Vous êtes sur la page 1sur 43

TOPICS OF RIEMANNIAN GEOMETRY

Nikolai V. MITSKIEVICH, Guadalajara-New Delhi

1 Algebra of Riemannian Geometry


PLEASE SEE THE PAGES 8 & 9 IN COLOURS DIRECTLY ON
THE SCREEN!!!
Quite a number of expressions and relations of Riemannian geometry
will be extensively used below. In existing literature [see for a traditional ap-
proach, e.g., Eisenhart (1926, 1933, 1972); Schouten and Struik (1935, 1938);
as to a presentation of general relativity from the mathematical viewpoint,
see Sachs and Wu (1977)], they are scattered in different parts of various pub-
lications, and the notations differ substantially from one source to another.
Many relations are naturally given without derivation. The reader may skip
this section but look into it merely for clarification of some formulae. It is
worth emphasizing that we take everywhere the speed of light equal to unity
(c = 1), the space-time signature being (+ − −−); the Greek indices are
four-dimensional, running from 0 to 3, the Einstein summation convention is
adopted, including collective indices. Symmetrization and antisymmetriza-
tion with respect to sets of indices are denoted by the standard Bach brackets,
(· · · ) and [· · · ] respectively, with the indices in the brackets. Tetrad indices
are written in separate parentheses, e.g., F(µ)(ν) . For important details of ab-
stract representation of the tensor calculus and its applications to field theory,
see (Israel 1970, Ryan and Shepley 1975, Eguchi, Gilkey and Hanson 1980,
Choquet-Bruhat, DeWitt-Morette and Dillard-Bleick 1982, von Westenholz
1986).
We admit that the tensor algebra in an arbitrary basis and in abstract
notations, is commonly known. Hence we begin with the Cartan forms. To
make notations shorter, we shall sometimes employ collective indices, e.g. in
the basis of a p-form [cf. Mitskievich and Merkulov (1985)]:

dxa := dxα1 ∧ · · · ∧ dxαp , thus a = {α1 , α2 , . . . , αp } (1.1)

(completely skew-symmetrized). Here exterior (wedge) product is supposed


to be an antisymmetrization of the tensorial product,

dxα1 ∧ · · · ∧ dxαp := dx[α1 ⊗ · · · ⊗ dxαp ] . (1.2)

1
Similarly, we shall write tetrad basis of a p-form as θ(a) (then a collective
or individual index in the parentheses pertains to a tetrad field); this con-
struction being analogous to dxa , and representing an exterior product of the
covector basis 1-forms, θ(α) = g (α) β dxβ . Here g (α) β are the covariant tetrad
components corresponding to the lower index β of a set of four independent
covectors (enumerated by the tetrad index (α)). We write the tetrad index
in its individual parentheses immediately after the root letter g, the same as
that of the metric tensor, only then followed by the (coordinated) component
number (here, β).1 Then a p-form α can be written as

α = αa dxa = α(a) θ(a) , while α ∧ β = (−1)pq β ∧ α (1.3)

where p = deg α, q = deg β (degrees — or ranks — of the forms α and β).


The axial tensor of Levi-Cività is defined with help of the corresponding
symbol,2
Eαβγδ = (−g)1/2 αβγδ , E αβγδ = −(−g)−1/2 αβγδ . (1.4)
We use the Levi-Cività symbol strictly as a symbol, that is, in the sense that
it represents a set of constant scalars, so that one cannot raise or lower its
indices; see however the footnote 2.
To find some important properties of αβγδ , and to relate these properties
to those of Kronecker’s deltas, let us somewhat formally consider an artificial
1
In the metric space, to which the pseudo-Riemannian space-time belongs, any tensor
(in particular, vector) can be taken in its covariant and contravariant form (with respect to
any of its indices) due to existence of the metric tensor which in fact is a generalization of a
mere identity matrix; thus to the covariant basis θ(α) always corresponds the contravariant
basis X(α) : this is the alternative use of the term “dual”. However such a duality has
nothing in common with the duality involving the Levi-Cività axial tensor.
2
This means that the Levi-Cività symbol itself can be simultaneously interpreted as a
contravariant axial tensor density (of weight +1) and a covariant axial tensor density of
weight −1: αβγδ = −(−g)1/2 E αβγδ = (−g)−1/2 Eαβγδ . Remember that a tensor density
is the corresponding tensor multiplied by (−g)w/2 where w is the weight of the tensor
density, so that, for example, the transformation law of a scalar density L (of weight w;
for w = 1, as the Lagrangean density usually is denoted), is L 0 0
(x ) = |J|
−w
L (x), |J|
being absolute value of the Jacobian of the coordinate transformation. Thus the reader
may find some discrepancies (from the physicists’ viewpoint) on p. 129 of the otherwise
quite good book by B.F. Schutz (1980). One has to take into account that Schutz simply
used the terminology utilized by mathematicians, in a contrast to physicists!

2
(however, well constructed) determinant
α α α α
δκ δλ δµ δν
δβ δβ δβ δβ

= Dαβγδ

κ λγ µ ν (1.5)
δγ δ δγ δγ κ λµν
κ λ µ ν
δδ δδ δδ δδ
κ λ µ ν

(this Dκαβγδ
λµν still represents merely a notation). Due to the well-known general
properties of determinants, we immediately see that Dκαβγδ λµν is skew symmet-
ric in both all upper and all lower indices (each time in four indices which
coincides with the 4 × 4-structure of the very determinant!). From (1.5) one
already can understand the rôle of Levi-Cività’s epsilon in dealing with 4× 4-
determinants in general. It is easy to figure out that Dκαβγδ 0123
λµν ≡ D0123 αβγδ κ λµν ,
0123
while D0123 ≡ 1. Moreover, an immediate opening of the determinant in the
left-hand side of (1.5) yields

Dκαβγδ α β γ δ
λµν ≡ 4!δ[κ δλ δµ δν] . (1.6)

Thus we have shown that

αβγδ κλµν = 4!δ[ακ δλβ δµγ δν]


δ
, (1.7)

or for the Levi-Cività axial tensors,

E αβγδ Eκλµν = −4!δ[ακ δλβ δµγ δν]


δ
. (1.8)

(In the last two expressions, it is remarkable how the products of epsilons and
of the Levi-Cività axial tensors — without any contractions! — are found to
be equal to products of the Kronecker deltas, also without contractions. Con-
secutive contractions of these relations (let us restrict them to those for the
Levi-Cività axial tensors) yield the general formula (we write it conveniently
using collective indices, both free and dummy ones)

Eag E bg = −p!(4 − p)!δab , (1.9)

so that in all particular cases we have as a continuation of (1.8) the following


relations:
E αβγν Eκλµν = −3!δ[ακ δλβ δµ]
γ β 
, E αβµν Eκλµν = −2 · 2!δ[ακ δλ] ,
αλµν α κ λµν
, (1.10)
E Eκλµν = −3!δκ , E Eκλµν = −4!

3
or, more symmetrically,

E αβγδ Eκλµν = −0!4!δ[ακ δλβ δµγ δν]


δ
, E αβγν Eκλµν = −1!3!δ[ακ δλβ δµ]
γ
,
β
E αβµν Eκλµν = −2!2!δ[ακ δλ] , E αλµν Eκλµν = −3!1!δκα , E κλµν Eκλµν = −4!0!.

