same
tangent vector at the point P ∈ M (where t = t0 and s = s0 ) iff
dxi (γ(t)) dxi (µ(s))
= . (3.7)
dt
t=t0 ds
s=s0
We can say that these numbers completely determine the rate of
change of any function along the curve γ or µ at P. So we can define
the tangent to the curve.
• The tangent vector to a curve γ at a point P on it is defined
as the map
d
γ̇P : C ∞ (M) → R, f 7→ γ̇P (f ) ≡ (f ◦ γ)|P . (3.8)
dt
As we have already seen, in a chart with coordinates {xi } we can
dxi (γ(t))
The numbers are thus the components of γ̇P . We will
dt
ϕ(P )
often write a tangent vector at P as vP without referring to the curve
it is tangent to.
We note here that there is another description of tangent vectors
based on curves. Let us write γ ∼ µ if γ and µ are tangent to each
other at the point P . It is easy to see, using Eq. (3.7) for example,
that this relation ∼ is transitive, reflexive, and symmetric. In other
words, ∼ is an equivalence relation, for which the equivalence class
[γ] contains all curves tangent to γ (as well as to one another) at P .
• A tangent vector at P ∈ M is an equivalence class of curves
under the above equivalence relation. 2
The earlier definition is related to this by saying that if a vector
vP is tangent to some curve γ at P , i.e. if vP = γ̇P , we can write
vP = [γ].
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 4
Tangent Space
• The set of all tangent vectors (to all curves) at some point P ∈
M is the tangent space TP M at P. 2
Proposition: TP M is a vector space with the same dimension-
ality n as the manifold M.
Proof: We need to show that TP M is a vector space, i.e.
XP + YP ∈ TP M , (4.1)
aXP ∈ TP M , (4.2)
∀XP , YP ∈ TP M, a ∈ R.
That is, given curves γ, µ passing through P such that XP =
γ̇P , YP = µ̇P , we need a curve λ passing through P such that
λ̇P (f ) = XP (f ) + YP (f )∀f ∈ C ∞ (M).
Define λ : I → Rn in some chart ϕ around P by λ = ϕ ◦ γ + ϕ ◦
µ − ϕ(P ). Then λ is a curve in Rn , and
λ=ϕ◦λ:I →M (4.3)
is a curve with the desired property. 2
Note: we cannot define λ = γ + µ − P because addition does not
make sense on the right hand side.
The proof of the other part works similarly. (Exercise!)
To see that TP M has n basis vectors, we consider a chart ϕ with
coordinates xi . Then take n curves λk such that
ϕ ◦ λk (t) = x1 (P ), · · · , xk (P ) + t, · · · , xn (P ) ,
(4.4)
i.e., only the k-th coordinate varies along t. So λk is like the axis of
the k-th coordinate (but only in some open neighbourhood of P ).
13
14 Chapter 4. Tangent Space
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
∂
Now denote the tangent vector to λk at P by , i.e.,
∂xk P
∂ d
f = λ̇ (f ) = (f ◦ λ ) . (4.5)
k k
∂xk P dt
P
P
where vPi = vP (xi ). Thus the vectors ∂x∂ k P span TP M . These are
Chapter 5
Dual space
The dual space is a vector space under the operations of vector ad-
dition and scalar multiplication defined by
16
17
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
∂
So λ vanishes on all vectors, since the form a basis. Thus
∂xj P
the dxi span TP∗ M , so TP∗ M is n-dimensional.
18 Chapter 5. Dual space
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 6
Vector fields
dt
with initial condition xi (0) = xi |P . This is a set of ordinary first order
differential equations. If v i are smooth, the theory of differential
20
21
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
d i
x (γ(t)) = v i (γ(t)) ≡ v(xi (γ(t))) , (6.4)
dt
γ(0) = P . (6.5)
d i
x (γQ (t)) = v(xi γQ (t)) ,
(6.6)
dt
γQ (0) = Q . (6.7)
The set of all (real) vector fields V (M) on a manifold M has the
structure of a (real) vector space under vector addition defined by
Chapter 7
Two important concepts are those of pull back (or pull-back or pull-
back) and push forward (or push-forward or pushforward) of maps
between manifolds.
• Given manifolds M1 , M2 , M3 and maps f : M1 → M2 , g :
M2 → M3 , the pullback of g under f is the map f ∗ g : M1 → M3
defined by
f ∗g = g ◦ f . (7.1)
f ∗g = g ◦ f . (7.2)
While this looks utterly trivial at this point, this concept will become
increasingly useful later on.
• Given two manifolds M1 and M2 with a smooth map f : M1 →
M2 , P 7→ Q the pushforward of a vector v ∈ TP M1 is a vector
f∗ v ∈ TQ M2 defined by
24
25
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
f∗ (v1 + v2 ) = f∗ v1 + f∗ v2 (7.5)
f∗ (λv) = λf∗ v . (7.6)
(g ◦ f )∗ = g∗ f∗ , i.e.
