Curvature: 3.1 Covariant Derivative
Curvature: 3.1 Covariant Derivative
Curvature
Problem Set #3: 3.1, 3.3, 3.5, 3.13, 3.15 (Due Monday Oct. 28th)
Midterm-exam: October 30th
29
CHAPTER 3. CURVATURE 30
µ ′ ′
∂x ∂ ∂x ν ν′ ∂x λ ∂xµ ∂xν ∂ ν ∂xµ ∂xν ν λ
V + Γ µ′ λ′ V = V + Γ V
∂xµ′ ∂xµ ∂xν ∂xλ ∂xµ′ ∂xν ∂xµ ∂xµ′ ∂xν µλ
′ λ′ ′
∂xµ ∂ ∂xν ν ν ′ ∂x λ ∂xµ ∂xν ν λ
V + Γ µλ V = Γ V
∂xµ′ ∂xν µλ
′ ′
∂xµ′ ∂xµ ∂xν ∂xλ
′ ′
′ ′ ∂xλ ∂xµ ∂xν ν λ′ ∂xλ ∂xµ ∂ ∂xν λ′
Γνµ′ λ′ V λ = Γ V − (3.5)
V .
∂xλ′ ∂xµ′ ∂xν µλ ∂xλ′ ∂xµ′ ∂xµ ∂xλ
′
Since this must be true for all V λ we get a transformation law for the con-
nection coefficients
′ ′
′ ∂xλ ∂xµ ∂xν ν ∂xλ ∂xµ ∂ 2 xν
Γνµ′ λ′ = λ′ µ′ Γ µλ − . (3.6)
∂x ∂x ∂xν ∂xλ′ ∂xµ′ ∂xµ ∂xλ
∇µ ων = ∂µ ων + Γ̃λµν ωλ (3.7)
where Γ̃λµν although so far unrelated does transforms exactly as Γλµν in (3.6).
To establish a relation between Γλµν and Γ̃λµν we demand that the covariant
derivative reduces to partial derivative for scalars
∇µ φ = ∂µ φ. (3.8)
Thus
∂µ ωλ V λ = ∇µ ωλ V λ
! " ! "
(∂µ ωλ ) V λ + ∂µ V λ ωλ = (∇µ ωλ ) V λ + ∇µ V λ ωλ
! " ! "
and therefore
∇µ ων = ∂µ ων − Γλµν ωλ . (3.11)
CHAPTER 3. CURVATURE 31
∇ρ gµν = 0. (3.14)
∇ρ g µν = 0 (3.15)
and thus the raising an lowering operators commute with covariant derivative
∇ρ V µ = ∇ρ (g µν Vν ) = g µν ∇ρ Vν . (3.16)
∂ρ gµν − ∂µ gνρ − ∂ν gµρ = Γλρµ gλν − Γλµρ gνλ + Γλρν gµλ − Γλνρ gλµ − Γλµν gλρ + Γλνµ
! " ! " ! "
(3.20)
gρλ
∂ρ gµν − ∂µ gνρ − ∂ν gµρ = −2Γλµν gλρ (3.21)
CHAPTER 3. CURVATURE 32
or
1
Γλµν = g λρ (∂µ gνρ + ∂ν gµρ − ∂ρ gµν ) . (3.22)
2
In Minkowski space described by Cartesian all of the Christoffel symbols
vanish, but this does not have to be the case in curvilinear coordinates. As
an example consider a two dimensional Euclidean space described by polar
coordinates with metric
ds2 = (dr)2 + r 2 (dθ)2 . (3.23)
The non-vanishing components of the inverse metric are
g rr = 1 (3.24)
g θθ = r −2 (3.25)
and for example the connection coefficient
1 rρ
Γrrr = g (∂r grρ + ∂r gρr − ∂ρ grr ) =
2
1 rr 1
= g (2∂r grρ − ∂ρ grr ) + g rθ (2∂r grρ − ∂ρ grr ) =
2 2
1 1
= 1 (2∂r grr − ∂r grr ) + 0 (2∂r grσ − ∂σ grr ) = 0 (3.26)
2 2
but
1 rρ
Γrθθ = g (∂θ gθρ + ∂θ gρθ − ∂ρ gθθ ) =
2
1 rr 1
= g (2∂θ gθr − ∂r grr ) + g rθ (2∂θ gθθ − ∂θ gθθ ) =
2 2
1 1
= 1 (2∂θ gθr − ∂r grr ) + 0 (2∂θ gθρ − ∂ρ gθθ ) =
2 2
1
= − ∂r grr = −r. (3.27)
2
It is a straightforward exercise to find all other coefficients,
Γrθr = Γrrθ = 0 (3.28)
Γθrr = 0 (3.29)
1
Γθrθ = Γθθr = (3.30)
r
θ
Γθθ = 0. (3.31)
Just like one can make the connection coefficients to be non-zero in flat
space it is possible to make the connection coefficients to vanish at some
point curved space but not everywhere.