The second expression in (1.4) is, of course, the result of raising the indices
in the first expression: E κλµν = (−g)1/2 αβγβ g ακ g βλ g γµ g δν = (−g)1/2 κλµν det(g στ )
= (−g)1/2 /gκλµν = −(−g)−1/2 κλµν , Q.E.D.;3 det(g στ )= {det (gστ )}−1 .
In our four-dimensional (essentially, even-dimensional) spacetime mani-
fold, the Levi-Cività axial tensor has the following properties:

Eag = (−1)p(4−p) Ega ≡ (−1)p Ega ,
. (1.11)
Eag E bg = −p!(4 − p)!δab ; #a = #b = p = 4 − #g.

While using here collective indices, we denote the number of individual in-
dices contained in them, by #. The Kronecker symbol with collective indices
is totally skew by a definition,
β1 β
δab ≡ δαβ11···β p p b v bv
···αp := δ[α1 · · · δαp ] , δ[a δu] ≡ δau , (1.12)

so that Aa δab ≡ Ab (let Aa be skew in all its individual indices). Dual con-
jugation of a form is denoted by the Hodge star ∗ [see Eguchi, Gilkey and
Hanson (1980)] put before the form. The star acts on the basis as
1
∗dxa := (4−p)! E a g dxg , ∗ α := αa ∗ dxa ,

√ , (1.13)
∗1 = 4!1 Eαβγδ dxα ∧ dxβ ∧ dxγ ∧ dxδ = −g(dx)

(dx) being the four-dimensional volume element corresponding to the four


covectors dx0 , dx1 , dx2 , and dx3 . Thus ∗ ∗ α = (−1)p+1 α (here even-
dimensional nature of the manifold is essential); in particular, ∗ ∗ 1 = −1.
It is worth stressing that the sequence of the (lower and upper) indices is of
great importance too, e.g. it was not the question of a chance when we wrote
E a g and not Eg a [which in fact is equal to (−1)p E a g , cf. (1.11)].
Scalar multiplication of Cartan’s forms is realized with help of consecutive
dual conjugations:
 
al p+1 (p + q)! [l a] #a = p,
∗(dxk ∧ ∗dx ) = (−1) δk dx , . (1.14)
p! #k = q = #l
3
Quod erat demonstrandum (Latin).

4
Here dxk , of course, does not mean lowering of the indices of coordinates
under the differential, this is, strictly speaking, the lowering of the collective
index nonly after differentiation. In (1.14) it is important that #k + 4 −
#(al) ≤4≥0
The proof of (1.14) consists of the following steps: (I):

E al g
∗ dxk ∧ ∗dxal = ∗ (dxk ∧ dxg ) ;

(4 − p − q)!

(II):
g Ek g b b
∗ (dxk ∧ dx ) = dx ;
p!
(III):
E alg Ekgb
∗ dxk ∧ ∗dxal = dxb ;

p!(4 − p − q)!
and this easily reduces to (IV):

(−1)p E alg Ebkg b (−1)p+1 (p + q)! al b (−1)p+1 (p + q)! [l a]


dx = δbk dx ≡ δk dx .
p!(4 − p − q)! p! p!

In particular, when we take in (1.14) a scalar multiplication of two 1-


forms, this gives

∗(dxλ ∧ ∗dxµ ) = ∗(dxκ ∧ ∗dxµ )g κλ = −g λµ . (1.15)

This last relation is equivalent to

∗(E µ πρσ dxλπρσ ) = E µ πρσ E λπρσ = −3! g λµ , (1.16)

as well as to
dxα · dxβ = g αβ (1.17)
(in fact, this property should be considered as introduction of a symmetric
metric tensor g in the manifold under consideration). For the more general
tetrad basis θ(α) we have

θ(α) · θ(β) = g (α)(β) . (1.18)

Contravariant objects are treated with the use of the corresponding coor-
dinated vector basis ∂µ or tetrad vector basis X(µ) (where necessary, their

5
tensor products). Here ∂µ is usually considered as the partial differen-
tiation operator ∂x∂ µ acting on (scalar) functions; see however comments
in the next footnotes. The scalar product operation is defined so that
dxα · ∂µ ≡ ∂µ · dxα = δµα . The tetrad basis is related to its tetrad com-
ponents (the covariant case was mentioned above; in the contravariant one,
X(σ) = g(σ) µ ∂µ , hence X(σ) · dxα = g(σ) α , while θ(σ) · ∂α = g (σ) α ). Thus
there always exists a pair of “conjugated” tetrads: one is the set of linearly
independent contravariant vectors, X(σ) , and another, a set of linearly in-
dependent covariant vectors, θ(σ) , so that g(σ) α g (σ) β = δβα and, equivalently,
g(ρ) α g (σ) α = δρσ : remember the alternative concept of “duality” mentioned
above (not to be confused with the Hodge conjugation).
The usual dual conjugation of a bivector will be also denoted by a star,
but this will be written over (or under) the pair of indices to which it is
applied, e.g.
1 κλµν
F κλ
∗ := E Fµν , (1.19)
2
∗∗
thus F αβ ≡ −Fαβ . It is obvious that Fκλ is skew (we called it bivector for this
reason), so that the dual conjugation does not lead to any loss of information.
Our choice of place for the star is justified here by the convenience in writing
down the so-called crafty identities where asterisks are applied to different
pairs of indices [see e.g. (Mitskievich and Merkulov 1985)]:

V κλ ∗
∗ µν = −Vµν κλ − δµν
κλ
Vστ στ + 2δµν
κτ
Vστ σλ + 2δµν
τλ
Vστ σκ . (1.20)

In fact, these different applications of asterisks simply point out non-coinciding


but equivalent ways to deal with the dual conjugation. For example, one
may write equally ∗F as (1/2)Fµν ∗ dxµν ≡ (1/2)F µν ∗
dxµν ; and it is clear
that F = (1/2)Fµν dxµν = −(1/2)F µν ∗
∗ dxµν . Let us consider now the tensor
used in (1.20) without leaving out important details. This tensor is skew in
its first pairs of indices as well as in their last pair (Vαβγδ = V[αβ][γδ] , but the
relationship between the pairs of indices is not fixed. Thus the tensor can
be considered as a double 2-form, V = Vαβγδ dxαβ ⊗ dxγδ . The left-hand side
in (1.20) can then be written as Vαβ γδ ∗ dxαβ ⊗ ∗∂γδ (we consider this as a
natural extension of the Hodge star concept, though with the basis ∂ one has
to deal with contravariant parts of objects; however, the rules introduced for
the usual Cartan forms, can be applied without contradictions in this case
too). In order not to formally contradict to the general amenity regulations
in the geometry, one may, of course, write Vαβ γδ ∗ dxαβ ⊗ ∗dxγδ , as this was

6
done in (1.14). Then this expression will be equal to

+δκγ δλδ δαµ δβν − δκγ δλδ δβµ δαν + δκγ δβδ δλµ δαν − δκγ δβδ δαµ δλν
 
 
γ δ µ ν γ δ µ ν γ δ µ ν γ δ µ ν
 
 +δκγ δαδ δβµ δλν − δκγ δαδ δλµ δβν + δλγ δκδ δβµδαν − δλγ δκδ δαµδβν

 

