Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
Download as pdf or txt
Download as pdf or txt
You are on page 1of 46

The Reissner-Nordström metric

Jonatan Nordebo

March 16, 2016

Abstract
A brief review of special and general relativity including some classi-
cal electrodynamics is given. We then present a detailed derivation of
the Reissner-Nordström metric. The derivation is done by solving the
Einstein-Maxwell equations for a spherically symmetric electrically
charged body. The physics of this spacetime is then studied. This
includes gravitational time dilation and redshift, equations of motion
for both massive and massless non-charged particles derived from the
geodesic equation and equations of motion for a massive charged par-
ticle derived with lagrangian formalism. Finally, a quick discussion of
the properties of a Reissner-Nordström black hole is given.

1
Contents
1 Introduction 3

2 Review of Special Relativity 3


2.1 4-vectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
2.2 Electrodynamics in Special Relativity . . . . . . . . . . . . . . 8

3 Tensor Fields and Manifolds 11


3.1 Covariant Differentiation and Christoffel Symbols . . . . . . 13
3.2 Riemann Tensor . . . . . . . . . . . . . . . . . . . . . . . . . . 15
3.3 Parallel Transport and Geodesics . . . . . . . . . . . . . . . . 18

4 Basics of General Relativity 19


4.1 The Equivalence Principle . . . . . . . . . . . . . . . . . . . . 19
4.2 The Principle of General Covariance . . . . . . . . . . . . . . 20
4.3 Electrodynamics in General Relativity . . . . . . . . . . . . . 21
4.4 Newtonian Limit of the Geodesic Equation . . . . . . . . . . 22
4.5 Einstein’s Field Equations . . . . . . . . . . . . . . . . . . . . 24

5 The Reissner-Nordström Metric 25


5.1 Gravitational Time Dilation and Redshift . . . . . . . . . . . 32
5.2 The Geodesic Equation . . . . . . . . . . . . . . . . . . . . . . 34
5.2.1 Comparison to Newtonian Mechanics . . . . . . . . . 37
5.2.2 Circular Orbits of Photons . . . . . . . . . . . . . . . . 38
5.3 Motion of a Charged Particle . . . . . . . . . . . . . . . . . . 38
5.4 Event Horizons and Black Holes . . . . . . . . . . . . . . . . 40

6 Summary and Conclusion 44

2
1 Introduction
In 1915 Einstein completed his general theory of relativity. It did not take
long before the first non-trivial exact solution for the Einstein field equa-
tions was found by Karl Schwarzschild in 1916 which corresponds to the
gravitational field of a spherically symmetric object [1]. In the same year
Hans Reissner generalized Schwarzschild’s solution to include an electri-
cally charged object [2]. Gunnar Nordström later (independently of Reiss-
ner) arrived at the same solution [3] now known as the Reissner-Nordström
metric.
The Reissner-Nordström metric is a famous solution to the Einstein
field equations. It describes the spacetime geometry around a spherically
non-rotating charged body. The universe at large appear to be electrically
neutral, so it is highly unlikely to find a macroscopic object that possess
a considerable amount of net charge. The Reissner-Nordström solution is
therefore not relevant to realistic situations in astrophysics. It does however
contribute in understanding the fundamental nature of space and time.
Also, the Reissner-Nordström metric is a more general solution than the
Schwarzschild metric, so by simply putting the electrical charge to zero we
obtain the Schwarzschild solution (which has plenty of practical applica-
tions).
In this paper we will first give a short introduction to general relativ-
ity. Our main goal is then to present a detailed derivation of the Reissner-
Nordström metric (which is often overlooked in many textbooks) without
assuming a static spacetime. We then proceed to study some of the physics
in this spacetime such as:

• Gravitational time dilation and redshift

• Equations of motion for non-charged particles (massive or massless)


derived with the geodesic equation

• Equations of motion for charged massive particles derived with the


Lagrangian formalism

• Event horizons and black holes

2 Review of Special Relativity


In this chapter we give an overview of special relativity. We do this since
general relativity generalizes special relativity, giving a description of grav-
ity as a geometric effect of space and time. Special relativity is a theory
regarding the relationship between time and space. It is based on two pos-
tulates:

3
(1) The laws of physics take the same form in all inertial reference frames.

(2) The speed of light is the same for all observers.

With these assumptions it becomes necessary to replace the Galilean trans-


formations of classical mechanics with the Lorentz transformations.
The Lorentz transformations predicts that events that occur at the same
time for one observer does not occur at the same time for an observer that
is moving relative to the first one. This means that the absolute time and
space that is used in the Galilean transformations must be abandoned and
we instead describe time and space as part of the same continuum know as
spacetime. Other predictions is that moving object will be shortened, mov-
ing clocks run slower and that addition of velocities is not as simple as with
Galilean transformations.
An event is an occurrence that is characterised by a definite time and
location relative to a reference frame. This means that an event can be
thought of as a point in spacetime. Consider an event in an inertial ref-
erence frame S that is given by the coordinates (ct, x, y, z), where c is the
speed of light. Suppose now that there is another inertial reference frame
S 0 with the spatial coordinate axes orientated as in S but is moving with
constant velocity v relative to S in the x-direction. Let also the origins co-
incide at time zero. The coordinates of the event in S 0 is defined to be
(ct0 , x0 , y 0 , z 0 ) . The Lorentz transformations, which can easily be derived if
one assumes that the transformations are linear, specifies that these coordi-
nates have the relation
 vx 
t0 = γ t − 2
c
0
x = γ (x − vt)
y0 = y
z0 = z

where
1
γ≡q .
v2
1− c2

These transformations of course reduce to the Galilean transformations if


v  c, since then γ ≈ 1 and t  vx/c2 .
In Euclidian space the distance between two points is invariant. In
spacetime, the "distance" s that is invariant is

s2 = c2 (∆t)2 − (∆x)2 − (∆y)2 − (∆z)2 .

4
If we look at an infinitesimal spacetime interval (which of course also is an
invariant) we have
ds2 = c2 dt2 − dx2 − dy 2 − dz 2 .
Note that it is also possible to choose the interval to be
ds2 = −c2 dt2 + dx2 + dy 2 + dz 2 .
The choice of sign is arbitrary and has no physical implication as long as
one is consistent. Throughout this paper we will use the sign convention of
the former infinitesimal interval, i.e. (+, −, −, −).
If we let (x0 , x1 , x2 , x3 ) ≡ (ct, x, y, z) we can write the infinitesimal in-
terval as
X3 X 3
ds2 = ηαβ dxα dxβ ,
α=0 β=0

where ηαβ is the metric tensor. When we use a Cartesian coordinate system,
the metric tensor is given by
 
1 0 0 0
0 −1 0
 
0
ηαβ =  ,
0 0 −1 0 

0 0 0 −1
and is known as the Minkowski metric. If we also use Einstein’s summation
convention, which implies summation over an index that appears both as
a subscript and a superscript, we can write the infinitesimal interval as
3 X
X 3
2
ds = ηαβ dxα dxβ ≡ ηαβ dxα dxβ . (2.1)
α=0 β=0

Consider now again the Lorentz transformations. They can be written


as
x0α = Λαµ xµ ,
where Λαµ is constant with the condition
ηαβ = Λµα Λν β ηµν .
The coordinate differentials transforms as
dx0α = Λαµ dxµ ,
and with the chain rule we can write that the coordinates transforms ac-
cording to
∂x0α µ
x0α = Λαµ xµ = x .
∂xµ

5
2.1 4-vectors
In special relativity a 4-vector is an object with four components that trans-
form in a specific way under Lorentz transformations. More specifically,
the components of a contravariant 4-vector V α transforms according to

V α = Λαµ V µ

while the components of a covariant vector Vα transforms as

Vα = Λαµ Vµ .

Here we use a superscript to denote a contravariant index while a subscript


denotes a covariant index.
With the metric tensor one can lower and raise indices. That is, we can
change a contravariant vector to a covariant and vice versa:

V α = η αµ Vµ
Vα = ηαµ V µ .

Here η αµ is the inverse of ηαµ and together they satisfy

η αµ ηµβ = δβα ,

where δβα is the usual Kronecker delta defined by


(
1 if α = β
δβα =
0 if α 6= β

For example, given the contravariant vector V α = (V 0 , V 1 , V 2 , V 3 ) the cor-


responding covariant vector is

Vα = (V0 , V1 , V2 , V3 ) = (V 0 , −V 1 , −V 2 , −V 3 ).

Raising and lowering indices does not only apply to vectors, but to tensors
of any rank.
With the metric tensor we can define an invariant scalar product for
4-vectors according to

V 2 = ηαβ V α V β = η αβ Vα Vβ = V α Vα .

In Euclidean space the scalar product of vectors is always nonzero. But


scalar products of 4-vectors in spacetime can be be either positive, zero or
negative. When V α Vα > 0 we call it a timelike vector. When V α Vα = 0 it is
called a null vector. And lastly, when V α Vα < 0 it is called a spacelike vector.

6
Consider now a moving particle. The proper time τ (i.e. the time mea-
sured by a clock following the particle) is independent of the coordinate
system used. The infinitesimal interval is given by

ds 1p 2
dτ = = dt − dx2 − dy 2 − dz 2
cs c
 2  2  2 r
1 dx dy dz v2 dt
= 1− − − = dt 1 − 2 = .
c dt dt dt c γ

Let xα = xα (τ ) = (ct, r) be the trajectory in spacetime of the particle, where


r is the position 3-vector given by r = (x, y, z). The 4-velocity uα of the
particle is defined by

dxα
 
α dt dt dr
u ≡ = c , = γ(c, v),
dτ dτ dτ dt

where v ≡ dr/dt is the 3-velocity. In Newtonian mechanics the momen-


tum mv is conserved. In relativistic mechanics it turns out that it is the
3-momentum defined by p ≡ γmv that is conserved. The 4-momentum is
then defined by

pα ≡ muα = mγ(c, v) = (mcγ , p),

where m is the rest mass of the particle. Let F ≡ dp/dt be the 3-force. The
kinetic energy of the particle is then given by
Z Z Z Z
dp
Ek = F · dx = · dx = v · dp = v · d(mγv).
dt
Using integration by parts we obtain
Z Z
2 2 m
Ek = mγv − mγv · dv = mγv − γd(v 2 )
2
mc2
= mγv 2 + − E0 = γmc2 − E0 ,
γ

where E0 is a constant of integration. By putting v = 0 so that Ek = 0 we


can determine E0 to be E0 = mc2 . E0 is interpreted as the "rest energy"
while γmc2 is interpreted as the total energy of the particle. This is an
interesting result as it indicates that mass itself has energy content. This is
famously known as the mass-energy equivalence.
Note that the first component in the 4-momentum was γmc, so if E is
the total energy of the particle we can write the 4-momentum as

pα = (E/c , p).