CHAPTER 3. CURVATURE 33
∇µ V µ = ∂µ V µ + Γµµλ V λ (3.32)
dxµ
# $
D
T µ1 ...µk ν1 ...νl = ∂µ T µ1 ...µkν1 ...νl = 0 (3.35)
dλ dλ
dxµ
# $
D
T µ1 ...µk ν1 ...νl = ∇µ T µ1 ...µkν1 ...νl = 0. (3.36)
dλ dλ
CHAPTER 3. CURVATURE 34
This is the parallel transport equation which, for example, take the fol-
lowing form for vectors
D µ d µ dxσ ρ
V = V + Γµσρ V = 0. (3.37)
dλ dλ dλ
It follows, for example, that the inner product of two parallel transported
vectors is preserved, i.e.
D D D D
(gµν V µ W ν ) = (gµν ) V µ W ν + gµν (V µ ) W ν + gµν V µ (V W ν ) = 0.
dλ dλ dλ dλ
(3.38)
Next we will obtain a formal solution of the parallel transport equa-
tion (3.37). Our task is to find the so-called parallel propagator matrix
P µρ (λ0 , λ) along trajectory γ(λ) such that
V µ (λ) = P µρ (λ, λ0 )V ρ (λ0 ). (3.39)
If we define a transition matrix
dxσ
Aµρ (λ) ≡ −Γµσρ (3.40)
dλ
then (3.37) can be written as a Schrodinger equation
d µ
V = Aµρ V ρ . (3.41)
dλ
By substituting (3.39) into (3.41) we get
d ( µ
P ρ (λ, λ0 )V ρ (λ0 ) = Aµσ P σρ (λ, λ0 )V ρ (λ0 )
) ( )
(3.42)
dλ
d µ
P (λ, λ0 ) = Aµσ P σρ (λ, λ0). (3.43)
dλ ρ
By integrating both side we get
* λ
P µρ (λ, λ0 ) = δρµ + Aµσ (η)P σρ (η, λ0)dη (3.44)
λ0
or in matrix notation
* λ * λ * η2
P (λ, λ0 ) = 1 + A(η1 )dη1 + A(η2 )A(η1 )dη1 dη2 + ... (3.46)
λ0 λ0 λ0
CHAPTER 3. CURVATURE 35
which can be simplified using a path-ordered product of matrices, P[A(ηn )A(ηn−1 )...A(η1 )],
as ∞
1 λ
+ *
P (λ, λ0) = 1 + P[A(ηn )...A(η1 )]dη1 dη2 ... (3.47)
n=1
n! λ0
which is the series expansion of an exponential,
#* λ $
P (λ, λ0) = P exp A(η)dη . (3.48)
λ0
In quantum field theory the same formula is known as Dyson’s formula which
is due to the fact that (3.41) is mathematically equivalent to the Schrodinger
equation. The parallel transport transformation around a closed loop is called
the holonomy of the connection around the loop. For the metric-compatible
connections the group of holonomy transformations at a point is a Lorentz
group. Note that the knowledge of the holonomy group at each point is
sufficient to determine the metric and thus might question what is more
fundamental: holonomies of all loops or metric at all point. In the canonical
quantum gravity the metric is treated as fundamental, when in the loop
quantum gravity the holonomies are more fundamental.