 

1  +δλ δα δκ δβ − δλ δα δβ δκ + δλ δβ δα δκ − δλ δβ δκ δα 
− V αβ γδ dxκλ ⊗ dxµν .
4 
 +δαγ δλδ δβµ δκν − δαγ δλδ δκµ δβν + δαγ δκδ δλµ δβν − δαγ δκδ δβµ δλν 

+δαγ δβδ δκµ δλν − δαγ δβδ δλµ δκν + δβγ δαδ δλµ δκν − δβγ δαδ δκµ δλν

 


 

+δβγ δκδ δαµ δλν − δβγ δκδ δλµ δαν + δβγ δλδ δκµ δαν − δβγ δλδ δαµ δκν
 
(1.21)
It is not too convenient to write down all 24 terms appearing as the pre-
liminary result of the multiplications in (1.21). Therefore let us look for
simplifications surely awaiting for our attention somewhere around the cor-
ner. The first of these simplifications are quite impatient, they fill the first
two adjacent places in the first and in the second lines, and they all four
contribute equally, yielding the term −V µν κλ dxκλ ⊗ dxµν (the denominator 4
just disappeared). Here you also see that the first two indices in V became
its last pair of indices, and vice versa, please take into account the crucial
in this sense basis factor dxκλ ⊗ dxµν . Well, we now have to deal with the
remaining 20 terms. Let it be a (not too bad) exercise since there already is
enough room for the comparison, see (1.20) vis. (1.21).
In electrodynamics [cf. (Wheeler 1962, Israel 1970, Mitskievich and
Merkulov 1985)] we consider here two particular cases to which the crafty
identities lead; these cases together make it possible (and easy) to explic-
itly represent any invariant built of the electromagnetic field tensor F as a
function of only two invariants, I1 = Fµν F µν and I2 = F µν ∗
F µν (the second
is, of course, only a pseudo-invariant, i.e., an axial scalar). First, we take
Vαβγδ = Fαβ Fγδ whose substitution into (1.20) with one additional contrac-
tion yields
1 λ
Fµν F λν − F µν

F λν
∗ = δ Fστ F στ . (1.22)
2 µ

Another choice of Vαβγδ = Fαβ F γδ , also with a contraction, leads to

1

F µν F λν = δµλ F στ∗ F στ . (1.23)
4
An example of application of these identities is worth being given: taking the
ξµ
(pseudo-)invariant Fµν F λν
∗ Fλξ F  and applying (1.23) to two interior factors,
one directly comes to Fµν 4 δξ I2 F ξµ = − 14 I1 I2 . By the way, all invariants
1 ν

7
and pseudo-invariants containing odd total number of usual and dually con-
jugated F s, identically vanish.

See the next two pages:

8
CORRECTION!!!

THE CALCULATION OF V ∗ µν
κλ ∗

+δκγ δλδ δαµ δβν − δκγ δλδ δβµ δαν +δκγ δβδ δλµ δαν − δκγ δβδ δαµ δλν
 
 
+δκγ δαδ δβµ δλν − δκγ δαδ δλµ δβν +δλγ δκδ δβµ δαν − δλγ δκδ δαµ δβν

 


 

 +δ γ δ δ δ µ δ ν − δ γ δ δ δ µ δ ν + δ γ δ δ δ µ δ ν − δ γ δ δ δ µ δ ν
 

∗ µν = V αβ
−4V κλ λ α κ β λ α β κ λ β α κ λ β κ α
=
∗ γδ γ δ µ ν γ δ µ ν γ δ µ ν γ δ µ ν

 +δ δ
α λ β κδ δ − δ δ
α λ κ βδ δ +δ δ
α κ λ βδ δ − δα κ δβ δλ
δ 

+δα δβ δκ δλ − δα δβ δλ δκ + δβ δα δλ δκ − δβγ δαδ δκµ δλν
γ δ µ ν γ δ µ ν γ δ µ ν

 


 

+δβγ δκδ δαµ δλν − δβγ δκδ δλµ δαν + δβγ δλδ δκµ δαν − δβγ δλδ δαµ δκν
 

= 4V µν κλ +2V στ στ (δκµ δλν − δλµ δκν )+4V σν σκ δλµ − 4V σµ σκ δλν −4V σν σλ δκµ +4V σµ σλ δκν
(we have already used the symmetry properties V µν κλ = V [µν] [κλ] for mutual
reduction of similar terms). Taking now into account that, for example,
∗ µν =
V σν σκ ≡ V στ σκ δτν , we combine some pairs of terms coming to −4V κλ ∗

= 4V µν κλ +2V στ στ (δκµ δλν − δλµ δκν )+4V στ σκ (δλµ δτν − δτµ δλν )−4V στ σλ (δκµ δτν − δτµ δκν ) .

Application of the notation δκµ δλν − δλµ δκν = 2δκλ


µν
with a division by (−4) yields
∗ µν = −V µν −V στ δ µν −2V στ δ µν + 2V στ δ µν .
V κλ ∗ κλ στ κλ σκ λτ σλ κτ

Finally, let us rewrite the just obtained identity without using colours and
interchanging the next to last term and the last one:
∗ µν = −V µν − V στ δ µν + 2V στ δ µν − 2V στ δ µν .
V κλ (1.24)
∗ κλ στ κλ σλ κτ σκ λτ

It is obvious that both left- and right-hand sides of this identity are equally
skew-symmetric in the free indices κ, λ as well as in µ, ν.
Contraction of the identity (1) in ν and λ yields
∗ µν = −V µν − V στ δ µν + 2V στ δ µν − 2V στ δ µν .
V κν (1.25)
∗ κν στ κν σν κτ σκ ντ

Though this is a special restricted case following from (1), we shall now see
its importance, in particular for electrodynamics. To this end we substitute
in (2) V µνκλ = F µν Fκλ :

∗ F µν = 1 I δ µ
F µν Fκν − F κν (1.26)
1

2 κ
9
where I1 = Fαβ F αβ is the electromagnetic first invariant. Another substitu-
∗ , leads to
tion, V µνκλ = F µν F κλ

µν ∗ 1 µ
Fκν F µν
∗ = F F κν = I2 δκ (1.27)
4
∗ . An example of appli-
where the second (pseudo-) invariant is I2 = F στ F στ
cation of the identities (3) and (4) is the straightforward determination of
any invariant being an arbitrary even power of F and/or ∗F (in the case of
odd number of factors such a product invariant identically vanishes; in many
cases, this is obvious without application of the above identities).
I gratefully acknowledge the detection of my error in three last coefficients
in formula (1.20) in my file on Topics of Riemannian Geometry by students
of Licenciatura on September 11, 2008.