7
Finally we will derive the so-called relativistic energy-momentum relation.
Since the inner product of 4-vectors is invariant, we must have that

pα pα = p0α p0α .

We can take the primed coordinate system to be an instantaneous rest frame.


In that case we have p0α = (mc, 0) while pα = (E/c, p). With the notation
p = |p|, the above can then simply be rewritten as

E 2 = p2 c2 + m2 c4 ,

which is the energy-momentum relation.

2.2 Electrodynamics in Special Relativity


Maxwell’s equations describe the generation and interaction of electric and
magnetic fields with each other and by charges and currents. With the
electric field E, magnetic field B, charge density ρ and current density J,
Maxwell’s equations are
ρ
∇·E= (2.2)
0
∇·B=0 (2.3)
∂B
∇×E=− (2.4)
∂t
∂E
∇ × B = µ0 J + µ0 0 , (2.5)
∂t
where 0 is the vacuum permittivity and µ0 is the vacuum permeability that
satisfy
1
c2 = .
µ0 0
If a particle has an electrical charge q, the electric and magnetic field will
exert a force (the Lorentz force) given by

F = q (E + v × B) .

The conservation of charge can be written as

∂ρ
+ ∇ · J = 0. (2.6)
∂t
This can easily be seen by considering a volume V with charge q. The net
current that flows into V is
ZZ ZZZ
dq
I= =− J · dS = − (∇ · J)dV,
dt
∂V V

8
where we used the divergence theorem in the last step. The derivative of q
can on the other hand be written as
ZZZ ZZZ
dq d ∂ρ
= ρdV = dV.
dt dt ∂t
V V

Comparing the two above equations we obtain equation (2.6).


The electric and magnetic fields can be expressed in terms of an elec-
tric scalar potential Φ and a magnetic vector potential A. The electric and
magnetic fields in terms of these potentials are given by

∂A
E = −∇Φ − (2.7)
∂t
B = ∇ × A. (2.8)

Maxwell’s equations can now be formulated as

∂ ρ
∇2 Φ +
(∇ · A) = − (2.9)
∂t 0

1 ∂2A
   
1 ∂Φ
∇2 A − 2 2 − ∇ ∇ · A + 2 = −µ0 J. (2.10)
c ∂t c ∂t
Though these equations look a bit more complicated, we have reduced the
number of components to solve for from 6 to 4. Also, the potentials are
not physically meaningful quantities that can be measured; the electric and
magnetic field are. In other words, the potentials are not uniquely deter-
mined by Maxwell’s equations. We can therefore make the replacements

A → A + ∇ψ
Φ → Φ − ∂ψ/∂t,

and it will not affect the electric and magnetic field. For example, the elec-
tric field will not change since

∂A0
   
0 0 ∂ψ ∂A ∂
E = −∇φ − = − ∇φ − ∇ − + ∇ψ
∂t ∂t ∂t ∂t
∂A
= −∇φ − = E.
∂t
To be able to make such replacements is know as gauge freedom, and these
transformations are known as gauge transformations. An often used gauge
is the so-called Lorenz gauge, which allows one to choose A and Φ so that
they satisfy the condition

1 ∂Φ
∇·A+ = 0.
c2 ∂t

9
Using the d’Alembertian operator defined by  ≡ ∇2 − c−2 ∂ 2 /∂t2 and the
Lorenz gauge, equation (2.9) and (2.10) can now be written as

 Φ = −ρ/0
 A = −µ0 J.

It is possible to write Maxwell’s equations on a simple form (known as


the covariant formulation of electromagnetism) that will be useful later. We do
this by defining the 4-potential Aα , the electromagnetic field tensor F αβ and
the 4-current j α by
 
α Φ
A ≡ ,A (2.11)
c
F αβ ≡ Aβ,α − Aα,β (2.12)
α
j ≡ (cρ, J), (2.13)

where we have used the contravariant 4-gradient defined by


∂ β
Aβ,α ≡ ∂ α Aβ ≡ A ,
∂xα
while the covariant 4-gradient is ∂α = gαµ ∂ µ . With the 4-potential the
Lorenz gauge condition can now simply be written as

∂α Aα = 0.

The electromagnetic field tensor is antisymmetric (F αβ = −F βα ), and can


in matrix form be written as
 
0 E1 /c E2 /c E3 /c
−E1 /c −B3 B2 
 
0
Fαβ = −E /c B
.
 2 3 0 −B1  
−E3 /c −B2 B1 0

It is straightforward to show that Maxwell’s equations can be written as

F αβ,β = −µ0 j α (2.14)


Fαβ,γ + Fβγ,α + Fγα,β = 0. (2.15)

The first one of these corresponds to equations (2.2) and (2.5) while the
second one is equivalent to (2.3) and (2.4).
The equation for charge conservation can also be written in a compact
way. We simply do this by taking the 4-divergence of the 4-current and set
it equal to zero:
 
1∂ ∂ρ
j α,α = , ∇ · (cρ, J) = + ∇ · J = 0.
c ∂t ∂t

10
This is exactly the same as equation (2.6).
Another object, which will be very central in general relativity, is the
electromagnetic stress-energy tensor given by [6]
 
αβ 1 1 αβ µν αµ β
T = η F Fµν − F F µ , (2.16)
µ0 4
which describes the flow of electromagnetic energy and momentum in space-
time. Properties for the stress-energy tensor is that it is symmetric (T αβ =
T βα ) and traceless:
 
α αβ 1 1 αβ µν αµ β
T ≡ T α = ηαβ T = ηαβ η F Fµν − ηαβ F F µ (2.17)
µ0 4
 
1 1 αβ µν αµ
= ηαβ η F Fµν − F Fαµ = 0, (2.18)
µ0 4

since ηαβ η αβ = η αα = 4.

3 Tensor Fields and Manifolds


In this chapter we review the foundations of the mathematical formulation
that is used in general relativity. In general relativity spacetime is being
treated as a 4-dimensional pseudo-Riemannian differentiable manifold, so
here we go through these concepts.
An n-dimensional manifold is a set of points that resembles an n-dimen-
sional Euclidean space near each point. Furthermore, points can be labeled
by a system of n real-valued coordinates xα = (x1 , x2 , . . . , xn ) such that
there is a one-to-one correspondence between the points and the labels.
The manifold does not need to be completely "covered" by one specific co-
ordinate system. We can describe the manifold by a collection of coordinate
systems where each coordinate system covers a subset of the manifold. At
points on the manifold that are covered by two different coordinate system
there is a set of equations that relates the coordinates of one system to the
coordinates of the other. Suppose there is another coordinate system with
the coordinates x0α . At points on the manifold where x0α and xα overlap,
each of these primed coordinates will be related by some function of the
the non-primed coordinate. We can write this as

x0α = x0α (x1 , x2 , . . . , xn ).

Similarly, we can write each xα as a function of the primed coordinates:

xα = xα (x01 , x02 , . . . , x0n ).

Note: in this section we let greek indices have the range 1, 2, . . . , n. In later
sections where we apply this to general relativity we will let greek indices

11
have the range 0, 1, . . . , n − 1, where the 0-component corresponds to the
time-component.
A differentiable manifold is a specific type of manifold. It essentially
just means that we are able to do differential calculus on the manifold.
More specifically, the partial derivatives ∂x0α /∂xµ and ∂xα /∂x0µ exist. These
partial derivatives now specify the transformation between the two coor-
dinate systems:
∂x0α µ ∂xα 0µ
x0α = x , x α
= x .
∂xµ ∂x0µ
From now on we will only work with differentiable manifolds and we can
therefore use the two synonymously.
A contravariant vector transform as
∂x0α µ
V 0α = V . (3.1)
∂xµ
while a covariant vector transform as
∂xµ
Vα0 = Vµ . (3.2)
∂x0α
From the transformation rules of contravariant and covariant vectors we
can generalize this to a tensor of any rank. Consider the direct product
T αβ ≡ V α Uβ of one contravariant vector and one covariant vector. This
object will transform as
 0α   ν 
0α ∂x µ ∂x ∂x0α ∂xν µ
T β = V Uν = T . (3.3)
∂xµ ∂x0β ∂xµ ∂x0β ν

It is now easy to see the generalization; a tensor T βα11βα22...β


...αm
n
will transform
as
∂x0α1 ∂x0α2 ∂x0αm ∂xν1 ∂xν2 ∂xνn µ1 µ2 ...µn
Tβ0α1 β1 α2 ...β
2 ...αn
= · · · · · · T .
m ∂xµ1 ∂xµ2 ∂xµm ∂x0β1 ∂x0β2 ∂x0βn ν1 ν2 ...νm
In special relativity the line element is given by ds2 = ηαβ dxα dxβ . We
will now use a more general (but still symmetric) metric tensor gαβ that will
give the line element for the manifold:

ds2 = gαβ dxα dxβ .

We define the inverse metric tensor g αβ such that

g αβ gβγ = g αγ = δγα .

It follows directly from the symmetry of gαβ that g αβ also is symmetric.