(g ◦ f )∗ v = g∗ f∗ v ∀v ∈ TP M1 . (7.7)
f∗ v = f∗ [γ] = [f ◦ γ] (7.8)
vPµ = vP (y µ ) (7.10)
But
so
∂ µ ∂
f∗ (y ) = (y µ ◦ f ) . (7.13)
∂xi P ∂xi P
Since f∗ is linear, we can use this to find the components of (f∗ v)Q
for any vector vP ,
i ∂
f∗ vP = f∗ vP
∂xi P
i ∂
= vP f∗
∂xi P
µ
i ∂y (x) ∂
= vP (7.16)
∂xi P ∂y µ f (P )
µ
µ i ∂y (x)
⇒ (f∗ vP ) = vP . (7.17)
i
∂x P
Note that since f∗ is linear, we know that the components of f∗ v
should be linear combinations of the components of v , so we can
27
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
already guess that (f∗ vP )µ = Aµi vPi for some matrix Aµi . The matrix
is made of first derivatives because vectors are first derivatives.
Another example of the pushforward map is the following. Re-
member that tangent vectors are derivatives along curves. Suppose
vP ∈ TP M is the derivative along γ . Since γ : I → M is a map,
we can consider pushforwards under γ , of derivatives on I . Thus for
γ : I → M , t 7→ γ(t) = P , and for some g : M → R ,
d d
γ∗ g = (g ◦ γ)|t=0
dt t=0 dt
= γ̇P (g)|t=0 = vP (g) , (7.18)
so
d
γ∗ = vP (7.19)
dt t=0
• We can use this to give another definition of integral curves.
Suppose we have a vector field v on M . Then the integral curve of v
passing through P ∈ M is a curve γ : t 7→ γ(t) such that γ(0) = P
and
d
γ∗ = v|γ(t) (7.20)
dt t
for all t in some interval containing P . 2
Even though in order to define the pushforward of a vector v
under a map f , we do not need f to be invertible, the pushforward
of a vector field can be defined only if f is both one-to-one and onto.
If f is not one-to-one, different points P and P 0 may have the
same image, f (P ) = Q = f (P 0 ) . Then for the same vector field v we
must have
Chapter 8
Lie brackets
Thus
(uv − vu)(f g) = f (uv − vu)(g) + (uv − vu)(f )g , (8.4)
which means that the combination
[u , v] := uv − vu (8.5)
28
29
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
er = ex cos θ + ey sin θ
eθ = −ex sin θ + ey cos θ , (8.12)
∂ ∂
with ex = and ey = being the Cartesian coordinate basis
∂x ∂y
vectors, and
x y p
cos θ = , sin θ = , r = x2 + y 2 (8.13)
r r
Using these expressions, it is easy to show that [er , eθ ] 6= 0 , so
{er , eθ } do not form a coordinate basis.
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 9
Lie algebra
(x • y) • z + (y • z) • x + (z • x) • y = 0 . (9.1)
2
The Jacobi identity is not really an identity — it does not hold
for an arbitrary algebra — but it must be satisfied by an algebra for
it to be called a Lie algebra.
Example:
i) The space Mn = {all n × n matrices} under matrix multipli-
cation, A • B = AB . This is an associative algebra since
31
32 Chapter 9. Lie algebra
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
A • B = [A , B] = AB − BA . (9.2)
f • g = [f , g]P.B. . (9.5)
Chapter 10
Local flows
ϕ∗ f = f ◦ ϕ : M1 → R , (10.1)
and ϕ∗ f ∈ C ∞ (M1 ) if ϕ is C ∞ .
The pushforward of a vector vP is defined by
ϕ∗ vP (f ) = vP (f ◦ ϕ) = vP (ϕ∗ f ) (10.2)
vP ∈ TP M1 , ϕ∗ vP ∈ Tϕ(P ) M2 . (10.3)
33
34 Chapter 10. Local flows
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
ϕ∗ [u , v] = [ϕ∗ u , ϕ∗ v] . (10.6)
Proof:
ϕ∗ [u , v](f ) = [u , v] (f ◦ ϕ) ◦ ϕ−1
= u (v (f ◦ ϕ)) ◦ ϕ−1 − u ↔ v , (10.7)
while [ϕ∗ u , ϕ∗ v](f ) = ϕ∗ u (ϕ∗ v (f )) − u ↔ v
= u (ϕ∗ v (f ) ◦ ϕ) ◦ ϕ−1 − u ↔ v
= u v (f ◦ ϕ) ◦ ϕ−1 ◦ ϕ ◦ ϕ−1 − u ↔ v
2
• A vector field v is said to be invariant under a diffeomorphism
ϕ : M → M if ϕ∗ v = v , i.e. if ϕ∗ (vP ) = vϕ(P ) for all P ∈ M . 2
We can write for any f ∈ C ∞ (M)
∗
(ϕ∗ v) (f ) = ϕ−1 (v (ϕ∗ f ))
⇒ ϕ∗ ((ϕ∗ v) (f )) = v (ϕ∗ f ) ,
⇒ ϕ∗ ◦ ϕ∗ v = v ◦ ϕ∗ . (10.9)
ϕ∗ ◦ v = v ◦ ϕ∗ . (10.10)
35
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Proof:
∂ d ∗ ∂f
f (x, t) = (φt g) = v(f ) ≡ v i i , (10.17)
∂t dt ∂x
36 Chapter 10. Local flows
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 11
Lie derivative
= u (v(f )) − v (u(f ))
t=0 t=0
= [u , v](f ) . (11.5)
t=0
37
38 Chapter 11. Lie derivative
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
£u (f g) = (£u f ) g + f (£u g) ,
and £u (f + ag) = £u f + a£u g . (11.9)
£u (v + aw) = £u v + a£u w ,
and £u (f v) = (£u f ) v + f £u v ∀f ∈ C ∞ (M)
(11.10)
.