10
For the Levi-Cività axial tensor one has
κλ ∗ κλ
E κλ
∗ µν ≡ E µν = −2δµν , (1.28)

for the Riemann–Christoffel curvature tensor,


∗ κλ κλ κλ [κ λ]
R αβ ∗ = −Rαβ + Rδαβ − 4R[α δβ] , (1.29)

and for the Weyl conformal curvature tensor,


∗ γδ
C αβ ≡ Cαβ γδ
∗ . (1.30)

By the way, a repeated application of the crafty identities to a construction


quadratic in the curvature tensor (with subsequent contractions, so that only
two indices remain free ones), leads to the well known Lanczos identities,

Rαβγ µ Rαβγν − 41 Rαβγδ Rαβγδ g µν − 2Rµαβν Rαβ + Rαβ Rαβ g µν −


(1.31)
−Rαµ Rαν + R Rµν − 41 R2 g µν = 0

[see our derivation of these identities (Mitskievich 1969) which differs radi-
cally from that by Lanczos (1938) who used less straightforward integral re-
lations]. Note that the invariant Rαβγδ R αβ γδ
∗ ∗ is known as the Bach–Lanczos
invariant, or the generalized Gauss–Bonnet invariant;4 like the electromag-
netic invariant Fαβ F αβ
∗ , it represents (in a four-dimensional world) a pure
divergence, so that it does not contribute to the field equations in four di-
mensions. Cf. also DeWitt (1965).

2 Differentiations in Riemannian Geometry


There are two closely related differentiation operators using the important
concept of connection in Riemannian geometry: the more modern nabla oper-
ator with which we begin this section, and the comparatively older ;-operator
to be encountered in more traditional treatment of the Riemannian geometry.
Finally, the Lie derivative does not use the concept of connection (though it
can be written with the help of covariant derivatives). But the Lie derivative
serves rather for more general description of covariant derivatives than the
4
The first of Lagrangians of the Lovelock (1971) series reduces to it.

11
latter ones are needed for its conceptual introduction. From these differential
operators one easily can pass to many variants of the crucially important con-
cept of Riemannian geometry, that of different kinds of transports in spaces
endowed with curvature. This last way of treating curved manifolds is natu-
rally related to one of the central object both of geometry and of physics in
curved manifolds, the geodesic equation.

2.1 The Nabla Operator


The covariant differentiation axioms [see e.g. Ryan and Shepley (1975)] are:
(1) ∇v T has the same tensor rank and valence properties as T , v being
always a vector.
(2) ∇v is a linear operation:
∇v (T + U ) = ∇v T + ∇v U,
∇au+bv T = a∇u T + b∇v T.

(3) The Leibniz property5 holds (including the contraction operation in the
Riemannian geometry).
(4) Action of ∇ on a function6 is ∇v f = vf , where v = v α ∂α ≡ v (α) X(α) .
(5) Metricity property: ∇v g = 0 (for a further generalization of the geometry,
the non-metricity tensor appears on the right-hand side).
(6) Zero torsion axiom:
∇u v − ∇v u = [u, v] + T, T = 0.
(in a generalization of the geometry, the torsion tensor7 T becomes different
from zero).
These axioms8 are most simply realized when the connection coefficients
are introduced,
(γ) (γ)
∇X(α) X(β) = Γ(β)(α) X(γ) ; ∇X(α) θ(γ) = −Γ(β)(α) θ(β) (2.1)
5
Distributive property when product expressions are differentiated.
6
A “scalar”; the concept of scalar is somewhat different for the covariant differentiation
denoted by a semicolon; see more in the footnote 9.
7
Like the curvature operator below, the torsion T (if 6= 0) here also describes
 a tensor,
now of the rank three, by the rule: ∇X(µ) X(ν) − ∇X(ν) X(µ) − X(µ) , X(ν) = T (λ) (µ)(ν) X(λ)
[cf. (3.2)]. Then, of course, the connection coefficients following from (2.1), should be
modified.
8
There could be also added the seventh (in some sense) “axiom”: [∇u , ∇v ] = ∇[u,v] +
R(u, v) [cf. the definition (3.1)], with R = 0, as an alternative to T = 0. Then, of course,

12
(only one of these two relations is independent). When the structure coeffi-
cients of a basis are introduced with help of
(γ) (γ) (γ) (γ)
[X(α) , X(β) ] = C(α)(β) X(γ) , C(α)(β) = Γ(β)(α) − Γ(α)(β) (2.2)
γ
(for a holonomic, or coordinated basis, Cαβ = 0; as usually, we write in
individual parentheses only the tetrad and not coordinated basis indices),
the general solution satisfying the whole set of axioms, from 1 to 6, is
(γ)  
Γ(α)(β) = 21 g (γ)(δ) X(β) g(α)(δ) + X(α) g(β)(δ) − X(δ) g(α)(β) 


h i  (2.3)
1 (γ) (δ) ()(γ) (δ) ()(γ)
+ 2 C(β)(α) + C()(β) g(α)(δ) g + C()(α) g(β)(δ) g . 

We give here the hint for deduction of (2.3): first, to look at the first
square brackets and  − X(δ) g(α)(β) =
 represent them as X(β) g(α)(δ) + X(α) g(β)(δ)
∇X(β) X(α) · X(δ) + ∇X(α) X(β) · X(δ) − ∇X(δ) X(α) · X(β) ; then to apply
nabla-differentiation and the definition of structure coefficients (2.2). Finally,
(γ) (γ) (γ) (γ)
to do a recombination of the expression Γ(α)(β) +Γ(β)(α) into 2Γ(α)(β) +Γ(β)(α) −
(γ) (γ)
Γ(α)(β) , the last two terms forming the structure coefficient C(β)(α) . Then one
has to rewrite the result already into the form of (2.3), and all is done.
In a coordinated basis (where the basis vectors and covectors are usually
written as ∂α and dxα ), the connection coefficients reduce to the Christoffel
symbols (which are written without parentheses before and after each index,
and which obviously are symmetric in the two lower indices):
1
Γγαβ = g γδ (gαδ,β + gδβ,α − gαβ,δ ) . (2.4)
2
Orthonormal bases and those of Newman–Penrose [see Penrose and Rindler
(1984a, 1984b)] are special cases of bases whose vectors have constant scalar
products (i.e., the corresponding tetrad components of the metric are con-
stant). For such bases the connection coefficients are skew in the upper and
the first lower indices; they are sometimes called Ricci rotation coefficients.
Thus, in a coordinated basis, the differentiation of A = Aα yields
∇∂γ A = (∂γ Aα )dxα + Aα ∇∂γ dxα = Aα,γ − Aβ Γβγα dxα =: Aα;γ dxα , (2.5)


the theory changes drastically, but in literature one can find an assertion that in this case
the former gravitational effects become those of the torsion. The author is not ready to
swear that this is completely true, and the following exposition strictly belongs to the
Riemannian geometry.

13
and for V = V α ∂α ,

∇∂γ V = V α ,γ + V β Γαγβ ∂α =: V α ;γ ∂α ,

(2.6)

in accordance with the (old) traditional notations for the ;-covariant deriva-
tive. From these definitions it is easy to see that the ;-covariant derivatives
of vectors, Aα;γ and V α ;γ , are components of rank 2 tensors. In general the
;-covariant derivatives of rank r tensors also are (in the sense of components)
tensors of the rank r + 1 (cf. the axiom 1 of the ∇-covariant derivative).
Let us revisit the definition of connection coefficients, using now the coor-
dinated basis. Then dxα = δβα dxβ ; it is probably better to highlight more the
sense of the index α in this expression (that it is not a component number
of a vector, but the number of a covector in a coordinated basis), entering it
into a circle (the parentheses already have other meanings): dx α
= δβ
α
dxβ .
Then

α
δβ;γ α
= δβ,γ − Γσγβ δσ
α
= −Γ α
γβ ,

since the Kronecker delta is a set of constants, and the encircled index does
not pertain to the coordinated components (now it has more in common with
a tetrad component number, but for a non-orthonormal tetrad). Thus we
come to an equivalent of (2.1) in a coordinated basis, ∇∂γ dxα = −Γαγβ dxβ .
β
The subindex in ∂ α = δ α ∂β can be treated similarly, only the sign will be
changed to the inverse one.