The signature is defined as the number of positive, negative and zero
eigenvalues of the metric tensor gαβ . For example, the Minkowski metric

12
will have the signature (+, −, −, −). A manifold with a metric with only
positive signs is called Riemannian. A generalization of the Riemannian
manifold is the pseudo-Riemannian manifold that do not need to have a sig-
nature with only positive signs. The signature used in describing spacetime
is clearly not of only positive signs, and is therefore a pseudo-Riemannian
manifold.
The metric tensor can be used to raise and lower indices:
gαβ T µβν = T µαν . (3.4)
A tensor with only up-indices is called contravariant while a tensor with
only down-indices is called covariant. If it has at least one up and one
down index it is called mixed. Furthermore, a tensor with m up indices
and n down indices will be denoted as an (m, n)-tensor.
A summation over an upper and lower index is called a contraction and
will yield a new tensor of lower rank. For example, consider a tensor T αβ γ .
If we do a contraction of α with β we get a new tensor
Rγ ≡ T αα γ = gαβ T αβγ .
It is straightforward to show that Rγ satisfy the transformation law
∂x0γ µ
R0γ = R .
∂xµ
The sum or difference of two tensors is a tensor of the same type. This
can easily be seen if we look at a specific example of a sum between two
second rank tensors. Consider the object Rαβ ≡ S αβ + T αβ , where S and T
are tensors. It will transform as
∂xα ∂x0ν 0 µ 0µ ∂xα ∂x0ν 0 µ
Rαβ = S αβ + T αβ =

S ν + T ν ≡ R ν.
∂x0µ ∂xβ ∂x0µ ∂xβ
It is then obvious that this generalizes to a sum or difference between two
tensors of any rank. It is now easy to convince oneself that any equation be-
tween two tensors with the same upper and lower indices will be invariant
under coordinate transformations.

3.1 Covariant Differentiation and Christoffel Symbols


In general, the differentiation of a tensor does not yield another tensor. This
can easily be seen by considering the contravariant vector V α which trans-
forms as
∂x0α µ
V 0α = V .
∂xµ
Differentiating w.r.t. x0β we get
∂V 0α ∂x0α ∂xν ∂V µ ∂ 2 x0α ∂xν µ
= + V .
∂x0β ∂xµ ∂x0β ∂xν ∂xµ ∂ν ∂x0β

13
Because of the second term on the right hand side, ∂V α /∂xβ does not trans-
form as a tensor. A derivative that do yield another tensor when operating
on a tensor is the covariant derivative, denoted by ∇µ (or by use of a semi-
colon). The covariant derivative of a contravariant vector V α is defined
by
∇β V α ≡ V α;β ≡ ∂β V α + Γαβµ V µ , (3.5)
where Γαβµ is the Christoffel symbols (also known as affine connections) which
can be thought of as a "correction" term. We claimed that ∇β V α is a tensor,
so it must transform as
∂xµ ∂x0α
∇0β V 0α = ∇µ V ν .
∂x0β ∂xν
From this we can work out how the Christoffel symbols must transform.
The both sides in the above equation can be expanded with the use of equa-
tion (3.5), then we transform V 0 to V :

LHS : ∇0β V 0α = ∂β0 V 0α + Γ0α


βλ V

∂xµ ∂x0α ν ∂xµ ∂ 2 x0α ν 0α ∂x



= ∂ µ V + V + Γ βλ Vρ
∂x0β ∂xν ∂x0β ∂xµ ∂xν ∂xρ

∂xµ ∂x0α ν ∂xµ ∂x0α ν ∂xµ ∂x0α ν ρ


RHS : ∇µ V = ∂µ V + Γ V .
∂x0β ∂xν ∂x0β ∂xν ∂x0β ∂xν µρ
By comparing the LHS to the RHS one can reach the conclusion that the
Christoffel symbols must transform as

∂xµ ∂xρ ∂x0α ν ∂xµ ∂xρ ∂ 2 x0α


Γ0α
βλ = Γ − .
∂x0β ∂x0λ ∂xν µρ ∂x0β ∂x0λ ∂xµ ∂xρ
From this transformation law we see that the affine connection is not a ten-
sor. With similar reasoning one can show that the covariant derivative of a
covariant vector is [8]

∇β Vα ≡ Vα;β = ∂β Vα − Γµαβ Vµ ,

and that for second order tensors it is

(2, 0)-tensor: ∇γ T αβ = ∂γ T αβ + Γαµγ T µβ + Γβµγ T αµ (3.6)


(1, 1)-tensor: ∇γ T αβ= + Γαµγ T µβ − Γµβγ T αµ
∂γ T αβ (3.7)
(0, 2)-tensor: ∇γ Tαβ = ∂γ Tαβ − Γµαγ Tµβ − Γµβγ Tαµ . (3.8)

The covariant derivative is clearly a generalization of the partial deriva-


tive. One important distinction however is that the order of which covari-
ant differentiations are done does matter (i.e. it does not commute). So in

14
general, changing the order of covariant differentiation changes the result.
Properties such as linearity

∇(T + S) = ∇T + ∇S

and product rule

∇(T ⊗ S) = (∇T ) ⊗ S + T ⊗ (∇S)

still holds.
If one require that the affine connection is torsion free, i.e. that it is sym-
metrical in its lower indices (Γαβγ = Γαγβ ), and that the covariant derivative
of the metric tensor is zero everywhere (a property called metric compati-
bility), then the affine connection is unique [8]. Properties that directly fol-
lows from this is that the covariant derivative of the inverse metric is zero
(∇γ g αβ = 0) and that it commutes with rasing and lowering indices:

gαβ (∇γ V β ) = ∇γ (gαβ V β ) = ∇γ Vα .

We can now derive an expression for the Christoffel symbols in terms of


the metric tensor and its first derivatives. Consider the explicit expression
of the covariant derivative of the metric tensor computed with equation
(3.8):
0 = ∇γ gαβ = ∂γ gαβ − Γλγα gλβ − Γλγβ gλα .
By doing three different permutations of the free indices and combining
these equations one end up with

0 = ∂γ gαβ − ∂α gβγ − ∂β gγα + 2Γλαβ gλγ .

Solving for the Christoffel symbol finally yields

1
Γλαβ = g λµ (∂α gβµ + ∂β gµα − ∂µ gαβ ). (3.9)
2

3.2 Riemann Tensor


Consider the covariant derivative of a covariant vector:

∇β Vα = ∂β Vα − Γµαβ Vµ .

Since ∇β Vα is a (0, 2)-tensor, according to equation (3.8), a second differen-


tiation yields

∇γ (∇β Vα ) = ∂γ (∇β Vα ) − Γµαγ (∇β Vµ ) − Γµβγ (∇µ Vα )

15
The three terms in the above expression can be written as
 
∂γ (∇β Vα ) = ∂γ ∂β Vα − ∂γ Γµαβ Vµ − Γµαβ (∂γ Vµ )
Γµαγ (∇β Vµ ) = Γµαγ ∂β Vµ − Γνµβ Vν


Γµβγ (∇µ Vα ) = Γµαγ ∂β Vµ − Γνµβ Vν ) .


Taking the differentiations in different order (i.e. interchanging β and γ)
one can show that
 
∇γ (∇β Vα ) − ∇β (∇γ Vα ) = Vµ ∂β Γµαγ − ∂γ Γµαβ + Γναγ Γµνβ − Γναβ Γµνγ .
(3.10)
This leads us to make the definition
Rµαβγ ≡ ∂β Γµαγ − ∂γ Γµαβ + Γναγ Γµνβ − Γναβ Γµνγ . (3.11)
Rµαβγ is called the Riemann curvature tensor (or simply Riemann tensor). That
it really is a tensor can be understood by noting that the left hand side
of equation (3.10) is a difference between two tensors and therefore is a
tensor itself. Then the right hand side, Vµ Rµαβγ , must of course also be a
tensor. But since Vµ and Rµαβγ are completely independent of each other
we conclude that Rµαβγ is a tensor.
Since the Christoffel symbols are constructed from the metric tensor and
its first derivatives, the Riemann tensor is constructed from the metric ten-
sor and its first and second derivatives. In fact, it turns out that the Riemann
tensor is the only tensor that can be constructed from the metric tensor and
its first and second derivatives [5].
If the Riemann tensor vanishes at each point on a manifold we can now
say that the order of differentiations for a (0, 1)-tensor field does not mat-
ter. For other types of tensor fields we can make similar calculations as
when deriving equation (3.10). This lets us state that the order of covari-
ant differentiations of a tensor field of any rank and type does not matter
if Rµαβγ = 0 at all points on the manifold. We can now also give a pre-
cise definition of curvature. If Rµαβγ = 0 at each point on a manifold, the
manifold is flat. Otherwise it is curved.
One property of the Riemann tensor, known as the cyclic identity, is
Rµαβγ + Rµβγα + Rµγαβ = 0,
which can be showed in a straightforward but tedious way. If we lower the
upper index of the Riemann tensor, i.e. Rαβγδ ≡ gαµ Rµβγδ , one can read off
the following symmetrical properties [7]:
Rαβγδ = Rγδαβ
Rαβγδ = −Rβαγδ
Rαβγδ = −Rαβδγ
Rααγδ = 0.

16
In a similar way as we got equation (3.10), one can show that the fol-
lowing holds for a second rank covariant tensor:

(∇α ∇β − ∇β ∇α )Tγδ = Rµγβα Tµδ + Rµδβα Tγµ . (3.12)

From the definition of the Riemann tensor we had that

(∇γ ∇β − ∇β ∇γ )Vα = Rµαβγ Vµ .

With the use of the product rule and the above equation we obtain

∇λ (∇γ ∇β − ∇β ∇γ )Vα = Vµ ∇λ Rµαβγ + Rµαβγ ∇λ Vµ .

Using the above equation together with (3.12) (but with ∇γ Vβ instead of
Tγβ ) one can show that the Riemann tensor must satisfy the relation [9]

∇δ Rµαβγ + ∇β Rµαγδ + ∇γ Rµαδβ = 0, (3.13)

which is known as the Bianchi identity. In a n-dimensional space, the Rie-


mann tensor has n4 components. But with all the properties that theRie-
mann tensor possess, one can show that there only exist n2 n2 − 1 /12
independent components [8]. For example, in a 4-dimensional manifold
we have 20 independent components of the Riemann tensor.
From the Riemann tensor we can construct the Ricci tensor, which we
define by

Rαβ ≡ Rµαµβ = ∂µ Γµαβ − ∂β Γµαµ + Γµνµ Γναβ − Γµνβ Γναµ . (3.14)

From the cyclic identity of the Riemann tensor, it can be showed that the
Ricci tensor is symmetric. The trace of the Ricci tensor is known as the Ricci
scalar:
R ≡ Rαα = g αβ Rαβ . (3.15)
Consider now the Bianchi identity given by equation (3.13). If we con-
tract µ with β and use that Rαβγδ = −Rαβδγ we obtain

∇δ Rαγ + ∇β Rβαγδ − ∇γ Rαδ = 0.