3.3 Geodesics
Geodesic connecting two points can be defined as a path which parallel trans-
ports its own tangent vector, i.e.
D dxµ d2 xµ ρ
µ dx dx
σ
= + Γ ρσ = 0. (3.50)
dλ dλ dλ2 dλ dλ
This is the geodesic equation which is obtained by substituting the tangent
µ
vector dx
dλ
into (3.37). For the metric-compatible connection coefficients the
geodesic is also the path of the largest proper time (there is no smallest
proper time since one can always construct a path of zero length made out
of null segments) defined as
,
dxµ dxν
*
τ [x] = dλ −gµν (3.51)
dλ dλ
CHAPTER 3. CURVATURE 36
δτ [x]
= 0. (3.52)
δxµ
By Taylor expanding the metric
λ = aτ + b (3.59)
d2 xµ ρ
µ dx dx
σ
dxµ
+ Γ ρσ = f (α) (3.60)
dα2 dα dα dα
where f (α) is some function which depends on the parametrization. Con-
versely if (3.60) is satisfied along some curve, then one can always find an
affine parameter λ for which (3.50) is satisfied.
In addition to parallel transport the geodesics xµ (λ) passing through some
point p can be used to map point from the tangent space Tp to a neighborhood
of p. Such mapping is called the exponential map
expp : Tp → M (3.61)
which is defined as
expp (k µ ) = xν (λ = 1) (3.62)
where
dxµ (λ = 0)
kµ = (3.63)
dλ
CHAPTER 3. CURVATURE 38
is the tangent vector at point p where we have set λ = 0. Of course the map
is well defined and invertible on a subset of Tp sufficiently close to k µ = 0,
and so (3.61) should not be taken literally. The range of the map can fail
to be all of the manifold given that there can be points not connected by a
geodesic, and the range of the map can fail to be all of the tangent space if
the manifold is geodesically incomplete (has singularities or boundaries).
[∇µ , ∇ν ]V ρ . (3.66)
which vanishes for Christoffel connections, and the first term defines the
Riemann tensor,
& '
Rρσµν = 2 ∂[µ Γρν]σ + Γρ[µ|λ Γλσ|ν] = ∂µ Γρνσ − ∂ν Γρµσ + Γρµλ Γλσν − Γρνλ Γλσµ . (3.69)
One can check that it is a legitimate (1,3) tensor, although it is not immedi-
ately clear that it is the same tensor as in (3.64) (see Wald for details). It is
also straightforward to determine the action of the commutator [∇ρ , ∇σ ] on
a tensor of arbitrary rank,
[∇ρ , ∇σ ]X µ1 ...µkν1 ...νl = −2Γλ[ρσ] ∇λ X µ1 ...µkν1 ...νl +Rµ1 λρσ X λµ2 ...µkν1 ...νl ...−Rλν1 ρσ X µ1 ...µkλν2 ...νl ....
(3.70)
Sometimes it useful to express the relevant quantities in a coordinate
independent way. Then one can think of the (1,2) torsion tensor as a map
from two vectors to a third vector,
and the (1,3) Riemann tensor as a map from three vectors to a fourth vector,
[X, Y ]µ ≡ X λ ∂λ Y µ − Y λ ∂λ X µ (3.73)
∇X ≡ X µ ∇ µ . (3.74)
CHAPTER 3. CURVATURE 40
Note that the third term in (3.72) vanishes when X and Y are coordinate
basis since
[∂µ , ∂ν ] = 0. (3.75)
In the case of metric compatible connection the Riemann tensor vanishes
if the metric is constant everywhere on the manifold,
gµν = const ⇒
∂σ gµν = 0⇒
Γρµν = 0⇒
∂σ Γρµν = 0⇒
Rρσµν = 0. (3.76)
at some point p and with basis vectors ê(µ) whose dot product is
This set of basis vectors can be parallel transported to some other point q.