2.2 ;-covariant Differentiation


Thus, covariant differentiation can also be denoted by a semicolon (;) before
the differentiation index (especially, in the elder literature9 ). Then, according
to the covariant differentiation axioms (in particular, the axiom 1, however
paradoxical this may appear), it leads to an increase by unity of the valence
of the corresponding tensor. As Trautman (1956, 1957) remarked, such a
covariant derivative can be defined as

Ta;α := Ta,α + Ta |τσ Γστα (2.7)


9
In the old-fashioned coordinated component notations, tensor components are usually
treated as synonyms of the corresponding tensors, so it is then quite awkward to consider
them as a set of scalars, but in the modern abstract (coordinate-free) notations, tensor —
and connection — components are dealt with just as a set of scalar functions. Here we
use a fairly innocuous mixture of both styles.

14
where the coefficients Ta |τσ determine the behavior of the quantity Ta (which
could be not merely a tensor, but also a tensor density or some more general
object) under infinitesimal transformations of coordinates,  being a dimen-
sionless infinitesimal scalar constant:
x0µ = xµ + ξ µ (x),
δTa := Ta0 (x0 ) − Ta (x) = Ta |τσ ξ σ ,τ , (2.8)
i.e. Ta |τσ = −1 ∂δTa /∂(ξ σ ,τ ). It is clear that the coefficients Ta |τσ possess a
property
(Sa Tb )|τσ = Sa |τσ Tb + Sa Tb |τσ (2.9)
which is closely related to the Leibniz property of differentiation. The co-
variant (in the sense of ∇u ) differentiation axioms are readily reformulated
in terms of ;-differentiation; in particular, the axiom 5 takes the form
gαβ;γ = 0 or, equivalently, Eκλµν;γ = 0 (2.10)
which (in each of these two forms) immediately yields the definition (2.4) of
Γγαβ .
We shall now show in general that after the ;-differentiation the resulting
object has property of a tensor of the same valences as those of the tensor
before the differentiation, plus one additional covariant valence. To this end
we shall use collective indices, but in another sense than in the Cartan for-
malism (now, without antisymmetrization). Thus the true analogue of the
oversimplified expressions (2.5) and (2.6) reads
∇∂γ Ta b dxa ⊗ ∂b = Ta b ;γ (dxa ⊗ ∂b ) .

(2.11)
Similarly, a chain of equations follows:
b b ∂xβ b
∇∂γ0 Ta0 (x0 ) (dx0a ⊗ ∂b0 ) = Ta0 0 0a 0
;γ (x ) (dx ⊗ ∂b ) =Ta ;β (x) (dxa ⊗ ∂b ) .
∂x0γ
(2.12)
Now note that the successive contraction of the basis objects with collective
indices yields
∂xa ∂x0l
(dx0a ⊗ ∂b0 ) · ∂k0 ⊗ dx0l = δka δbl , (dxa ⊗ ∂b ) · ∂k0 ⊗ dx0l =
 
, (2.13)
∂x0k ∂xb
so that finally
l ∂xβ ∂xa ∂x0l b
Tk0 ;γ (x0 ) = Ta ;β (x). (2.14)
∂x0γ ∂x0k ∂xb
This is precisely the transformation law we have spoken about above.

15
2.3 The Lie Derivative
The definition of covariant derivative (2.7) may be deduced from the concept
of Lie derivative,

£ξ Ta := Ta,α ξ α − Ta |τσ ξ σ ,τ ≡ Ta;α ξ α − Ta |τσ ξ σ ;τ , (2.15)

which can be written for objects of the most general nature as

£ξ Ta := −1 (Ta (x) − Ta0 (x)) (2.16)

[cf. Yano (1955)].10 Due to the property (2.9), the definition (2.15) of the Lie
derivative under an infinitesimal transformation yields the Leibniz property.
However, if we consider the finite-difference ‘Lie derivative’ (2.16) when it
is not subjected to an infinitesimal transport (the parameter  is now finite,
./
so that we have to introduce instead of £ a new notation, say, £ ), then a
generalized Leibniz property has to be introduced,
./
 ./   ./   ./   ./ 
£ξ (Sa Tb ) = − £ξ Sa £ξ Tb + £ξ Sa Tb + Sa £ξ Tb .

Seemingly, it is quite different from the former one (see the first right-hand
side term). However this situation is typical for finite differences, automati-
cally changing to the usual Leibniz property when one goes to the infinitesi-
mal case (in the lower order of ), and only the last two right-hand side terms
survive.
10
The quantity Ta0 (x) which will appear also in the proof of Noether’s theorem, describes
the change of dependence of the components Ta due to the transformation of coordinates
from xµ to x0µ . This may seem to be somewhat artificial, so we give here an explanation
for it in a form of parable. Consider a region of space-time filled with integral curves
of the vector field ξ µ (x) and two points, P and P 0 , belonging to the same curve, such
that the coordinates of P 0 in the system {x0 } are numerically equal to those of P in {x}:
x0µ |P 0 = xµ |P . This corresponds to the situation when at both points there are indepen-
dent observers having exactly the same mentality so that they ascribe to their respective
positions in space-time the same numerical characteristics thus using these two coordi-
nate systems, {x} and {x0 }, respectively. Let these observers have all necessary devices
for measurement and reproduction of the field Ta at their respective points. Moreover,
suppose that each of them can communicate with the other telepathically. Then the in-
formation about measurement of Ta made by the observer at P 0 , sent instantaneously to
the observer at P and reproduced by him, is related to the independent measurement of
Ta by this latter observer directly through the Lie derivative (2.16).