Multiplying with g αγ and using that

g αγ ∇β Rβαγδ = ∇β Rβγ γδ = −∇β Rγβγδ


 
= −∇β g βα Rγ αγδ = −∇β Rβδ

we get
∇δ R − 2 ∇β Rβδ = 0.

17
The above equation can be written in an equivalent form as
 
µν 1 µν
∇µ R − g R = 0.
2
This expression leads us to define a new tensor, which is of great impor-
tance in general relativity, called the Einstein tensor:
1 αβ
Gαβ ≡ Rαβ − g R.
2
The Einstein tensor is symmetric (since both Rαβ and g αβ are symmetric)
and divergence-less.

3.3 Parallel Transport and Geodesics


Parallel transport can be thought of as transporting a tensor along a curve
on a manifold while keeping the tensor constant. For simplicity, consider
a vector. On a flat manifold one can parallel transport the vector along a
curve from one point to another and the result of this transport does not
depend on how the curve looks like. This is very easy to see if one uses
Cartesian coordinates. Then we can parallel transport a vector by simply
keeping the components constant.
In a curved manifold it is not that simple anymore. The easiest way to
understand this is probably to consider parallel transport of a vector on a 2-
sphere. Imagine a vector at one point on the sphere. If one move this vector
around a closed curve while letting it point in the "same direction" it will
not be parallel to the "original" vector when it comes back to the starting
point.
What this really means is that there is no well-defined way to globally
say that two vectors are parallel or not. We can only compare two vectors
if they are at the same point (i.e. they are elements of the same tangent
space).
Now we will consider how to mathematically describe pararell trans-
port. Consider first a flat manifold where a curve is given by xµ (λ), where λ
is a parameter. On this flat manifold the requirement that a tensor T α1 ...αmβ1 ...βn
is constant along this curve is simply (with the chain rule)
d α1 ...αm dxµ
0= T β1 ...βn = ∂µ T α1 ...αmβ1 ...βn ,
dλ dλ
for all λ. This will obviously not hold for a curved manifold. The gener-
alization of this is to simply change the partial derivative by a covariant
derivative. So a tensor T α1 ...αmβ1 ...βn is said to be parallel transported along
the curve xµ (λ) if it for all λ satisfy
dxµ
0= ∇µ T α1 ...αmβ1 ...βn .

18
We will only consider the case when the tensor is a vector. For a contravari-
ant vector V α , the above equation reads

dxµ dxµ  dV α dxµ ν


0= ∇µ V α = ∂µ V α + Γαµν V ν = + Γαµν V . (3.16)
dλ dλ dλ dλ
We have now a clear definition of parallel transport. Next step is to talk
about geodesics, which is a central part of general relativity. A geodesic can
be thought of as the generalization of a straight line to curved spaces. We
will define a geodesic to be a curve that parallel transport its own tangent
vector. If one uses the Christoffel connection (as we do), another equiva-
lent definition is to say that a geodesic is the path of the shortest distance
between two points (which clearly is a generalization of a straight line) [8].
As before, let a curve be given by xα (λ). The tangent vector along this
curve is simply dxα /dλ. The condition that this tangent vector is parallel
transported along the curve is given by (3.16). So if xα is to be a geodesic it
must satisfy
d2 xα µ
α dx dx
ν
0= + Γ µν .
dλ2 dλ dλ
This is called the geodesic equation. It is a nonlinear second order differential
α
equation. If one has has an initial position xα (λ0 ) and direction dx dλ λ0 it

will give a unique geodesic.

4 Basics of General Relativity


With the basics of the mathematical foundation of general relativity de-
scribed in the previous chapter, we now use it to discuss the physics of
the theory. We start with the equivalence principle which is an important
principle for generalizing the physics of special relativity to include grav-
ity, and then introduce the principle of general covariance (which let us
easily generalize electrodynamics to curved spacetime). Finally we show
how motion in general relativity reduces to Newtonian mechanics and in-
troduce the Einstein field equations.

4.1 The Equivalence Principle


The (strong) equivalence principle can be stated as:

At every spacetime point in a gravitational field it is possible to choose


a locally inertial coordinate system such that, within a sufficiently small
region of the point, the laws of nature are the same as in special relativ-
ity (i.e. non-accelerated coordinate system in absence of gravitation).

19
There is great similarity between the equivalence principle and that a curved
Riemannian manifold appears locally flat. Because of this resemblance one
may expect that spacetime in general relativity can be described with a
pseudo-Riemannian manifold.
In a locally inertial Cartesian coordinate system with coordinates Lα the
metric is given by
ds2 = ηαβ dLα dLβ .
By using that
∂Lα µ
dLα = dx
∂xµ
for any change to the arbitrary coordinates xα , we have that

ds2 = g̃αβ dxα dxβ

where we defined
∂Lµ ∂Lν
g̃αβ ≡ ηµν .
∂xα ∂xβ
Locally in the coordinate system of Lα the equations of motion of a free
particle is
d2 Lα
= 0,
dλ2
where λ is a parameter (for massive particles this parameter can be taken to
be the proper time, but not for massless particles such as photons). Chang-
ing to the coordinates xµ and using the chain rule we can write the equa-
tions of motion as
d2 xα dxµ dxν
2
+ Γ̃αµν = 0,
dλ dλ dλ
where we defined
∂xα ∂ 2 Lσ
Γ̃αµν = .
∂Lσ ∂xµ ∂xν
This has exactly the same form as the geodesic equation derived in section
3.3. In fact, one can show that g̃αβ and Γ̃αµν has the exact same relation to
each other as gαβ and Γαµν has (as derived in section 3.1) [5]. So in general
relativity, where we treat spacetime as a 4-dimensional pseudo-Riemannian
manifold, we can express the equations of motion geometrically. That is,
a particle in free fall will follow a path that is a geodesic, given by the
geodesic equation.

4.2 The Principle of General Covariance


We will now discuss a very useful way of generalizing results that are valid
in special relativity to be valid in general relativity known as the principle of
general covariance. This principle states that a physical equation holds true
in all coordinate systems if:

20
(1) The equation holds true in absence of gravitation (i.e it holds true in
special relativity).
(2) It is a tensor equation (i.e it preserves its form under a general coord-
inate transformation).

By the equivalence principle one can write down an equation that holds
in a locally inertial coordinate system and then make a general coordinate
transformation to find the corresponding equation in that coordinate sys-
tem. With the principle of general covariance finding the equations that
holds for all coordinate systems is much simpler. It can be seen that it
follows from the equivalence principle by considering any equation that
satisfy condition (1) and (2). Since the equation is generally covariant it
preserves its form under a general coordinate transformation, so if it is true
in any coordinate system it is true in all coordinate systems. The equiva-
lence principle tells us that at every point in spacetime there exists locally
inertial coordinate systems in which the effects of gravity are absent. Since
we assumed that our equation holds in special relativity (i.e. no gravity)
and therefore holds in these locally inertial systems, it must hold in all co-
ordinate systems.
Note that any equation can be made generally covariant by working out
what it looks like in arbitrary coordinate systems. So in it self the principle
of general covariance has no physical meaning.
Our method now to find equations that are valid in a general gravita-
tional field is by simply take the valid equations (and definitions) of special
relativity and replace partial derivatives (with respect to coordinates) by
covariant derivative and the Minkowski metric ηαβ by the general metric
tensor gαβ .

4.3 Electrodynamics in General Relativity


Recall that in special relativity, when no gravitational field is presence,
Maxwell’s equations can be written as

F αβ,β = −µ0 j α
Fαβ,γ + Fβγ,α + Fγα,β = 0

if one is using a Cartesian coordinate system. When in a gravitational


field these equations hold only for a locally inertial coordinate system at
a point. But according to the principle of general covariance, if we simply
change the partial derivatives to covariant derivatives the equations would
hold for any coordinate system. That is, in a general coordinate system

21
Maxwell’s equations are

F αβ;β = −µ0 j α
Fαβ;γ + Fβγ;α + Fγα;β = 0.

The electromagnetic tensor (when in a coordinate basis) can still be defined


by
Fαβ = Aβ,α − Aα,β
since the Christoffel symbols is symmetric in the lower indices. Instead of
raising and lowering indices with ηαβ , the more general metric tensor gαβ
is now used. In special relativity the conservation of charge is expressed
by j α,α = 0. In the same way as before, we change the derivative and this
generalizes to
j α;α = 0.
Note that this only corresponds to a local conservation law of electric charge
and not a global one. For the stress-energy tensor given by equation (2.16)
in special relativity we only need to swap from the Minkowski metric ten-
sor to the general metric tensor. The result is
 
αβ 1 1 αβ µν αµ β
T = g F Fµν − F F µ , (4.1)
µ0 4

while it still is symmetric T αβ = T βα and traceless:




T ≡ T αα = gαβ T αβ = 0.

4.4 Newtonian Limit of the Geodesic Equation


In the Newtonian theory of gravity the equation of motion for a particle in
free fall outside a spherically symmetric body with total mass M is
GM
a = −∇Φ = − r̂, (4.2)
r2
where Φ is the potential and r is the distance from the center of the body.
Newton’s theory does not hold in general relativity, but it should be re-
covered as a good approximation when the gravitational field is weak and
static while the particles are moving much slower than the speed of light.
Recall the geodesic equation, which determines the path taken by a particle
in free fall:
d2 xα µ
α dx dx
ν
+ Γ µν = 0.
d2 τ dτ dτ
For slow moving particles we have that

dx0 dt dxi dxi dt


=c  = .
dτ dτ dτ dt dτ

22
So the geodesic equation can be approximated with

d2 xα dt 2
 
α
+ Γ00 c = 0.
d2 τ dτ

For a static field we have that


1 1
Γα00 = g αµ (g0µ,0 + gµ0,0 − g00,µ ) = − g αµ g00,µ ,
2 2
since gαβ,0 = 0. Thus we have that

d2 xα 1 αµ dt 2
 
− g g00,µ c = 0. (4.3)
d2 τ 2 dτ

The weak field approximation allows us to write the metric tensor as the
Minkowski metric plus a small perturbation term:

gαβ = ηαβ + hαβ .