Since the Riemann tensor is zero the result of the transport does not depend
on the trajectory and the dot product remains unchanged,
Then according to the Forbenius’s Theorem one can always find coordinates
xµ such that
∂
ê(µ) = µ (3.82)
∂x
in which the metric would have the form of ηµν .
CHAPTER 3. CURVATURE 41
Not all of the components of the Riemann tensor are independent and
one can show that the components must obey,
Some other useful tensor that can be formed from the Riemann tensor
are:
• Ricci tensor:
Rµν = Rλµλν (3.88)
which is symmetric for Christoffel connections.
• Ricci scalar:
R = Rµµ (3.89)
which can be shown (using Bianchi identity) to obey
1
∇µ Rρµ = ∇ρ R. (3.90)
2
• Einstein tensor:
1
Gµν = Rµν − R gµν (3.91)
2
which is also symmetric for Christoffel connections and can be shown
to obey
∇µ Gµν = 0. (3.92)
• Weyl tensor:
1! " 1
Cρσµν = Rρσµν − gρ[µ Rν]σ − gσ[µ Rν]ρ + R gρ[µ gν]σ (3.93)
2 3
which is essentially the Riemann tensor with all of the contractions
removed (i.e. with zero traces) while maintaining its symmetries.
CHAPTER 3. CURVATURE 42
Then
and
Rθφθφ = a2 sin2 θ. (3.97)
It follows that
Rθθ = g φφ Rφθφθ = 1
Rθφ = Rφθ = 0
Rφφ = g θθ Rθφθφ = sin2 θ (3.98)
and
R = g θθ Rθθ + g φφ Rφφ = 2a−2 . (3.99)
Note that the Ricci scalar is positive as it should be for a positively curved
space such as the two-sphere, but for negatively curved spaces such as a
saddle it would be negative. The considered two-sphere is an example of a
maximally symmetric space generically defined by
xµ (s, t) ∈ M. (3.101)
[S, T ] = 0. (3.104)
This means that they can be used as basis vectors for a coordinate system
of our two-dimensional surface. Then the quantities defined as
and
aµ = (∇T V )µ = T ρ ∇ρ V µ (3.106)
we can call the “relative velocity” and “relative acceleration” of the geodesics.
For a connection with vanishing torsion equation (3.104) implies
S ρ ∇ ρ T µ = T ρ ∇ρ S µ
aµ = T ρ ∇ρ (T σ ∇σ S µ )
= T ρ ∇ρ (S σ ∇σ T µ )
= (T ρ ∇ρ S σ ) ∇σ T µ + T ρ S σ (∇ρ ∇σ T µ )
(S ρ ∇ρ T σ ) ∇σ T µ + T ρ S σ ∇σ ∇ρ T µ + Rµνρσ T ν
! "
=
= (S ρ ∇ρ T σ ) ∇σ T µ + S σ ∇σ (T ρ ∇ρ T µ ) − (S σ ∇σ T ρ ) (∇ρ T µ ) + Rµνρσ T ν T ρ S σ
= S σ ∇σ (T ρ ∇ρ T µ ) + Rµνρσ T ν T ρ S σ
= Rµνρσ T ν T ρ S σ (3.107)
where in the last line we used that T ρ is a tangent vector to geodesic and
thus
T ρ ∇ρ T µ = 0. (3.108)
The resulting equation is known as the geodesic deviation equation,
D2 µ
aµ = S = Rµνρσ T ν T ρ S σ (3.109)
dt2
which can be interpreted as the gravitational tidal force due to curvature
Rµνρσ .
CHAPTER 3. CURVATURE 44
φ∗ f ≡ f ◦ φ. (3.113)
Such map is called the pullback since it is defined by pulling back the action
of function f from manifold N to manifold M (i.e. in the direction opposite
to the way φ was defined).