£u+av f = £u f + a£v = 0 ∀a ∈ R
and [£u , £v ]f = £[u ,v] f = 0 . (11.12)
φ−t v = v
⇒ £u v = v . (11.13)
Thus again the vector fields leaving w invariant form a Lie algebra.
• Let us also define the corresponding operations for 1-forms. As
we mentioned in Chap. 6, a 1−form is a section of the cotangent
bundle [
T ∗M = TP∗ M . (11.15)
P
Alternatively, a 1-form is a smooth linear map from the space of
vector fields on M to the space of smooth functions on M ,
£u ω(v) = £u ωi v i = uj j ωi v i
∂x
∂ωi ∂v i
= uj j v i + uj ωi j . (11.26)
∂x ∂x
41
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
We already know the left hand side of this equation from Eq. (11.26),
and the right hand side can be calculated in a chart as
We will not discuss this in detail, but only mention that it leads to
the same Leibniz rule as in Eq. (11.27), and the same description in
terms of components as in Eq. (11.29).
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 12
Tensors
∗ ∗
Ap,q
P
: TP M × · · · × TP M × TP M × · · · × TP M → R
| {z } | {z }
q times p times (12.5)
43
44 Chapter 12. Tensors
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
We can define the components of this tensor in the same way that
we did for the (1, 2) tensor. Then a (p, q) tensor has components
which can be written as
a ···a
Ab11···bqp .
• Some special types of (p, q) tensors have special names. A (1, 0)
tensor is a linear map AP : TP∗ M → R , so it is a tangent vector. A
(0, 1) tensor is a cotangent vector. A (p, 0) tensor has components
with p upper indices. It is called a contravariant p−tensor. A
(0, q) tensor has components with q lower indices. It is called a
covariant q−tensor. 2
It is possible to add tensors of the same type, but not of different
types,
a ···a a ···a a ···a
Ab11···bqp + Bb11···bqp = (A + B)b11···bqp . (12.7)
• A tensor field is a rule giving a tensor at each point. 2
We can now define the Lie derivative of a tensor field by using
Leibniz rule in a chart. Let us first consider the components of a
tensor field in a chart. For a (1, 2) tensor field A , the components
in a chart are
∂ ∂
Akij = A( i , j ; dxk ) . (12.8)
∂x ∂x
The components are functions of x in a chart. Thus we can write
this tensor field as
∂
A = Akij dxi ⊗ dxj ⊗ , (12.9)
∂xk
where the × indicates a ‘product’, in the sense that its action on two
vectors and a 1-form is a product of the respective components,
∂
dx ⊗ dx ⊗ k (u, v; ω) = ui v j ωk .
i j
(12.10)
∂x
0
Under a change of charts, i.e. coordinate system xi → x0i , the
components of the tensor field change according to
∂ 0 0 0 ∂
A = Akij dxi ⊗ dxj ⊗ = Aki0 j 0 dx0i ⊗ dx0j ⊗ 0k0 (12.12)
∂xk ∂x
Since 0
0 ∂x0i i ∂ ∂xi ∂
dx0i = dx , = (12.13)
∂xi ∂x0i0 ∂x0i0 ∂xi
0
(i and i are not equal in general), we get
0 0
∂ 0 ∂x0i i ∂x0j ∂xk ∂
Akij dxi ⊗ dxj ⊗ k
= Aki0 j 0 i
dx ⊗ j
dxj ⊗ 0k0 k .
∂x ∂x ∂x ∂x ∂x
(12.14)
Equating components, we can write
0 0
0 ∂x0i ∂x0j ∂xk
Akij = Aki0 j 0 (12.15)
∂xi ∂xj ∂x0k0
0
0 ∂xi ∂xj ∂x0k
Aki0 j 0 = Akij . (12.16)
∂x0i0 ∂x0j 0 ∂xk
∂ ∂f
From now on, we will use the notation ∂i for and ∂i f for
∂xi ∂xi
unless there is a possibility of confusion. This will save some space
and make the formulae more readable.
We can calculate the Lie derivative of a tensor field (with respect
to a vector field u, say) by using the fact that £u is a derivative on
the modules of vector fields and 1-forms, and by assuming Leibniz
rule for tensor products. Consider a tensor field
m···n
T = Ta···b ∂m ⊗ · · · ⊗ ∂n ⊗ dxa ⊗ · · · ⊗ dxb . (12.17)
Then
m···n
£u T = (£u Ta···b ) ∂m ⊗ · · · ⊗ ∂n ⊗ dxa ⊗ · · · ⊗ dxb
m···n
+Ta···b (£u ∂m ) ⊗ · · · ⊗ ∂n ⊗ dxa ⊗ · · · ⊗ dxb + · · ·
m···n
+Ta···b ∂m ⊗ · · · ⊗ ∂n ⊗ (£u dxa ) ⊗ · · · ⊗ dxb , + · · ·
(12.18)
where the dots stand for the terms involving all the remaining up-
per and lower indices. Since the components of a tensor field are
functions on the manifold, we have
m···n
£u Ta···b = ui ∂i Ta···b
m···n
, (12.19)
46 Chapter 12. Tensors
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 13
Differential forms
ω (v1 , · · · , vi , · · · , vj , · · · , vp ) = −ω (v1 , · · · , vj , · · · , vi , · · · , vp )
(13.1)
for any pair i, j . 2
A 0-form is defined to be a function, i.e. an element of C (M) , ∞
Similarly, for a p-form ω , the components are ωi1 ···ip , and compo-
nents are multiplied by (−1) whenever any two indices are inter-
changed.
n
It follows that a p-form has independent components in n-
p
dimensions.