We will ignore all terms higher than first order. This means that all products
of h (or its derivatives) with h (or its derivatives) are neglected. With the
condition δγα = g αβ gβγ , we must have that

g αβ = η αβ − hαβ ,

where hαβ = η αµ η βν hµν . To first order, and from the constancy of the
Minkowski metric we see that

g αµ g00,µ = (η αµ − hαµ )(η00,µ + h00,µ ) = η αµ h00,µ

So equation (4.3) now becomes

d2 xα dt 2
 
1 αµ
= η h00,µ c .
dτ 2 2 dτ

Because η αµ is diagonal, we have that the case α = 0 implies that dt/dτ is


constant. We can therefore use

d2 xα d2 xα dt 2
 
= .
dτ 2 dt2 dτ

Only considering the spacelike components, the Minkowski tensor η αµ is


equal to −δ αµ and we have

d2 xi 1
2
= − c2 δ ij h00,j ,
dt 2

23
where i and j can be 1,2 or 3. But the left hand side of this equation is just
the usual components of the 3-acceleration. h00,j is just the gradient of h00 ,
so comparing with equation (4.2) we find that


h00 = ,
c2
and since g00 = η00 + h00 we conclude that
2Φ 2GM
g00 = 1 + 2
=1− 2 .
c c r
Thus the 00-component of the metric tensor seems to be closely related to
the Newtonian potential energy.

4.5 Einstein’s Field Equations


In 1915 Einstein formulated the equations that govern how spacetime is
being curved by matter and energy. These will be refereed to as Einstein’s
field equations and may be written as

8πG
Gαβ = Tαβ , (4.4)
c4
where G is Newton’s gravitational constant, Tαβ is the stress energy tensor
and Gαβ is the Einstein tensor defined by

1
Gαβ ≡ Rαβ − R gαβ . (4.5)
2
In total there are 4×4 = 16 equations. But since Gαβ (and Tαβ ) is symmetric
this reduces to 10 independent equations. As we found earlier in section
3.2, the Einstein tensor satisfy Gαβ ;α = 0, which reduces the number even
further to 10 − 4 = 6 independent equations.
Another useful form of Einstein’s field equations can be obtained if we
take the trace on both sides. The trace of Gαβ is
 
αβ αβ 1
g Gαβ = g Rαβ − R gαβ = R − 2R = −R
2

since g αβ gαβ = 4. If we define T ≡ T αα = g αβ Tαβ we get the relation

8πG
R=− T.
c4
This allows us to write Einstein’s field equations in the equivalent form:
 
8πG 1
Rαβ = 4 Tαβ − T gαβ . (4.6)
c 2

24
Einstein’s field equations together with the geodesic equations, which gov-
ern how particles only influenced by gravity move, build up the core of
general relativity.
It should be noted that Einstein’s field equations can not be derived
from any underlying principle. Nevertheless, one can make reasonable ar-
guments that this is a good candidate. One of these is that it reduces to the
Poisson equation
∇2 Φ = 4πGρ,
where Φ is the gravitational potential and ρ is the mass density when con-
sidering the 00-component in a weak-field approximation [7].

5 The Reissner-Nordström Metric


In this chapter we will first derive the Reissner-Nordström metric and then
describe some physics in this metric such as time dilation, gravitational
redshift, equations of motion for both a charged and non-charged particle
and lastly black holes and event horizons.
The Reissner-Nordström metric is a solution to Einstein’s field equa-
tions that describes the spacetime around a spherically symmetric non-
rotating body with mass M and an electric charge Q. Other than spheri-
cal symmetry we also have the assumption that the space is empty from
matter (there is only an electromagnetic field). When Q → 0 the metric
should approach the Schwarzschild metric. Another property the metric
should have is that the spacetime is asymptotically flat. In other words, as
the distance from the body approaches infinity, the metric must approach
the Minkowski metric.
Because of the spherical symmetry the most natural coordinate system to
use is of course the spherical coordinate system. In flat spacetime, the met-
ric when using spherical coordinates is

ds2 = c2 dt2 − dr2 − r2 dθ2 − r2 sin2 θ dφ2 .

For curved spacetime (but still spherically symmetric) this can be general-
ized to

ds2 = A(t, r)c2 dt2 − B(t, r)dr2 − r2 dθ2 − r2 sin2 θ dφ2 .

One could make the assumption that A and B is independent of time, but
we will keep the time-dependence for now as we will find later that A and
B must necessarily be independent of time.
Recall Einstein’s field equations:
 
8πG 1
Rαβ = 4 Tαβ − T gαβ .
c 2

25
Since we are working in vacuum (no matter) with an electromagnetic field
we will use the electromagnetic stress-energy tensor given by
 
1 1 µν νµ
Tαβ = gαβ Fµν F − gβν Fαµ F . (5.1)
µ0 4

Remember also that the electromagnetic stress-energy tensor was traceless,


which means that Einstein’s field equations can be written as

8πG
Rαβ = Tαβ .
c4
Finally we also need the source-free Maxwell equations given by

F αβ;β = 0 (5.2)
Fαβ;γ + Fβγ;α + Fγα;β = 0. (5.3)

The calculation of the Ricci tensor is done by first calculating the Christof-
fel symbols with

1
Γλαβ = g λµ (∂α gβµ + ∂β gµα − ∂µ gαβ ),
2
and then computing the Ricci tensor with

Rαβ = Rµαµβ = ∂µ Γµαβ − ∂β Γµαµ + Γµνµ Γναβ − Γµνβ Γναµ .

as they were derived in section 3.1 and 3.2, respectively. It is a straight-


forward but tedious task, so the calculation of both the Christoffel symbols
and the Ricci tensor was done with the help of a computer algebra software
(Maple 17). All the non-zero Christoffel symbols are given by


Γ000 =
2Ac
Ḃ Ḃ
Γ101 = Γ110 = , Γ011 =
2Bc 2Ac
A0 A0
Γ001 = Γ010 = , Γ100 =
2A 2B
1 B0
Γ212 = Γ221 = , Γ111 =
r 2B
1 r
Γ313 = Γ331 = , Γ122 =−
r B
r sin2 θ
Γ323 = Γ332 = cot θ, Γ133 =−
B
Γ233 = − sin θ cos θ

26
where a dot represent differentiation w.r.t. t, and a prime w.r.t. r. For the
Ricci tensor all non-zero components are:
!
A0 A0 B 0 A00 A0
 
B̈ Ḃ Ȧ Ḃ
R00 = − + + + − + − (5.4)
4B A B 2B Br 2Bc2 4Bc2 A B
!
A0 A0 B 0 A00 B0
 
B̈ Ḃ Ȧ Ḃ
R11 = + − + − − − (5.5)
4A A B 2A Br 2Ac2 4Ac2 A B
 0
B0

r A 1
R22 = − − − +1 (5.6)
2B A B B
R33 = R22 sin2 θ (5.7)

R01 = R10 = . (5.8)
Brc
This is as far as we can get in generalizing a spherically symmetric gravita-
tional field. To determine A and B any further we need to invoke Einstein’s
field equations, and that means we must specify the stress-energy tensor.
In our case the stress-energy tensor was given in terms of the metric tensor
and the electromagnetic tensor Fαβ . From the spherical symmetry we have
that the electric field can only have a radial component. Also, this radial
component must not depend on θ or φ, so we have that
Er = E1 = E1 (t, r) = c F01 = −c F10 .
All the other components are zero since there are no currents or magnetic
monopoles. In matrix form we have
 
0 Er /c 0 0
−Er /c 0 0 0
Fαβ = 
 0
.
0 0 0
0 0 0 0
The components of the stress-energy tensor can now be computed with
equation (5.1). Consider the first term in the parenthesis. Carrying out the
summation gives
1 1  1
gαβ Fµν F µν = gαβ Fµ0 F µ0 + Fµ1 F µ1 = gαβ F10 F 10 + F01 F 01

4 4 4
1 01
 1
= gαβ 2F01 F = gαβ F01 F 01 .
4 2
For the second term we get
gβν Fαµ F νµ = gβν Fα0 F ν0 + gβν Fα1 F ν1 = gβ1 Fα0 F 10 + gβ0 Fα1 F 01 ,
and we can write equation (5.1) as
 
1 1
Tαβ = gαβ F01 F 01 − gβ1 Fα0 F 10 − gβ0 Fα1 F 01 .
µ0 2

27
The components the stress-energy tensor can now easily be obtained. We
have:
 
1 1 01 01
T00 = g00 F01 F − g00 F01 F
µ0 2
1 1
=− g00 F01 F 01 = − A F01 F 01
2µ0 2µ0

 
1 1 01 01
T11 = g11 F01 F − g11 F01 F
µ0 2
1 1
=− g11 F01 F 01 = B F01 F 01
2µ0 2µ0

1 1 2
T22 = g22 F01 F 01 = − r F01 F 01 (5.9)
2µ0 2µ0

1
T33 = g33 F01 F 01 = T22 sin2 θ.
2µ0

All other (non-diagonal) components of the stress-energy tensor turns out


to be zero. Since T01 = 0 we also have that R01 = 0 and with equation (5.8)
we conclude that Ḃ = 0, which means that B can not depend on t.
Note that
T00 T11
+ = 0.
A B
This in turn implies that
 0
B0

R00 R11 1 A
0= + = + .
A B rB A B

From the above expression we obtain

A0 B 0 ∂
0= + = ln(AB).
A B ∂r
This means that the product AB must be constant with respect to r. We can
write this as
AB = f (t),
where f (t) is some function that does not depend on r. We will however
show later that f must equal unity, but we will keep it as it is for now. As
we now have the relation g00 = −f /g11 (since A = g00 and B = −g11 ), we
can easily show that

F01 = g00 g11 F 01 = −f F 01 .

28
We will now solve Maxwell’s equations. Equation (5.3) does not give us
any more information because of the non-existence of magnetic monopoles,
and it can be seen that it is directly satisfied by considering the case α = 0,
β = 1 and γ = 0:

F01;0 + F10;0 + F00;1 = F01;0 − F01;0 = 0.

Similarly for the other cases it can be showed that Fαβ;γ + Fβγ;α + Fγα;β is
identically zero. Equation (5.2) will however give us a more explicit form
of Er . Using equation (3.6) for the covariant derivative of a second rank
contravariant tensor, equation (5.2) reads

0 = F αβ;β = ∂β F αβ + Γαµβ F µβ + Γβµβ F αµ . (5.10)

For α = 1 the above becomes

0 = ∂0 F 10 + Γ1µβ F µβ + Γβµβ F 1µ .