Clearly, it is not be possible to define a pushforward of a function, but
it is possible to define a pushforward of a vector V which, as we know, is
nothing but a map from functions on manifolds to real numbers. Then the
pushforward of the vector field defined on manifold M, i.e.
V : F (M) → R. (3.114)
φ∗ V : F (N) → R (3.115)
(φ∗ V )α ∂α f = V µ ∂µ (φ∗ f )
= V µ ∂µ (f ◦ φ)
∂y α
= V µ µ ∂α f (3.117)
∂x
or
∂y α
(φ∗ V )α = V µ . (3.118)
∂xµ
CHAPTER 3. CURVATURE 45
(3.128)
CHAPTER 3. CURVATURE 46
or
(φ∗ T ) ω (1) , ..., ω (k), V (1) , ..., V (l) = T φ−1 ∗ ω (1) , ..., φ−1 ∗ ω (k) , φ∗ V (1) , ..., φ∗V (l) .
! " !! " ! " "
(3.129)
−1
Clearly the pullback of φ is nothing but a pushforward of φ .
Consider an example of a two-sphere S 2 with coordinates (θ, ϕ) embedded
into a three dimensional Euclidean space R3 with coordinates (x, y, z) such
that there is a natural map
φ : S 2 → R3 (3.130)
defined by
φ(θ, ϕ) = (sin θ cos ϕ, sin θ sin ϕ, cos θ). (3.131)
Then the metric on R3 as a (0, 2) tensor can be pullback to S 2 . Then the
pullback matrix operator
# $
cos θ cos ϕ cos θ sin φ − sin θ
φ∗ = (3.132)
− sin θ sin ϕ sin θ cos φ 0
such that
φs+t = φs ◦ φt (3.136)
we can consider trajectory of a given point p under φt . These trajectories
xµ (t) can used to define vector fields by evaluating tangent vector at t = 0:
. µ/
dx
V µ (x) = . (3.137)
dt t=0
∆t T µ1 ...µkν1 ...νl (p) = φt∗ T µ1 ...µkν1 ...νl (p) − T µ1 ...µkν1 ...νl (p)
! "
(3.139)
which is a map from (k, l) to (k, l). The Lie derivative is linear
Although the Lie derivative does not depend on the coordinate system it
does not require specification of connections. It is easy to show that the Lie
derivative reduces to an ordinary derivatives for scalars
LV f = V (f ) = V µ ∂µ f (3.143)
Then the Lie derivative of a one-form can be derived by considering the Lie
derivative of a scalar
LV (ωµ U µ ) = V ν ∂ν (ωµ U µ )
(LV ωµ ) U µ + ωµ (LV U)µ = V ν U µ ∂ν ωµ + V ν ωµ ∂ν U µ
(LV ωµ ) U µ + ωµ (V ν ∂ν U µ − U ν ∂ν V µ ) = V ν U µ ∂ν ωµ + V ν ωµ ∂ν U µ
(LV ωµ ) U µ = V ν U µ ∂ν ωµ + ωµ U ν ∂ν V µ
L V ωµ = V ν ∂ν ωµ + ων ∂µ V ν . (3.145)
LV T µ1 ...µkν1 ...νl = V σ ∂σ T µ1 ...µkν1 ...νl −(∂λ V µ1 ) T λµ2 ...µkν1 ...νl ...+ ∂ν1 V λ T µ1 ...µkλν2 ...νl ....
! "
(3.146)
Although the above expression is coordinate independent one can rewrite it
in a more covariant form
LV T µ1 ...µkν1 ...νl = V σ ∇σ T µ1 ...µkν1 ...νl −(∇λ V µ1 ) T λµ2 ...µkν1 ...νl ...+ ∇ν1 V λ T µ1 ...µkλν2 ...νl ....
! "
(3.147)
A diffeomorphism φ is called a symmetry of some tensor T if it leaves
the tensor invariant, i.e.