Any 1-form produces a function when acting on a vector field. So
given a pair of 1-forms A, B, it is possible to construct a 2-form ω
47
48 Chapter 13. Differential forms
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
by defining
ω = A ∧ B = −B ∧ A , (13.4)
1
ωij dxi ⊗ dxj − dxj ⊗ dxi (u, v)
ω(u, v) =
2!
1
= ωij ui v j − uj v i = ωij ui v j .
(13.9)
2!
49
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
2
A 2-form in three dimensions can be written as
1
ω = ωij dxi ∧ dxj
2!
= ω12 dx1 ∧ dx2 + ω23 dx2 ∧ dx3 + ω31 dx3 ∧ dx1 (13.15)
50 Chapter 13. Differential forms
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
2
In three dimensions, consider two 1-forms α = αi dxi , β = βi dxi .
Then
1
α∧β = (αi βj − αj βi ) dxi ∧ dxj
2!
= αi βj dxi ∧ dxj
= (α1 β2 − α2 β1 ) dx1 ∧ dx2
+ (α2 β3 − α3 β2 ) dx2 ∧ dx3
+ (α3 β1 − α1 β3 ) dx3 ∧ dx1 . (13.16)
The components are like the cross product of vectors in three dimen-
sions. So we can think of the wedge product as a generalization of
the cross product.
• We can also define the wedge product of a p−form α and a
q−form β as a (p + q)−form satisfying, for any p + q vector fields
v1 , · · · , vp+q ,
1 X
α ∧ β (v1 , · · · , vp+q ) = (−1)deg P α ⊗ β (P (v1 , · · · , vp+q )) .
p!q!
P
(13.17)
Here P stands for a permutation of the vector fields, and deg P is 0 or
1 for even and odd permutations, respectively. In the outer product
on the right hand side, α acts on the first p vector fields in a given
permutation P , and β acts on the remaining q vector fields. 2
The wedge product above can also be defined in terms of the
components of α and β in a chart as follows.
1
α= αi ···i dxi1 ∧ · · · ∧ dxip
p! 1 p
1
β = βj ···j dxj1 ∧ · · · ∧ dxjq
q! 1 q
1
αi1 ···ip βj1 ···jq dxi1 ∧ · · · ∧ dxip ∧ dxj1 ∧ · · · ∧ dxjq .
α∧β =
p!q!
(13.18)
Note that α ∧ β = 0 if p + q > n , and that a term in which some i
is equal to some j must vanish because of the antisymmetry of the
wedge product.
a × (b × c) = (a × b) × c . (13.20)
α ∧ β = (−1)pq β ∧ α . (13.21)
α ∧ β = (−1)pq β ∧ α (13.24)
as wanted. 2
• The wedge product defines an algebra on the space of differential
forms. It is called a graded commutative algebra . 2
• Given a vector field v , we can define its contraction with a
p-form by
ιv ω = ω(v, · · · ) (13.25)
with p − 1 empty slots. This is a (p − 1)-form. Note that the position
of v only affects the sign of the contracted form. 2
Example: Consider a 2-form made of the wedge product of two
ϕ∗ (λ ∧ µ) = ϕ∗ λ ∧ ϕ∗ µ
ϕ∗ (dxi ∧ dxj ) = ϕ∗ dxi ∧ ϕdxj . (13.34)
and we can continue this for any number of basis 1-forms. So for any
p-form ω , let us define the pullback ϕ∗ ω by
Chapter 14
Exterior derivative
This should also satisfy Leibniz rule, but the algebra of p-forms is
not a commutative algebra but a graded commutator algebra, i.e.,
involves a factor of (−1)pq for exchanges. So we need
or alternatively,
This will be the Leibniz rule for wedge products. Note that it gives
the correct result when one or both of α, β are 0-forms, i.e., functions.
The two formulas are identical by virtue of the fact that dβ is a
(q + 1)-form, so that
α ∧ dβ = (−1)p(q+1) dβ ∧ α . (14.4)
We will try to define the exterior derivative in a way such that it has
these properties.
54
55
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
This clearly has the first property of linearity. To check the (graded)
Leibniz rule, let us write α ∧ β in components. Then
1
∂i αi1 ···ip βj1 ···jq dxi ∧ dxi1 ∧ · · · ∧ dxjq
d(α ∧ β) =
p!q!
1
∂i αi1 ···ip βj1 ···jq + αi1 ···ip ∂i βj1 ···jq dxi ∧ dxi1 ∧ · · · ∧ dxjq
=
p!q!
1
∂i αi1 ···ip βj1 ···jq dxi ∧ dxi1 ∧ · · · ∧ dxip ∧ dxj1 ∧ · · · dxjq
=
p!q!
1
(−1)p αi1 ···ip ∂i βj1 ···jq dxi1 ∧ · · · ∧ dxip ∧ dxi ∧ dxj1 ∧ · · · dxjq
+
p!q!