The second term in the above equation vanish

Γ1µβ F µβ = Γ1µ0 F µ0 + Γ1µ1 F µ1 = Γ110 F 10 + Γ101 F 01 = 0,

since Γ101 = Γ110 = 0 (or since F 01 = −F 10 together with Γ101 = Γ110 ). The
third term also vanish

Γβµβ F 1µ = F 10 Γ000 + Γ101 + Γ202 + Γ303 = 0




since all the Christoffel symbols in the parenthesis equals zero. Thus we
simply end up with
0 = ∂0 F 10 .
This of course implies that F 10 , and therefore Er , must not depend on time.
That is, we have
Er = Er (r).
Using α = 0 in equation (5.10) we obtain

0 = ∂1 F 01 + Γ0µβ F µβ + Γβµβ F 0µ . (5.11)

Similarly as before the second term vanish but in this case the third term
does not. We have that

Γβµβ F 0µ = Γβ1β F 01 = F 01 Γ010 + Γ111 + Γ212 + Γ313



 0
B0

01 A 2 2
=F + + = F 01
2A 2B r r
since
A0 B0
   
1 ∂ 1 ∂
+ = ln(AB) = ln (f ) = 0.
2A 2B 2 ∂r 2 ∂r

29
Equation (5.11) now reads

∂ 01 2 01
0= F + F
∂r r
which is an ordinary first order differential equation with the (easily checked)
solution
const.
F 01 = ,
r2
which let us write
const.
Er = .
r2
By the Gauss’s flux theorem we can conclude that the constant must equal
Q/4π0 , and we have
Q
Er = .
4π0 r2
This is not an unfamiliar expression. It is really Coulomb’s law, although
one must of course remember that r is just the one of our chosen coordi-
nate and do not necessarily measures the "real" radial distance when in a
Reissner-Nordström spacetime.
We are now close to get the final form of the Reissner-Nordström metric.
We only need to get a more explicit form of A and B in terms of r. This can
now be done by considering one of Einstein’s field equations, namely

8πG
R22 = T22 .
c4
For the left hand side we have
 0
B0

r A 1 1 ∂
R22 = − − − +1=− (rA) + 1
2B A B B f ∂r

which one gets by substituting B = f /A and B 0 = −f A0 /A2 and using the


product rule. For the right hand side we use equation (5.9) and obtain

1 ∂ 1 8πG 1
− (rA) + 1 = r2 Er2 .
f ∂r f c4 2µ0 c2

With Er2 = Q2 /(4π0 r2 )2 this can be written as

∂ GQ2
(rA) = f − .
∂r 4πc6 µ0 20 r2

If we now integrate and use that c2 µ0 = 1/0 this becomes

C(t) GQ2
A=f+ + ,
r 4π0 c4 r2

30
where C(t) is a function that may depend on time. When Q = 0 the metric
must reduce to the Schwarzschild metric. And as we showed in section 4.4,
when gravity is weak (i.e. when r is large) the metric tensor component
g00 must approach 1 − 2GM/c2 r. So at this limit, if the geodesics of the
metric should agree with the motion of Newtonian gravity, we must have
that f = 1 (which implies that AB = 1) and that C(t) = −2GM/c2 ≡ −rs . It
is worth noting that from the relation AB = f (t) one
p could directly get rid
of f (t) by redefining the time coordinate as dt → f (t)dt.
The constant rs is commonly known as the Schwarzschild radius. If we
also define
2 GQ2
rQ ≡ ,
4π0 c4
A and B can finally be written as
2
rs rQ
A=1− + 2
r r
!−1
2
1 rs rQ
B= = 1− + 2 ,
A r r

and the metric tensor in matrix form as


  
2
rQ
rs
 1 − r + r2 0 0 0 
  2
−1 
rQ
 
gαβ = 
 0 − 1 − rrs + r2
0 0 .

 

 0 0 −r2 0 

0 0 0 −r2 sin2 θ

We now have the complete Reissner-Nordström metric derived from Ein-


stein’s field equations together with Maxwell’s equations. An interesting
note however; we did not include any magnetic monopoles in this deriva-
tion. There is no experimental evidence of the existence of magnetic monopoles
so it is only natural to leave them out. Still, there is no theoretical argu-
ments that they should not exist, so it can be pleasing to assume that they
exist and see how the metric would look like if they did.
Let P be the magnetic charge of the body. Besides the radial electric field
component we would now also have a radial magnetic field with magni-
tude Br . This magnetic field would go like [8]

P
Br ∝ ,
r2
which corresponds to the electromagnetic tensor component

F23 = r2 Br sin θ.

31
The diagonal components of the stress-energy tensor would also get one
extra term containing Br2 . The implication of this is that we could make
our "no magnetic charge"-solution to a "with magnetic charge"-solution by
simply replacing Q2 by Q2 + P 2 /c2 , if one uses SI-units where the magnetic
charge is measured in ampere·meters (A · m).
This discussion about magnetic monopoles was just a quick add-on to
the Reissner-Nordström metric but we will not use it any further. That is,
in the following sections we will use P = 0.

5.1 Gravitational Time Dilation and Redshift


Consider some fixed point in space. With constant r, θ and φ we have that
dr, dθ and dφ is zero and the metric becomes
!
2
2 rs rQ
ds = 1 − + 2 c2 dt2 .
r r

Using that ds2 = c2 dτ 2 we obtain


s
2
rs rQ
dτ = dt 1− + 2. (5.12)
r r

dt can be interpreted as an infinitesimal time interval measured by a ob-


server that is infinitely far away from a gravitational body while dτ is the
interval measured by an observer at a distance r from the center of the
body. Suppose that the quantity inside the square root in equation (5.12) is
positive and less than unity (this is the case for any physically real situation
as discussed briefly in section 5.4). We then have that dτ < dt. This means
that the far-away observer qwill measure the clock that is closer to the body
run slower by a factor of 1 − rs /r + rQ 2 /r 2 .

Suppose now that an electromagnetic wave is transmitted radially out-


wards (or inwards) from a point (r1 , θ, φ) to another point at (r2 , θ, φ). Let
t1 be the coordinate time of emission and t2 be the coordinate time of recep-
tion. The electromagnetic wave is traveling along a radial null geodesic, so
we can use 0 = dθ = dφ and the metric becomes

c2 2
0 = ds2 = A dt2 − dr .
A
Using λ as an parameter for the geodesic the above can be written as
 2  2
2 dt 2 dr
A =c .
dλ dλ

32
If we let λ1 correspond to t1 and λ2 to t2 , solving for dt/dλ and integrating
we obtain Z λ2
1 dr
t2 − t1 = c dλ.
λ1 A dλ

Let now t01 be the coordinate time when the electromagnetic wave has os-
cillated exactly one period after first emission at r = r1 . Similarly, let t02
be the coordinate time when the electromagnetic wave has oscillated one
period after receiving the signal at r = r2 . Since the integral above does not
depend on t we have that

t2 − t1 = t02 − t01 ,

or equivalently as

∆t2 ≡ t02 − t2 = ∆t1 ≡ t01 − t1 .

From this we see that the coordinate time period at the point of emission
and the coordinate time period at the point of reception are equal. A real
clock however measures the proper time, so with equation (5.12) the peri-
ods in proper time are
s
2
rQ
rs
∆τ1 = ∆t1 1 − + 2
r1 r1

and s
2
rQ
rs
∆τ2 = ∆t2 1− + 2.
r2 r2
The proper frequency of the electromagnetic wave is simply the reciprocal
of the proper period and since ∆t1 = ∆t2 we must have the relation
!1/2 !−1/2
2
rQ 2
rQ
rs rs
f2 = f1 1− + 2 1− + 2 , (5.13)
r2 r2 r1 r1

where f1 and f2 are the proper frequencies measured by an observer lo-


cated at r1 and r2 , respectively.
Suppose that both r1 and r2 are large enough so that the quantities in-
side the square roots are positive. If r1 < r2 (traveling radially outwards),
we have that
!1/2 !1/2
2
rQ rQ2
rs rs
1− + 2 < 1− + 2
r2 r2 r1 r1

which means that f2 will be smaller than f1 and the wave is said to be
redshifted. Letting r1 > r2 (traveling radially inwards) the result would
instead be that f2 is larger than f1 and the wave is said to be blueshifted.

33
5.2 The Geodesic Equation
We will now look at the equations that describes the motion of a massive
non-charged freely falling particle or a photon in a Reissner-Nordström
spacetime. Both the particle and the photon will follow a geodesic, so find-
ing out the path they take is the same as solving the geodesic equation. The
massive particle will follow a timelike geodesic while the photon follows a
null geodesic. Note: in this and the following sections we put c = 1. We do
this because it let us skip a lot of tedious work remembering to use c when
dealing with the t-component.
Let xα = xα (λ) be a curve parameterized by λ. For xα to be a geodesic
it must satisfy the geodesic equation given by
d2 xα µ
α dx dx
ν
+ Γ µν = 0. (5.14)
d2 λ dλ dλ
For the massive particle it would be natural to have the geodesic parame-
terized by the proper time, and we will actually do this later. But since the
photon follows a null geodesic it can not be parameterized by the proper
time so we will stick with the more general parameter λ so that we can treat
both the massive particle and the photon as far as possible.
Using all the non-zero Christoffel symbols written down in the begin-
ning of this chapter we have that for α = 0, equation (5.14) reads
d2 t A0 dt dr
+ = 0. (5.15)
dλ2 A dλ dλ
Considering α = 1 we obtain

d2 r A0 dt 2 B 0 dr 2
   
0= 2 + +
dλ 2B dλ 2B dλ
 2
r sin θ dφ 2
2  
r dθ
− − . (5.16)
B dλ B dλ
For α = 2 and α = 3 we get
 2
d2 θ 2 dθ dr dφ
0= 2 + − sin θ cos θ (5.17)
dλ r dλ dλ dλ
2
d φ 2 dφ dr dφ dθ
0= 2 + + 2 cot θ . (5.18)
dλ r dλ dλ dλ dλ
By the spherical symmetry we must of course have that the trajectory is
in a plane. We can therefore, without any loss of generality, put θ = π/2.
This means that the derivatives of θ vanish and equation (5.17) is instantly
satisfied. With this simplification equation (5.16) and (5.18) now become

d2 r A0 dt 2 B 0 dr 2 r dφ 2
     
+ + − =0 (5.19)
dλ2 2B dλ 2B dλ B dλ

34
and
d2 φ 2 dφ dr
+ = 0, (5.20)
dλ2 r dλ dλ
respectively. If we divide the above equation (5.20) with dφ/dλ and note
that  −1 2  
dφ d φ d dφ
2
= ln
dλ dλ dλ dλ
and
2 dr d
ln r2