16
It is important that both objects in the parentheses in (2.16), have the
same argument (non-primed one). This leads to a possibility to extend the
operation Ta;α even to quantities of the type of a connection.
Concerning the transformations of coordinates, we usually consider only
those which are reducible to infinitesimal ones (i.e., we speak about the con-
nected component of the identity transformation). In this case it is always
possible to recover finite transformations through integration of the infinites-
imal ones, but of course not such as inversions. Thus we may write
x0µ = xµ + ξ µ , ξ µ = ξ µ (x), (2.17)
 being an infinitesimal constant, and ξ an arbitrary differentiable vector field
(note that a vector field is inextricably related to the general concept of con-
gruence which is central in the definition of the Lie derivative). The standard
transformation coefficients of tensor components by virtue of infinitesimality
read
∂x0 µ µ µ ∂xµ µ µ
ν
= δ ν + ξ ,ν , 0 ν = δν − ξ ,ν , (2.18)
∂x ∂x
(up to the first order of magnitude terms). In this sense
Γλµν;α = Rλ (µν)α . (2.19)
The Lie derivative (2.15) reduces to the gradient projected on ξ when it is
applied to a scalar function. Since the set of coordinates xµ is a collection of
four scalar functions, this means that
£ξ x µ = ξ µ (2.20)
for any vector field ξ.
Concerning the Lie derivative, it is worth also mentioning the relations
£ξ gµν = ξµ;ν + ξν;µ , £ξ Γλµν = ξ λ ;µ;ν − ξ κ Rλ µνκ ≡ ξ λ (;µ;ν) − ξ κ Rλ (µν)κ . (2.21)
A vector field satisfying the equation
ξµ;ν + ξν;µ = 0, (2.22)
is called the Killing field. Its existence imposes certain limitations on the
geometry, — existence of an isometry: £ξ gµν = 0 [see (Yano 1955, Eisenhart
1933, Ryan and Shepley 1975)]. The Lie derivative plays an important role
in the monad formalism, as we shall see below; this is why a version of this
formalism well adapted to description of the canonical structure of the field
theory, is called the “Lie-monad formalism” [see (Mitskievich and Nesterov
1981, Mitskievich, Yefremov and Nesterov 1985)].

17
3 The Curvature in Riemannian Geometry
It is convenient to define the concept of curvature in the Riemannian ge-
ometry by introducing the curvature operator [see e.g. (Ryan and Shepley
1975)],
R(u, v) := ∇u ∇v − ∇v ∇u − ∇[u,v] . (3.1)
This operator has the following five properties:
1) It is skew in u and v.
2) It annihilates the metric, R(u, v) g = 0.
3) It annihilates any scalar function, R(u, v) f = 0 (e.g., R(u, v) Γαβγ = 0).
4) This is a linear operator (cf. the axiom 2 of the covariant differentiation).
5) The Leibniz property holds for it (this fact is remarkable since the curva-
ture operator involves the second, not first order differentiation).
The curvature tensor components (which are scalars from the point of
view of the ∇-differentiation) are determined as

R(κ) (λ)(µ)(ν) := θ(κ) · R(X(µ) , X(ν) )X(λ) . (3.2)

These are the coefficients (components) of decomposition of the curvature


tensor with respect to the corresponding basis, and in the general case they
are
(α) (α) () (α)
R(α) (β)(γ)(δ) = X(γ) Γ(β)(δ) − X(δ) Γ(β)(γ) + Γ(β)(δ) Γ()(γ)
(3.3)
() (α) () (α)
−Γ(β)(γ) Γ()(δ) − C(γ)(δ) Γ(β)() .
It is worth remembering that some authors use a definition of curvature with
the opposite sign, e.g. Penrose and Rindler (1984a,b) and Yano (1955) [see
also a table in (Misner, Thorne and Wheeler 1973)]. The structure of the
curvature operator yields a simple property,

Rκ λµν wλ = wκ ;ν;µ − wκ ;µ;ν . (3.4)

It is also easy to show that the standard algebraic identities hold,


Rκλµν = R[κλ][µν] = Rµνκλ ,
(3.5)
νλ
Rκ[λµν] = 0, or equivalently, R ∗ µν =0

(this last identity bears the name of Ricci, but sometimes it is called “al-
gebraic Bianchi identity”). Important role is also played by the differential

18
Bianchi identity,

Rκλ[µν;ξ] = 0, equivalently, R αβ
∗ κλ;β = 0. (3.6)

The Ricci curvature tensor we define as Rµν ≡ Rνµ := Rλ µνλ , cf. Eisen-
hart (1926, 1933), and Misner, Thorne and Wheeler (1973). There are
also introduced the scalar curvature, R := Rαα , the Einstein tensor Gµν =
Rµν − (1/2)gµν R forming the left-hand side of Einstein’s equations, and the
Weyl conformal curvature tensor C,
1
Cκλµν := Rκλµν + gκ[µ Rν]λ + gλ[ν Rµ]κ − Rgκ[µ gν]λ (3.7)
3
(here written in the four-dimensional spacetime manifold). The latter (taken
with one contravariant index, and in a coordinated basis), does not feel a
multiplication of the metric tensor by an arbitrary good function, so it is
conformally invariant. Remember that the Weyl tensor works only for space-
time dimensionality D ≥ 4. When D = 2, Gµν ≡ 0, and for D = 3 the
Cotton tensor (Cotton 1899) is used when the conformal correspondence of
space-times and the classification à la Petrov are considered, see (Garcia et
al. 2004).

3.1 Cartan’s structure equations


In the formalism of Cartan’s exterior differential forms, see (Israel 1970, Ryan
and Shepley 1975), the operation of exterior differentiation is introduced,

d := θ(α) ∇X(α) ∧ ≡ dxα ∇∂α ∧, (3.8)

where ∧ is the discussed above wedge product (the operation of exterior


multiplication), so that dd ≡ 0. The operation d has its most simple repre-
sentation in a coordinated basis (since the Christoffel symbols are symmetric
in their two lower indices). It is however especially convenient to work in an
orthonormalized or Newman–Penrose basis when the connection 1-forms
(α)
ω (α) (β) := Γ(β)(γ) θ(γ) (3.9)

are skew-symmetric (ω (α) (β) = −ω(β) (α) ). In the general case they are found
as a solution of the system of Cartan’s first structure equations,

dθ(α) = −ω (α) (β) ∧ θ(β) , (3.10)

19
and relations for differential of the tetrad metric components,11

dg(α)(β) = ω(α)(β) + ω(β)(α) . (3.11)

Then one has just 24 + 40 equations for determining all the 64 com-
ponents of the connection coefficients, so that the solution of this set of
equations should be unique. It is easy to accustom to find the specific so-
lutions without performing tedious formal calculations; then the work takes
surprisingly short time. Since reference frame characteristics (vectors of ac-
celeration and rotation, and rate-of-strain tensor) below will turn out to be
a kind of connection coefficients, this approach allows to calculate promptly
also these physical characteristics, though for the rate-of-strain tensor the
computation will be rather specific.
In their turn, the curvature tensor components are also computed easily
and promptly, if one applies Cartan’s second structure equations,

Ω(α) (β) = dω(α)(β) + ω (α) (γ) ∧ ω (γ) (β) . (3.12)

Here the calculations are much more straightforward than for the connection
1-forms, and they take less time. It remains only to read out expressions for
the curvature components according to the definition
1
Ω(α) (β) := R(α) (β)(γ)(δ) θ(γ) ∧ θ(δ) (3.13)
2
being in conformity with the curvature operator properties via (3.12). The
Ricci and Bianchi identities read in the language of Cartan’s forms as

Ω(α) (β) ∧ θ(β) = 0 (3.14)

and
dΩ(α) (β) = Ω(α) (γ) ∧ ω (γ) (β) − ω (α) (γ) ∧ Ω(γ) (β) (3.15)
respectively.
The detailed deduction of the second Cartan structure equation (3.12)
from the definitions of curvature (3.3) and the curvature 2-form (3.13) can
be done as follows. When the (skew) symmetry properties are taken into
11
For deduction of these relations consider the following calculations: dg(α)(β) = d(X(α) ·
X(β) ) = θ(µ) ∇X(µ) (X(α) ·X(β) ) = ω (γ) (α) (X(γ) ·X(β) )+ω (γ) (β) (X(α) ·X(γ) ) = ω (γ) (α) g(γ)(β) +
ω (γ) (β) g(α)(γ) = ω(β)(α) + ω(α)(β) .