= dα ∧ β + (−1)p α ∧ dβ . (14.6)
dA = (dAν ) ∧ dxν
= ∂µ Aν dxµ ∧ dxν
1
= (∂µ Aν − ∂ν Aµ ) dxµ ∧ dxν
2
⇒ (dA)µν = ∂ µ Aν − ∂ ν Aµ . (14.10)
56 Chapter 14. Exterior derivative
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Then
2x2 2y 2
1 1
dα = − dx ∧ dy − − dy ∧ dx
x2 + y 2 (x2 + y 2 )2 x2 + y 2 (x2 + y 2 )2
2 x2 + y 2
= 2 dx ∧ dy − 2 dx ∧ dy = 0 . (14.15)
x + y2 (x2 + y 2 )2
Introduce polar coordinates r, θ with x = r cos θ , y = r sin θ .
Then
r cos θ (sin θdr + r cos θdθ) r sin θ (cos θdr − r sin θdθ)
α= −
r2 r2
2 2 2
r cos θ + sin θ dθ
= = dθ . (14.16)
r2
Thus α is exact, but θ is multivalued so there is no function f
such that α = df everywhere. In other words, α = dθ is exact only
in a neighbourhood small enough that θ remains single-valued.
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 15
Volume form
n
The space of p-forms in n dimensions is dimensional. So the
p
space of n-forms in n dimensions is 1-dimensiona, i.e., there is only
one independent component, and all n-forms are scalar multiples of
one another.
Choose an n-form field. Call it ω. Suppose ω 6= 0 at some point
P . Then given any basis {eµ } of TP M , we have ω(e1 , · · · en ) 6= 0
since ω 6= 0 . Thus all vector bases at P fall into two classes, one for
which ω(e1 , · · · en ) > 0 and the other for which it is < 0 .
Once we have identified these two classes, they are independent of
ω . That is, if ω 0 is another n-form which is non-zero at P , there must
be some function f 6= 0 such that ω 0 = f ω . Two bases which gave
positive numbers under ω will give the same sign — both positive or
both negative — under ω 0 and therefore will be in the same class.
• So every basis (set of n linearly independent vectors) is a mem-
ber of one of the two classes. These are called righthanded and
lefthanded. 2
• A manifold is called orientable if it is possible to define a con-
tinuous n-form field ω which is non-zero everywhere on the manifold.
Then it is possible to choose a basis with the same handedness ev-
erywhere on the manifold continuously. 2
Euclidean space is orientable, the Möbius band is not.
• An orientable manifold is called oriented once an orientation
has been chosen, i.e. once we have decided to choose basis vectors
58
59
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
= ω(cell) . (15.1)
Adding up the contributions from all cells and taking the limit of cell
size going to zero, we find
Z Z
ω= f dn x . (15.2)
U ϕ(U )
∂x1 1 ∂x1 2
dx1 = dy + 2 dy
∂y 1 ∂y
∂x 2 ∂x2 2
dx2 = dy 1
+ dy
∂y 1 ∂y 2
1 2
∂x1 ∂x2
1 2 ∂x ∂x
⇒ dx ∧ dx = − 2 2 dy 1 ∧ dy 2
∂y 1 ∂y 1 ∂y ∂y
1 2
= Jdy ∧ dy , (15.3)
Chapter 16
Metric tensor
ii) symmetric:
g(v, w) = g(w, v); (16.2)
iii) non-degenerate:
g(v, w) = 0 ∀w ⇒ v = 0. (16.3)
2
• If for some v, w 6= 0 , we find that g(v, w) = 0 , we say that v, w
are orthogonal. 2
• Given a metric g on V , we can always find an orthonormal
basis {eµ } such that g(eµ , eν ) = 0 if µ 6= ν and ±1 if µ = ν . 2
• If the number of (+1)’s is p and the number of (−1)’s is q , we
say that the metric has signature (p, q) .
We have defined a metric for a vector space. We can generalize
this definition to a manifold M by the following.
61
62 Chapter 16. Metric tensor
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
where vµ = gµν v ν .
T µν ←→ T µν ←→ Tµ ν ←→ Tµν (16.6)
µνρ ···
T ←→ T µν ρ ··· ←→ T µν
ρ ··· ←→ Tµνρ ···
←→ · · · (16.7)
2
Given a manifold with metric, there is a canonical volume form
dV (sometimes written as vol) , which in a coordinate chart reads
q
dV = | det gµν |dx1 ∧ · · · ∧ dxn . (16.10)
Note that despite the notation, this is not a 1-form, nor the gradient
of some function V . This is clearly a volume form because it is an
n-form which is non-zero everywhere, as gµν is non-degenerate.
We need to show that this definition is independent of the chart.
Take an overlapping chart. Then in the new chart, the corresponding
volume form is
q
dV 0 = | det gµν0 |dx01 ∧ · · · ∧ dx0n . (16.11)
α β
= g A−1 µ ∂α , A−1 ν ∂β
α β
= A−1 µ A−1 ν gαβ . (16.13)
64 Chapter 16. Metric tensor
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Thus q q
0 −1
| det gµν | = |det A| | det gµν | , (16.15)
and so dV 0 = dV .