=
r dλ dλ
we obtain  
d 2 dφ
ln r = 0.
dλ dλ
This means that the quantity inside the logarithm is a constant of motion
and we put it equal to a constant L:


r2 = L. (5.21)

If we compare this with the Newtonian theory, L correspond to angular
momentum per unit mass. So this equation basically states that angular
momentum is conserved.
In a similarly way, we can divide equation (5.15) by dt/dλ and obtain
 
d dt
ln A = 0,
dλ dλ

which again means that the quantity inside the logarithm must be a con-
stant which we will denote by e:

dt
A ≡ e. (5.22)

The interpretation of the constant of motion e will be made later. We can
now use equation (5.21) and (5.22) in equation (5.19) and obtain
2
d2 r e 2 A0 B0 L2

dr
+ + − = 0.
dλ2 2BA2 2B dλ Br3

Multiplying the above equation with 2B dr/dλ we can write it as


0
2
dr d2 r L2 dr

2 A dr dr dr
0 = 2B 2
+ e 2
+ B0 −2
dλ dλ A dλ dλ dλ r3 dλ
"   #
d dr 2 e2 L2
= B − + 2 .
dλ dλ A r

35
The expression inside the square bracket must be a constant
 2
dr e2 L2
B − + 2 = const ≡ −e20 ,
dλ A r
where we implicitly defined e0 (it will be clear later why we defined it this
way). Since B = 1/A, an equivalent form to the above is
 2
L2
 
dr
= e2 − A e20 + 2 . (5.23)
dλ r
We now have dr/dλ as a function of r alone. But we are also able to express
dr/dφ as a function of r. If we divide the above equation with φ̇2 = L2 /r4
we get
 2
r4 k2 r2 e20
 
dr 2
= 2 −r A 1+ 2 . (5.24)
dφ L L
Recall that A was defined as
!
2
rs rQ
A= 1− + 2 ,
r r
so we have that
r 2 e2 r2 e2
   
r2 A 1 + 20 = r2 − rrs + rQ 2
1 + 20

L L
!
2 2
2
rQ e0 rs e2 e2
= rQ − rs r + 1 + 2 r2 − 20 r3 + 02 r4 .
L L L
Substituting this into equation (5.24) we finally get
!
 2 2 e2
dr 2
rQ 0 rs e2 1
r2 + 20 r3 − 2 e20 − e2 r4 (5.25)

= −rQ + rs r − 1 + 2
dφ L L L
We have now obtained dr/dφ in terms of r. But as will be shown below this
can be simplified if we look at two special cases, namely null geodesics and
timelike geodesics corresponding to paths taken by massless and massive
particles, respectively.
Remember that when c = 1 and θ = π/2, the metric is given by
1 2
ds2 = dτ 2 = A dt2 − dr − r2 dφ2 .
A
From equation (5.21), (5.22) and (5.23) we have that
L2 2
dφ2 = dλ
r4
e2
dt2 = 2 dλ2
A
L2
  
2 2 2
dr = e − A e0 + 2 dλ2
r

36
which let us write the differential proper time simply as

dτ 2 = e20 dλ2 .

From this we see that for null geodesics e0 must equal zero, and for timelike
geodesics e0 must be larger than zero. So for timelike geodesics parameter-
ized with the proper time, e0 simply equals unity and equation (5.25) reads
!
 2 2
rQ
dr rs 1
= −rQ + rs r − 1 + 2 r2 + 2 r3 − 2 1 − e2 r4 ,
2

(5.26)
dφ L L L

while for null geodesics equation (5.25) reduces to


2
e2 4

dr 2
= −rQ + rs r − r2 + r . (5.27)
dφ L2

5.2.1 Comparison to Newtonian Mechanics


We are now going to make a comparison of the recently found equation of
motion with the well known case of Newtonian mechanics. Using Q = 0,
we can rewrite equation (5.23) as
2
L2 2GM L2

2 dr 2 2GM
e − e20 = + − e 0 − .
dτ r2 r r3

This can be considered as an energy equation where the first two terms on
the R.H.S. correspond to the kinetic part and the third term to the potential
part. In Newtonian mechanics the total energy E of a particle with mass m
that is influenced only by gravity from a spherically symmetric object with
mass M satisfies
2E 2GM L2 2GM
= ṙ2 + r2 φ̇2 − = ṙ2 + 2 − .
m r r r
By comparing the two above equations we can see that e2 − e20 corresponds
to the total energy per unit mass. But since E is the Newtonian total energy
(i.e. no rest energy) we can make the interpretation that e2 corresponds
to the total relativistic energy per unit mass while e20 corresponds to the
rest energy per unit mass of the particle. So our conclusion earlier that
e0 = 1 (when using τ as the parameter) for massive particles and e0 = 0 for
massless particles now have an intuitive understanding. The big difference
from Newtonian mechanics is that general relativity introduces an extra
term which depend on r−3 which get dominant when r is small. When
r is large (i.e. for weak gravity) this extra term can be neglected and the
equation reduces to the Newtonian one.

37
5.2.2 Circular Orbits of Photons
One interesting special case is that photons (or other massless particles) can
dr d2 r
be in a circular orbit. For circular orbits we must have dλ = 0 and dλ 2 = 0.
From equation (5.19) we must have that
2 2
A0
 
dt dφ
−r =0
2 dλ dλ

while from the metric (since dτ = 0) we get


 2  2
dt 2 dφ
A −r = 0.
dλ dλ

From the two expressions above we must therefore have that rA0 = 2A,
which explicitly reads
!
2 2
rs 2rQ rs rQ
− 2 =2 1− + 2 .
r r r r

This can be written as the quadratic equation

3
0 = r2 − rs r + 2rQ
2
2
with the solution s 2
3rs 3rs 2.
r± = ± − 2rQ
4 4
We have three different cases depending on the value of the expression
inside the square root. These three cases correspond to the existence of two,
one or zero real-valued solutions. The only possible orbit for the photon is
however the solution given by r+ , as will be clear when discussing black
holes and event horizons later in section 5.4. The reason is that r− will be
inside of the event horizon except in the case when 2rQ = rs , then the r−
photon orbit and the event horizon are at the same radius.

5.3 Motion of a Charged Particle


In the Reissner-Nordström spacetime we have a static electric field, so ask-
ing how the motion of a charged particle would be is natural. We will now
derive the equations describing this charged particle with help of the la-
grangian formalism, which we assume that the reader is familiar to.
The Lagrangian for a charged particle is [10]

1
L = gαβ ẋα ẋβ + qAα ẋα
2

38
where q is the charge per unit mass of the particle, Aα the four-potential
and a dot represents a differentiation with respect to the proper time. If we
had no electromagnetic field present the Lagrangian would only consist of
the first term. In flat spacetime this first term would be the kinetic part
(using ηαβ instead of gαβ ), while in a gravitational field it also contains the
gravitational potential energy. The second term arises because we have
potential energy from the electromagnetic field, and splitting it up in its
timelike and spacelike coordinates we have

qAα ẋα = qA0 u0 + qAi ui = qΦu0 + qAi ui ,

where uα is the 4-velocity. The first term in this expression clearly corre-
sponds to the static electric potential energy while the second term corre-
sponds to a "magnetic potential energy". If one make a gauge transforma-
tion of the electric and magnetic potentials the Lagrangian will change by
an addition of a total derivative of a function. This will however not change
the equations of motion.
Since we only have a radial electric field and no magnetic field the only
non-zero component of the four-potential is

Q Q̃
A0 = ≡
4π0 r r
and the Lagrangian now becomes

1h 2 i q Q̃
L= A ṫ − A−1 ṙ2 − r2 θ̇2 − r2 sin2 θ φ̇2 + ṫ.
2 r
The motion of the particle is determined by solving the Euler-Lagrange
equations given by  
d ∂L ∂L
α
− α = 0.
dτ ∂ ẋ ∂x
Since the Lagrangian does not explicitly depend on t or φ we see from the
above equation that ∂L/∂ ṫ and ∂L/∂ φ̇ are both constants:

∂L q Q̃
= A ṫ + ≡ e = const (5.28)
∂ ṫ r
∂L
= r2 φ̇ ≡ L = const. (5.29)
∂ φ̇

The only difference between equation (5.28) and the equation for e in the
case of a massive non-charged particle is the "extra" term q Q̃/r which clearly
corresponds to a static electric potential energy. As before, the particle must
move in a plane and we can put θ = π/2. So for the θ-component (α = 2) the

39
Euler-Lagrange equation is immediately satisfied and for the r-component
(α = 1) we get
 
d ∂L ∂L
0= −
dτ ∂ ṙ ∂r
1   q Q̃
= −A−1 r̈ − A0 ṫ2 + A−2 A0 ṙ2 − 2rφ̇2 − 2 ṫ
2 r
which is the same as equation (5.19) except for the last term containing Q̃.
We will however not use the above equation. Instead we will use the metric
which, when θ = π/2, is given by

ds2 = dτ 2 = A dt2 − A−1 dr2 − r2 dφ2 .

If we divide by dτ 2 and multiply with A this can be written as

ṙ2 + A − A2 ṫ2 + A r2 φ̇2 = 0.

Using equation (5.28) and (5.29) we obtain


!2
2 q Q̃ L2
0 = ṙ + A − e − +A 2
r r
!2
L2
 
2 q Q̃
= ṙ + A 1 + 2 − e − .
r r

By dividing the above equation with φ̇2 = L2 /r4 we can write it as


 2 !2
2 4
 
dr r r q Q̃
= −r2 A 1 + 2 + 2 e −
dφ L L r
!
2 − q 2 Q̃2
rQ
2
= −rQ + rs r − 1 + r2
L2
1   1
+ 2 rs − 2eq Q̃ r3 − 2 1 − e2 r4 .