φ∗ T = φ∗ T = T. (3.148)
If the symmetry is generated by a vector field V µ (x) then the symmetry
implies that the Lie derivative vanishes
LV T = 0. (3.149)
LV gµν = 0 (3.151)
or
has three Killing vectors (i.e. satisfy the Killing equations (3.152)) corre-
sponding to two translation
X = (1, 0) (3.154)
Y = (0, 1) (3.155)
U ν ∇ν (Vµ U µ ) = U ν U µ ∇ν Vµ + Vµ (U ν ∇ν U µ )
1 ν µ
= (U U ∇ν Vµ + U µ U ν ∇µ Vν )
2
= 0. (3.158)
3.7 Tetrads
So far we were using the coordinate basis for the tangent space
ê(µ) = ∂µ (3.159)
where the eaµ matrix an invertible matrix whose inverse eµa expresses tetrad
basis vectors in coordinate basis
As usual (with the abuse of notations) the component eaµ and eµa are often
called the tetrads and inverse tetrads respectively. Clearly, they must satisfy
or
eaµ eµb = δba , (3.165)
and (3.161) can be rewritten in the indices notation as
or
gµν = eaµ ebν ηab (3.167)
which is why tetrads are sometimes called the “square root” of the metric.
After defining the orthonormal basis for vectors we can define an orthonormal
basis for one-forms by setting
Then one can show that using the very same inverse tetrads eµa and tetrads
eaµ we can respectively express
and
θ̂(a) = eaµ θ̂(µ) . (3.170)
CHAPTER 3. CURVATURE 51
Of course, no only the basis vectors but any vectors expressed in terms of
the coordinate basis can be re-expressed in terms of tetrad basis. In other
words
V = V a ê(a) = V µ ê(µ) (3.171)
implies
V a = eaµ V µ . (3.172)
Similarly for tensors
and one example of such transformation we had already seen before for metric
tensor (3.166).
For a non-coordinate tetrad basis we can illustrate how the change of
tetrad basis would change components of tensors even without changing co-
ordinates. If we go from one tetrad basis to another the orthonormality must
be preserved and thus the metric ηab must remain flat. Of course the group
of such transformation is known as Lorentz group and the transformation
matrices are known as Lorentz matrices. In contrast to the global Lorentz
transformations in special relativity we are now free to have different changes
of basis at different points (local Lorentz transformations), i.e.
∇µ X ab = ∂µ X ab + ωµ ac X cb − ωµ c b X ac . (3.177)
CHAPTER 3. CURVATURE 52
Of course in mixed basis one can get terms with connection coefficients and
spin coefficients in the same expression, e.g.
But since the same object (e.g. covariant derivative of vector) can be
written in coordinate basis
and
ωµab = eaν eλb Γνµλ − eλb ∂µ eaλ (3.182)
which is equivalent to
∇µ eaν = 0. (3.183)
Just like the connection coefficients the spin connections are not legitimate
tensors, but the lower Greek index does transform as a one-form. The trans-
formation law for other indices is given by
(dX)µν a = ∂µ Xν a − ∂ν Xµ a (3.185)
CHAPTER 3. CURVATURE 53
is a legitimate tensor. Then one can view torsion as a vector valued two-
form Tµν a and Riemann curvature tensor as a (1,1) valued two-form Rabµν .
By suppressing indices of the differential forms the torsions and curvature
tensors we can written as
T a = dea + ω ab ∧ eb (3.187)
and
Rab = dω ab + ω ac ∧ ω cb (3.188)
known as Maurer-Cartan structure equations. Similarly the Bianchi
identities (3.85) and (3.86) can be written respectively as
dT a + ω ab ∧ T b = Rab ∧ eb (3.189)
and
dRab + ω ac ∧ Rcb − Rac ∧ ω cb = 0. (3.190)
The metric compatibility (i.e. ∇µ gνλ = 0)
dea = −ω ab ∧ eb = ω ab ∧ eb (3.193)
The two equations can be used to solve for the spin connections in terms of
tetrads.