• This is called the metric volume form and written as
p
dV = |g|dx1 ∧ · · · ∧ dxn (16.16)
in a chart. 2
When we write dV , sometimes p we mean the n-form as defined
above, and sometimes we mean |g|dn x , the measure for the usual
integral. Another way of writing the volume form in a chart is in
terms of its components,
p
|g|
dV = µ1 ···µn dxµ1 ∧ · · · ∧ dxµn (16.17)
n!
where is the totally
p antisymmetric Levi-Civita symbol, with
12···n = +1 . Thus |g| µ1 ···µn are the components of the volume
form.
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Chapter 17
Hodge duality
where ω is a p-form. 2
The ? operator acts on forms, not on components.
Example: Consider R3 with metric +++, i.e. gµν =
diag(1, 1, 1) . Then |g| ≡ g = 1 , g µν diag(1, 1, 1) . Write the coordi-
nate basis 1-forms as dx, dy, dz . Their components are clearly
(17.3)
65
66 Chapter 17. Hodge duality
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
2
Example: p = n . Then
p
|g|
(?ω) = µ1 ···µn g µ1 ν1 · · · g µn νn ων1 ···νn . (17.5)
n!
For the volume form,
p
|g|
dV = µ1 ···µn dxµ1 ∧ · · · ∧ dxµn
p n!
(dV )ν1 ···νn = |g|ν1 ···νn
|g|
(?dV ) = µ ···µ g µ1 ν1 · · · g µn νn ν1 ···νn
n! 1 n
|g| |g| n!
= n!(det g)−1 = = sign(g) = (−1)s ,(17.6)
n! n! g
where s is the number of (−1) in gµν . 2
So we find that
and
?(?dV ) = (−1)s (?1) = (−1)s dV , (17.8)
i.e., (?)2 = (−1)s on 0-forms and n-forms.
In general, on a p-form in an n-dimensional manifold with signa-
ture (s, n − s) , it can be shown in the same way that
where the sum over I means a sum over all possible index sets I =
I1 < · · · < Ip , but there is no sum over the indices {I1 , · · · , Ip }
themselves, in a given index set the Ik are fixed. Using the dual of
basis p−forms, and Eq. (13.13), we get
X
?ω = ωI1 ···Ip ? (dxI1 ∧ · · · ∧ dxIp )
I
p
X |g|
= ν ···ν µ ···µ g µ1 I1 · · · g µp Ip ωI1 ···Ip dxν1 ∧ · · · ∧ dxνn−p .
(n − p)! 1 n−p 1 p
I
(17.13)
The sum over I is a sum over different index sets as before, and
the Greek indices are summed over as usual. Thus we calculate
p
|g| X
?ω ∧ ω = ν1 ···νn−p µ1 ···µp g µ1 I1 · · · g µp Ip ωI1 ···Ip ×
(n − p)!
I,J
dxν1 ∧ · · · ∧ dxνn−p ∧ ωJ1 ···Jp dxJ1 ∧ · · · ∧ dxJp
p
|g| X
We see that the set {ν1 , · · · , νn−p } cannot have any overlap with the
set J = {J1 , · · · , Jp }, because of the wedge product. On the other
hand, {ν1 , · · · , νn−p } cannot have any overlap with {µ1 , · · · , µp } be-
cause is totally antisymmetric in its indices. So the set {µ1 , · · · , µp }
must have the same elements as the set J = {J1 , · · · , Jp } , but they
may not be in the same order.
Now consider the case where the basis is orthogonal, i.e. g µν is
diagonal. Then g µk Ik = g Ik Ik etc. and we can write
p
|g| X
?ω ∧ ω = ν1 ···νn−p I1 ···Ip g I1 I1 · · · g Ip Ip ωI1 ···Ip ωJ1 ···Jp ×
(n − p)!
I,J
dxν1 ∧ · · · ∧ dxνn−p ∧ dxJ1 ∧ · · · ∧ dxJp . (17.15)
We see that in each term of the sum, the indices {I1 · · · Ip } must be
the same as {J1 · · · Jp } because both sets are totally antisymmetrized
with the indices {ν1 · · · νn−p }.
Since both sets are ordered, it follows that we can replace J by
I,
p
|g| X
?ω ∧ ω = ν1 ···νn−p I1 ···Ip g I1 I1 · · · g Ip Ip ωI1 ···Ip ωI1 ···Ip ×
(n − p)!
I
In each term of this sum, the indices {ν1 · · · νn−p } are completely
determined, so we can replace them by the corresponding ordered
set K = K1 < · · · < Kn−p , which is completely determined by the
set I , so that
p X
?ω ∧ ω = |g| K1 ···Kn−p I1 ···Ip ω I1 ···Ip ωI1 ···Ip ×
I
dxK1 ∧ · · · ∧ dxKn−p ∧ dxI1 ∧ · · · ∧ dxI(17.17)
p
.
Chapter 18
Maxwell equations
∇·E =ρ (18.1)
∂E
∇×B− =j (18.2)
∂t
∇·B =0 (18.3)
∂B
∇×E+ = 0. (18.4)
∂t
The electric and magnetic fields are all vectors in three dimensions,
but these equations are Lorentz-invariant. We will write these equa-
tions in terms of differential forms.