L L
This is the equation that govern the motion of a charged particle in the
Reissner-Nordström metric. And as it must be, setting q = 0 reduces this to
equation (5.26) which describes a free falling non-charged massive particle.

5.4 Event Horizons and Black Holes


To build up the discussion in a logical way, we will in this section first
consider the Schwarzschild black hole and then proceed to the Reissner-
Nordström black hole. The Schwarzschild metric (when c = 1) is given
by
ds2 = gtt dt2 + grr dr2 − r2 dθ2 − r2 sin2 θ dφ2

40
where
1 rs
gtt = − =1− .
grr r
The metric becomes singular at r = rs because when r → rs we have that
grr → ±∞. Another singular point is clearly r = 0. Note that for any spher-
ically symmetric object (e.g. a star) with a radius larger than rs there will be
no singularity. This is because the Schwarzschild solution is only valid out-
side the object where there is vacuum, which imply that all components of
the stress-energy tensor are zero. Inside the object the stress-energy tensor
does not vanish and one would obtain a solution that does not have any
singularities.
Consider now an object that do have a radius less than rs . Then the
Schwarzschild solution does hold at r = rs and we do have a singularity
at r = rs . An object with the property that its radius is less than rs is called
a black hole. This name is justified (as will be shown later) by the fact that
no massive particle nor light can escape if at a distance closer than rs . This
"boundary" in spacetime is called an event horizon and it marks the surface
for which events inside of it can not affect the outside.
As discussed in section 5.1, gravitational time dilation and redshift will
appear in a gravitational field. When dealing with black holes a few in-
teresting things happen. To an observer, a clock near the black hole will
appear to run slower than a clock further away. By equation (5.12) (letting
rQ = 0), one can see that the time dilation will be infinite for a clock falling
towards the black hole as it approaches the event horizon. This leads to
that it would require an infinite time to reach the event horizon for an object
falling towards the black hole, as seen from an outside observer. Also, from
equation (5.13) one can see that the redshift of an electromagnetic wave
traveling outwards goes to infinity when the point of emission approaches
the event horizon.
Up till now we have worked with the spherical coordinates (t, r, θ, φ).
But what if we were to choose a different coordinate system, would there
still be any singularities? As it turns out, the singularity at r = rs can
actually be made to disappear if one uses the right set of coordinates, and
one can see that it would not take an infinite amount of proper time for an
object to fall past the event horizon. To see that r = rs is merely an apparent
singularity and that not even light can escape from a black hole, we will use
the so-called Lemaître coordinates λ and ρ which differentials transform as
r 
rs rs −1
dλ = dt + 1− dr
r r
r 
r rs −1
dρ = dt + 1− dr.
rs r

41
Taking the difference between the two above equations we have
r r  r
r rs rs −1 r
dρ − dλ = − 1− dr = dr. (5.30)
rs r r rs

Integrating the above results in


 2/3
3
r= rs1/3 (ρ − λ) ,
2

and one will find that the metric can be written as


rs 2
ds2 = dλ2 − dρ − r2 dθ2 − r2 sin2 θ dφ2 . (5.31)
r
With this metric we see only one singularity located at r = 0, while no
singularity arise at r = rs . However this singularity at the center will not
disappear by any transformation, it is really a true singularity.
Consider now a photon that travels along a radial trajectory. We have
that ds2 = dθ2 = dφ2 = 0 and equation (5.31) can be written as
r
rs
dλ = ± dρ
r
where a plus sign correspond to the photon traveling outward while a neg-
ative sign correspond to inward motion. Using the above and equation
(5.30) one can obtain  r 
rs
dr = ±1 − dλ.
r
From this we see that if r < rs , the expression inside the parenthesis is neg-
ative which means that dr is always negative (if dλ is positive). So photons,
regardless if they were emitted outward or inward, end up at the center
of the black hole. We did assume that the photon was traveling radially,
so strictly the conclusion of course only hold for that special case. But if a
photon that was emitted directly outward just inside the event horizon can
not escape, it should not be to hard to believe that a photon emitted in any
direction (when r < rs ) would also be forced to travel towards the center.
We will now consider a Reissner-Nordström black hole. For the Schwar-
zschild black hole we had one event horizon which, when using the coor-
dinates (t, r, θ, φ), we localized by finding singularities in the metric. Recall
the Reissner-Nordström metric given by

ds2 = gtt dt2 + grr dr2 − r2 dθ2 − r2 sin2 θ dφ2

where
2
1 rs rQ
gtt = − =1− + 2.
grr r r

42
A quick inspection suggest that possible event horizons should occur when
2 /r 2 which yields the quadratic equation
0 = gtt = 1 − rs /r + rQ

0 = r2 − rs r + rQ
2

with the solutions


1 q 
2 .
r± = rs ± rs2 − 4 rQ
2
From this we see that, depending on the relative values of rs and rQ , there
is two, one or zero real-valued solutions. A schematic plot of gtt for the
three different cases is showed in figure 1.

2 /r 2 for the three different cases.


Figure 1: The function gtt (r) = 1 − rs /r + rQ
gtt = 0 indicates that there is an event horizon.

Consider the situation when rs > 2 rQ . In this case there is two coor-
dinate singularities at r+ and r− . Note that these two singularities occur
because of our choice of coordinate system. As in the Schwarzschild metric
one can choose a coordinate system in which there exist no singularities
except at r = 0 (which still is a true singularity). The metric in this case can

43
be divided into three regions:
Region 1: r+ < r < ∞
Region 2: r− < r < r+
Region 3: 0 < r < r−
An object coming from region 1 and falling into region 2 would have the
same experience as when crossing the event horizon in the Schwarzschild
black hole. For an outside observer the infalling object would be infinitely
redshifted and it would never reach the event horizon. The proper time for
reaching the event horizon for the object would however be finite. Once
inside region 2 all massive particles and photons necessarily move in the
direction of decreasing r. This unavoidable decrease in r does however stop
when reaching region 3, and the object is therefore not doomed to end up at
the singularity at r = 0. If now moving back (increasing r) to region 2 again,
the object can only move in direction of increasing r and ultimately come
out beyond the horizon at r = r+ [8]. However, it is highly speculative if
this journey through the black hole is physically real.
Consider now the situation when rs < 2 rQ . In this case there is no
singularities when r > 0, and therefore no event horizons. The singular-
ity at r = 0 does still exist, which means that there is no event horizon
preventing someone far away to directly observe this singularity. A singu-
larity with this property (i.e. no event horizon "hiding" it) is called a naked
singularity. It is widely believed, but not proven, that no naked singular-
ity (except maybe the one occurring in the Big Bang model) exist in the
universe [11]. This assumption is called the weak cosmic censorship hypoth-
esis and was formulated by Roger Penrose in 1969. So this solution when
rs < 2 rQ is therefore usually considered to be unphysical.
Lastly, when rs = 2 rQ there exist only one horizon, and the black hole
is called extremal. The event horizon is located at r = rs /2, and in this
case gtt is positive on both sides which means that an observer inside the
event horizon does not necessarily move towards the singularity at r =
0. However, this extremal black hole with only one horizon seems to be
unstable since adding any nonzero mass would turn it to the "regular" case
when rs > 2 rQ .

6 Summary and Conclusion


In this paper we solved the Einstein-Maxwell equations for a spherically
symmetric charged body and found the Reissner-Nordström metric given
by
! !−1
2 2
2 rs rQ 2 2 rs rQ
ds = 1 − + 2 c dt − 1 − + 2 dr2 − r2 dΩ2 .
r r r r

44
It was later shown how this gravitational field give rise to phenomena such
as gravitational time dilation and redshift. With the use of the geodesic
equations we derived the equations of motion for both massive and mass-
less non-charged particles while we used the lagrangian formalism to de-
rive the equations of motion for massive charged particles. We also found
out how these equations corresponds to the well known equations derived
from Newtonian mechanics. Lastly we discussed the properties of a Reissner-
Nordström black hole. Depending on the relative value of rs and rQ there is
two, one or zero event horizons that corresponds to apparent singularities
in the metric. The only true singularity is found in the center of the black
hole.
The Reissner-Nordström metric is a generalization to the Schwarzschild
metric, it can however itself be generalized to the so-called Kerr-Newman
metric. It is a solution to the Einstein-Maxwell equations for an electri-
cally charged rotating axially symmetric body. That is, in addition to the
Reissner-Nordström metric it has a non-zero angular momentum which
has the consequence that it no longer exhibit spherical symmetry. Some
methods that we used in deriving the Reissner-Nordström metric can also
be used when deriving the Kerr-Newman metric.

References
[1] Schwarzschild, K., 1916, Über das Gravitationsfeld eines Massen-
punktes nach der Einsteinschen Theorie, Sitzungsber. Preuss. Akad. D.
Wiss. 50, pages 189-196.

[2] Reissner, H., 1916, Über die Eigengravitation des elektrischen Feldes
nach der Einsteinschen Theorie, Annalen der Physik. 50, pages 106-120.

[3] Nordström, G., 1918, On the Energy of the Gravitational Field in Ein-
stein’s Theory, Verhandl. Koninkl. Ned. Akad. Wetenschap., Afdel. Natu-
urk., 26, pages 1201-1208.

[4] Jackson, J. D., 1998, Classical Electrodynamics, John Wiley & Sons Inc,
3rd edition,

[5] Weinberg, S., Gravitation and Cosmology, John Wiley & Sons Inc, 1st
edition, 1972.

[6] Wheeler, J. A., Misner, C. W and Thorne, K. S, 1973, Gravitation, W. H.


Freeman and Co, 1st edition, page 141.

[7] Foster, J. and Nightingale, J. D., 2005, A Short Course in General Relativ-
ity, Springer, 3rd edition,

45
[8] Carroll, S. M, 2003, Spacetime and Geometry: An introduction to General
Relativity, Addison Wesley, 1st edition.

[9] Wald, R. M, 1984, General Relativity, The University of Chicago Press,


1st edition, page 39-40.

[10] Chandrasekhar, S., 1983, The Mathematical Theory of Black Holes, Oxford
University Press, 1st edition, page 224.

[11] Singh, T. P., 1997, Singularities and Cosmic Censorship, J. Astrophys.


Astr., 18, pages 335–338.

46

You might also like