20
account,
 aswell as the definition of d (3.8), we immediately rewrite Ω(α) (β)
(α) (α)
as dΓ(β)(δ) ∧θ(δ) +ω (α) () ∧ω () (β) +Γ(β)() θ(γ) ∧ω () (γ) where we also used the
definition of 1-form of connection (3.9) (and we shall use it below). Then the
(α)
next identical rewriting yields dω (α) (β) + Γ(β)(δ) ω (δ) () ∧ θ() + ω (α) () ∧ ω () (β) +
(α)
Γ(β)() θ(γ) ∧ ω () (γ) where both terms with Γ identically cancel, and we finally
come to (3.12).

3.2 Einstein’s field equations


From the Bianchi identities (3.6), Rκλµν;ξ + Rκλνξ;µ + Rκλξµ;ν = 0, it follows
after two contractions (in κ and ξ and then with multiplication by g λµ ) that
 
κ 1 κ
Rν − Rδν = 0. (3.16)
2 ;κ

The tensor Gµν := Rνµ − 21 Rδνµ whose covariant divergence thus identically
vanishes, is called Einstein’s conservative tensor. The variational principle
applied to the first term (simply proportional to the scalar curvature R) in
the action integral for Einstein’s field, directly yields this tensor as the left-
hand side of equations; the same principle yields, via the term interpreted
as Lagrangian density of the physical field(s) to be used as source(s) of the
gravitational field, the energy-momentum tensor of the physical field(s) [the
other path to deduce these both tensors is to use the Noether theorem, see,
e.g., Mitskievich (2006)]. All these variation results, naturally, appear with

the corresponding coefficients including, in particular, −g. Historically, the
conservative tensor was first proposed by Einstein to be equated with the
energy-momentum tensor (multiplied by a certain coefficient) to intuitively
come to the equations of gravitational field.

3.3 Parallel Transport and it Analogues


Let us give here a scheme of definition of transports. Denote first as δT v :=
v|Q − v|P →Q the change of a vector (or of some other object) under its
transport between points P and Q [see (Mitskievich, Yefremov and Nesterov
1985)]. Then for the parallel transport
δP v
= ∇u v (3.17)

21
and for the Fermi–Walker transport [cf. Synge (1960)]
δFW v
= ∇u v + η[(v · ∇u u)u − (v · u)∇u u] =: ∇FW
u v (3.18)

where uα = dxα /dλ is the tangent vector to the transport curve (for the
Fermi–Walker transport, it is essential to choose canonical parameter along
the transport curve in such a way that the norm u·u = η −1 would be constant
and non-zero — so the curve itself should be non-null for the Fermi–Walker
transport). Using the definitions (2.7), it is easy to generalize the transport
relations to arbitrary tensors or tensor densities; so for the Fermi–Walker
transport
δFW Ta
= Ta;α uα + ηTa |τσ (uτ ;α uσ − uτ uσ ;α )uα . (3.19)

From (3.18), a definition of the Fermi–Walker connection seems to be di-
rectly obtainable, together with the corresponding covariant FW differentia-
tion operation which should lead also to the FW-curvature [cf. (Mitskievich,
Yefremov and Nesterov 1985)]:

RFW (w, v) := ∇FW FW FW FW FW


u ∇v − ∇v ∇u − ∇[u,v] . (3.20)

The reader may find that in this generalization a difficulty arises: the “deriva-
tive” ∇FW
u does not possess the linearity property. This trouble is however
readily removable if we begin with introduction of another type of derivative,
namely ∇τu (see below) which possesses all necessary properties of a covariant
derivative.

3.4 The Geodesic Equation


We give here the (equivalent to each other) forms of the geodesic equation
(describing parallel transport of a vector along itself) for a canonical (affine)
parameter (then absolute value of the vector does not suffer changes):
∇u u = 0,






u ∧ ∗du = 0,




(3.21)
d2 xα α dxβ dxγ
+ Γ = 0,


dλ2 βγ dλ dλ 





d dxβ 1 σ
dx dx τ 

(g αβ dλ
) = 2
gστ,α dλ dλ
.

22
A geodesic may be, of course, also null (lightlike), but if it is time-like, one has
dλ = ds. If a non-canonical parameter is used, an extra term proportional
to u should be added in the equations (3.21), with a coefficient being some
function of the (non-canonical) parameter.
Below we shall make use of the just mentioned generalization of the FW-
derivative (3.18) along a time-like congruence with a tangent vector field τ
normalized by unity (the monad). It is convenient to denote such a derivative
as ∇τu ,
∇τu T = ∇u T + Ta |µν (τµ;α τ ν − τµ τ ν ;α ) uα B a , (3.22)
where B a is the (coordinated) basis of T : T = Ta B a . Then ∇FW u
u ≡ ∇u . The
definition (3.22) can be also written as
[τ ]
Ta;µ = Ta;µ + Ta |ρσ (τ σ τρ;µ − τ σ ;µ τρ ). (3.23)
This τ -derivative acts as the usual ∇u upon any scalar function f , and it
annuls the vector τ , the metric tensor g, and the construction b = g − τ ⊗ τ
(the three-metric or the projector in the monad formalism):
∇τu f = uf, ∇τu τ = ∇τu g = ∇τu b = 0 (3.24)
(action of a τ -derivative on a tensor is understood in the sense of (3.22); u
is considered as a partial differentiation operator in uf ). The corresponding
τ -curvature operator is denoted as Rτ (w, v) [see (Mitskievich, Yefremov and
Nesterov 1985)].

3.5 The Laplacian and De Rhamian Operators


Alongside with exterior differentiation d which increases the degree of a form
by unity, in Cartan’s formalism another differential operation is also used
which diminishes this degree by unity,
δ := − ∗ d ∗ . (3.25)
In fact this is the divergence operation,
δ(αµh dxµh ) = pαµh ;µ dxh , p = #h + 1, (3.26)
which does not however fulfil the Leibniz property: when applied to a prod-
uct, e.g., of a function and a 1-form, f a, it results in
δ(f a) = f δa − ∗(df ∧ ∗a),

23
while it acts on a function (0-form) simply annihilating it, δf ≡ 0. One
might call this a generalized Leibniz property. Of course, similarly to the
identity dd ≡ 0, there holds also the identity δδ ≡ 0. A combination of both
operations gives the de Rham operator (de-Rhamian),12

4 := dδ + δd ≡ (d + δ)2 (3.27)