Consider R4 with Minkowski metric gµν = diag(−1, 1, 1, 1) . For
the magnetic field define a 2-form
B = Bx dy ∧ dz + By dz ∧ dx + Bz dx ∧ dy . (18.5)
E = Ex dx + Ey dy + Ez dz . (18.6)
70
71
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
?F = ?(B + E ∧ dt)
= B1 dt ∧ dx + B2 dt ∧ dy + B3 dt ∧ dz
+E1 dy ∧ dz + E2 dz ∧ dx + E3 dx ∧ dy . (18.11)
72 Chapter 18. Maxwell equations
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
d?F = (∇ · E) dx ∧ dy ∧ dz
+ (∂t E1 − (∇ × B)1 ) dt ∧ dy ∧ dz
+ (∂t E2 − (∇ × B)2 ) dt ∧ dz ∧ dx
+ (∂t E3 + (∇ × B)3 ) dt ∧ dx ∧ dy . (18.12)
j µ ∂µ = ρ∂t + j 1 ∂1 + j 2 ∂2 + j 3 ∂3 . (18.13)
?j = −ρ dx ∧ dy ∧ dz + j1 dt ∧ dy ∧ dz
+j2 dt ∧ dz ∧ dx + j3 dt ∧ dx ∧ dy . (18.15)
Comparing this equation with Eq. (18.12) we find that the other two
Maxwell equations can be written as
Chapter 19
Stokes’ theorem
73
74 Chapter 19. Stokes’ theorem
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
A = Ai dxi
dA = ∂i Aj dxi ∧ dxj (19.4)
d d
A evaluated on ∂D can be written as A where is tangent
dt dt
dxi
d
to ∂D . So we can write A = Ai dt , and
dt dt
Z Z
i
Ai dx = ∂i Aj dxi ∧ dxj
∂D D
Z
= (∂1 A2 − ∂2 A1 ) dx1 ∧ dx2
D
Z
= (∂1 A2 − ∂2 A1 ) d2 x . (19.5)
ϕ(D)
ω = f dx1 ∧ · · · ∧ dxn
1
= f µ1 ···µn dxµ1 ∧ · · · ∧ dxµn . (19.6)
n!
75
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
(−1)n−1 p
= µ1 ···µn−1 µ ∂µn |g|v µ dxµ1 ∧ · · · ∧ dxµn
(n − 1)!
(19.12)
76 Chapter 19. Stokes’ theorem
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
because all terms of the form b ∧ ιv α gives zero for any choice of
(n − 1) vectors on ∂U .
Then we have
Z Z
1 p µ
p ∂µ ( |g|v )(vol) = b(v)α
|g|
U ∂U
Z
= (nµ v µ ) α . (19.18)
∂U
Chapter 20
Lie groups
78
79
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
I} . 2
Example: The Symplectic group Sp(n) , defined as the sub-
group of U (2n) given by AT JA = J , where
−1n×n
0
J=
1n×n 0
2
Example: O(p, q) = {R ∈ GL(p+q, R) | RT ηp,q R = ηp,q } , where
1p×p
0
ηp,q =
0 −1q×q
2
Example: U (p, q) = {U ∈ GL(p + q, C) | U † ηp,q U = ηp,q } .
Example The Special Orthogonal group SO(n) is the sub-
group of O(n) for which determinant is +1. Similarly, the Special
unitary group SU (n) is the subgroup of U (n) with determinant
+1. Similarly for SO(p.q) and SU (p, q) .
The group U (1) is the group of phases U (1) = {eiφ |φ ∈ R} . As a
manifold, this is isomorphic to a circle S 1 .
Chapter 21
X = rg∗ (X) ∀g ∈ G ,
i.e. rg∗ (Xg0 ) = Xg0 g ∀g, g 0 ∈ G . (21.3)
81
82 Chapter 21. Tangent space at the identity
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Thus if X, Y ∈ L(G) ,
LX g ≡ LX
g := lg∗ X ∀g ∈ G (21.9)
lg0 ∗(LX X
g ) = lg 0 ∗ (lg∗ X) = lg 0 g∗ X = Lg 0 g . (21.10)
LX Y
g = Lg ∀g ∈ G , (21.13)
so
lg−1 ∗ LX Y
g = lg −1 ∗ Lg ⇒ X=Y (∈ Te G) . (21.14)
Xe = lg−1 ∗ LX
g for any g ∈ G . (21.15)
Xe = LX
e . (21.16)
Then
[u , v] = [Lu , Lv ]e .
(21.18)
Note that since commutators are defined for vector fields and not
vectors, the Lie bracket on Te G has to be defined using the com-
mutator of left-invariant vector fields on G and the isomorphism
Te G ↔ L(G) .
• If for an n-dimensional Lie group G , {t1 , · · · , tn } is a set of basis
vectors on Te G ' L(G) , the Lie bracket of any pair of these vectors
for some set of real numbers Cijk . These numbers are known as the
[ti , tj ] = [tj , ti ]
X X
k k
⇒ Cij tk = Cji tk
k k
k k
⇒ Cij = Cji , (21.21)
Chapter 22
85
86 Chapter 22. One parameter subgroups
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
γ X (0) = LX
e =X
and γ̇ X (t) = LX
γ(t) = lγ(t)∗ X (22.1)
We see that γ(t + τ ) and γ(τ )γ(t) are both integral curves of LX , i.e.
both satisfy the equation of the integral curve of LX , and at t = 0
both curves are at the point γ(τ ) . Thus by uniqueness these two are
the same curve,
γ̇(t) = LX
γ(t) = lγ(t)∗ X , (22.6)
γ(t) = I + tM (22.11)
γ(a) = exp ai ti
(22.12)
[Mµν , Mαβ ] = δνα Mµβ − δµα Mνβ + δµβ Mνα − δνβ Mµα . (22.15)
Chapter 23
Fiber bundles
S
Consider a manifold M with the tangent bundle T M = TP M .
P ∈M
Let us look at this more closely. T M can be thought of as the original
manifold M with a tangent space stuck at each point P ∈ M . Thus
there is a projection map π : T M → M , TP M 7→ P , which
associates the point P ∈ M with TP M .
Then we can say that T M consists of pooints P ∈ M and vectors
v ∈ TP M as an ordered pair (P, vP ) . Then in the neighbourhood of
any point P , we can think of T M as a product manifold, i.e. as
the set of ordered pairs (P, vP ) .
This is generalized to the definition of a fiber bundle. Locally
a fiber bundle is a product manifold E = B × F with the following
properties.
• B is a manifold called the base manifold, and F is another
manifold called the typical fiber or the standard fiber.
• There is a projection map π : E → B , and if P ∈ B , the pre-
image π −1 (P ) is homeomorphic, i.e. bicontinuously isomorphic, to
the standard fiber. 2
E is called the total space, but usually it is also called the bun-
dle, even though the bundle is actually the triple (E, π, B) .
• E is locally a product space. We express this in the following
way. Given an open set Ui of B , the pre-image π −1 (Ui ) is homeomor-
phic to Ui ×F , or in other words there is a bicontinuous isomorphism
ϕi : π −1 (Ui ) → Ui × F . The set {Ui , ϕi } is called a local trivializa
89
90 Chapter 23. Fiber bundles
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
ei = aji ej . (23.1)
gij = Φi ◦ Φj −1 : Ui ∩ Uj → G . (23.2)
In the other notation we have been using, vα and vβ are images of the
same vector vx ∈ Vx , and vβ = D(gβα )vα . A gauge transformation
T acts as
Chapter 24
Connections
Dv (s + αt) = Dv s + α Dv t
Dv (f s) = v(f )s + f Dv s
Dv+f w (s) = Dv s + f Dw s , (24.2)
95
96 Chapter 24. Connections
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
= ∂µ si ei + si Aµij ej
= ∂µ si + Aµj i sj ei , (24.5)
= gDv g −1 φ ,
(24.8)
and thus D0 is defined on all φ for which Dv φ is defined, i.e. D0 exists
because D does. We have of course assumed that g(x) is differentiable
as a function of x.
Let us now check that D0 is a connection according to our defi-
nitions. D0 is linear since
Dv0 (φ1 + αφ2 ) = gDv g −1 (φ1 + αφ2 )
= gDv g −1 φ1 + αgDv g −1 φ2
= gDv f g −1 φ
= gv(f )g −1 φ + gf Dv g −1 φ
= v(f )φ + f gDv g −1 φ
= gDv (g −1 φ) + αgDw g −1 φ
= Dv0 φ + αDw
0
φ. (24.11)
So D0 is a connection, i.e. there is a connection D0 satisfying
Dv0 (gφ) = g (Dv φ) .
Since φ is a section, i.e. φ ∈ Γ(E), so is g −1 φ, and thus
Dv g −1 φ ∈ Γ(E) and also gDv g −1 φ ∈ Γ(E) . Therefore, Dv0 maps
sections to sections, Dv0 : Γ(E) → Γ(E) . This completes the defini-
Aµ = Aµij ei ⊗ θj ,
A0µ = A0µij ei ⊗ θj , (24.13)
then given by
Dv0 φ = gDv g −1 φ
Dµ0 φ = gDµ g −1 φ
⇒
h i j i
∂µ φi + A0µij ei = g ∂µ g −1 φ + Aµij g −1 φ ei ,
⇒
(24.14)
Chapter 25
Curvature
Remember that
D v s = D v µ ∂µ s = v µ D µ s
= v µ ∂µ si + Aµij sj ei
= v(si )ei + v µ Aµij sj ei . (25.2)
99
100 Chapter 25. Curvature
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Also,
D[v ,w] s = [v , w](si )ei + [v , w]µ Aµij sj ei , (25.6)
so that
F (v, w)s = v µ wν ∂µ Aν ij − ∂ν Aµij + Aµik Aν kj − Aν ik Aµkj sj ei .
(25.7)
Thus we can define Fµν by
so that
• It is not very difficult to work out that the curvature acts linearly
on the module of sections,
Thus
Dλ (Fµν s) − Fµν Dλ s = [Dλ , [Dµ , Dν ]]s . (25.16)
Considering C ∞ sections, and noting that maps are associative under
map composition, we find that
where we have defined (Dλ Fµν ) by (Dλ Fµν )ij in this. Then this is a
Leibniz rule,
0
Given D and g such that g(x) ∈ G , we have D given by
Dv0 φ = gDv g −1 φ .
(25.22)
102 Chapter 25. Curvature
c Amitabha Lahiri: Lecture Notes on Differential Geometry for Physicists 2011
Then
and thus
F 0 (u, v) φ ≡ Du0 Dv0 − Dv0 Du0 − D[u
0
,v] φ
= gF (u, v)g −1 φ
0
⇒ Fµν = g ◦ Fµν ◦ g −1 . (25.24)