(remember divergence of a gradient in Euclidean geometry). A form anni-


hilated by the de-Rhamian, is called harmonic form (remember harmonic
functions and the Laplace operator). In contrast to δ, the de-Rhamian acts
non-trivially on 0-forms, and in contrast to the exterior differential d, it acts
non-trivially on 4-forms too (also in the four-dimensional world). The follow-
ing classification elements of forms are also used: an exact form is that which
is an exterior derivative of another form; a closed form is that whose exterior
differential vanishes (an exact form is always closed, though the converse is
in general not true); a co-exact form is a result of action by δ on some form
of a higher (by unity) degree; a co-closed form is itself identically annihilated
by δ. In (topologically) simplest cases it is possible to split an arbitrary form
into sum of a harmonic, exact, and co-exact forms [see Eguchi, Gilkey and
Hanson (1980), Mitskievich, Yefremov and Nesterov (1985)].
For an arbitrary form α,

4α = αa;µ ;µ dxa − pRσ τ µ ν αµh |τσ dxνh , #a = p. (3.28)

The number of individual indices entering the collective index a of the com-
ponents of the form α, may be any from 0 to 4, while the quantity αa |τσ is
determined by (2.8). The expression (3.28) can be reduced identically to

p(p − 1) στ
4α = (p!)−1 [αµνh;λ ;λ + pRµσ ασνh + R µν αστ h ]dxµνh (3.29)
2
where a = µνh, #a = 2 + #h = p, and p may be equal also to zero.
We compare here explicitly the results of action of de-Rhamiam on forms
of all degrees in a four-dimensional world:

f : 4f = f,µ ;µ , (3.30)

A = Aα dxα : 4A = (Aα;µ ;µ + Rασ Aσ )dxα , (3.31)


12
The same symbol, 4, also denotes the usual Laplacian.

24
1 1
F = Fαβ dxαβ : 4F = (Fαβ;µ ;µ + 2Rασ Fσβ + Rστ αβ Fστ )dxαβ , (3.32)
2 2
T = 3!1 Tαβγ dxαβγ : 4T = 3!1 (Tαβγ;µ ;µ + 3Rασ Tσβγ + 3Rστ αβ Tστ γ )dxαβγ

= (Ã;µ σ α
α;µ + Rα Ãσ ) ∗ dx ,
(3.33)
1 1
V = Vαβγδ dxαβγδ : 4V = Vαβγδ;µ ;µ dxαβγδ = f˜;µ ;µ
∗ 1. (3.34)
4! 4!
From the last two expressions it is clear that there exists a symmetry with
respect to an exchange of a basis form to its dual conjugate with a simultane-
ous dual conjugation of components of the form [compare (3.30) and (3.34),
(3.31) and (3.33)]. Here we made use of obvious definitions
1 1
Ãδ = − Tαβγ E αβγδ , f˜ = − Vαβγδ E αβγδ ,
3! 4!
alongside with the easily checkable identity,

(2Rασ Eσβγδ + 3Rστ αβ Eστ γδ )E αβγδ ≡ 0.

References
Choquet-Bruhat, Y., DeWitt-Morette, C., and Dillard-Bleick, M. (1982)
Analysis, Manifolds, and Physics (Amsterdam: North-Holland).
Cotton, E. (1899) Ann. Fac. Sci. Toulouse II 1, 385.
DeWitt, B.S. (1965) Dynamical Theory of Groups and Fields (New
York: Gordon and Breach).
Eguchi, T., Gilkey, P.B., and Hanson, A.J. (1980) Phys. Reports 66,
213.
Eisenhart, L.P. (1926) Riemannian Geometry (Princeton, N.J.: Prince-
ton University Press).
Eisenhart, L.P. (1933) Continuous Groups of Transformations (Prince-
ton, N.J.: Princeton University Press).
Eisenhart, L.P. (1972) Non-Riemannian Geometry (Providence, RI:
American Mathematical Society).

25
Garcı́a, A.A., Hehl, F.W., Heinicke, Ch., and Macı́as, A. (2004) Class.
Quantum Grav. 21, 1099.
Islam, J.N. (1985) Rotating fields in general relativity (Cambridge,
Cambridge University Press).
Israel, W. (1970) Commun. Dublin Inst. Adv. Studies A 19, 1.
Lanczos, C. (1938) Ann. Math. 39, 842.
Lanczos, C. (1962) Revs. Mod. Phys. 34, 379.
Lovelock, D. (1971) J. Math. Phys. 12, 498.
Misner, C.W., Thorne, K.S., and Wheeler, J.A. (1973) Gravitation (San
Francisco: W.H. Freeman).
Mitskievich, N.V. (1969) Physical Fields in General Relativity (Moscow:
Nauka). In Russian.
Mitskievich, N. (2006) Relativistic Physics in Arbitrary Reference Frames
(New York: Nova Science Publishers, Inc.).
Mitskievich, N.V., and Merkulov, S.A. (1985) Tensor Calculus in Field
Theory (Moscow: Peoples’ Friendship University Press). In Russian.
Mitskievich, N.V., and Nesterov, A.I. ((1981) Exper. Techn. der Physik
29, 333.
Mitskievich, N.V., Yefremov, A.P., and Nesterov, A.I. (1985). Dy-
namics of Fields in General Relativity (Moscow: Energoatomizdat) In
Russian.
Penrose, R., and Rindler, W. (1984a) Spinors and Space-Time. Vol. 1.
Two-Spinor Calculus and Relativistic Fields (Cambridge: Cambridge
University Press).
Penrose, R., and Rindler, W. (1984b) Spinors and Space-Time. Vol.
2. Spinor and Twistor Methods in Space-Time Geometry. (Cambridge:
Cambridge University Press).
Ryan, M.P., and Shepley, L.C. (1975) Homogeneous Relativistic Cos-
mologies (Princeton, N.J.: Princeton University Press).

26
Sachs, R.K., and Wu, H. (1977) General Relativity for Mathematicians
(New York, Heidelberg, Berlin: Springer).
Schmutzer, E. (1989) Grundlagen der Theoretischen Physik (Mannheim,
Wien, Zürich: BI-Wissenschaftsverlag), Teile 1 & 2.
Schmutzer, E. (2004) Projektive Einheitliche Feldtheorie mit Anwen-
dungen in Kosmologie und Astrophysik (Frankfurt am Main: Wis-
senschaftlicher Verlag Harry Deutsch GmbH). Mit einem Anhang von
A.K. Gorbatsievich.
Schouten, I.A., and Struik, D.J.(1935) Einführung in die neueren Meth-
oden der Differentialgeometrie. Vol. 1 (Gronongen: Noordhoff).
Schouten, I.A., and Struik, D.J. (1938) Einführung in die neueren
Methoden der Differentialgeometrie. Vol. 2 (Gronongen: Noordhoff).
Schutz, B.F. (1980) Geometrical Methods of Mathematical Physics (Cam-
bridge: Cambridge University Press).
Synge, J.L. (1960) Relativity: The General Theory (Amsterdam: North-
Holland).
Synge, J.L. (1965) Relativity: The Special Theory (Amsterdam: North-
Holland).
Trautman, A.(1956) Bulletin de l’Académie Polonaise des Sciences,
Série de sciences math., astr. et phys. IV, 665 & 671.
Trautman, A. (1957) Bulletin de l’Académie Polonaise des Sciences,
Série de sciences math., astr. et phys. V, 721.
Vladimirov, Yu.S., Mitskievich, N.V., and Horsky, J. (1987) Space,
Time, Gravitation (Moscow: Mir). Revised English translation of a
Russian edition of 1984 (Moscow: Nauka).
Westenholz, C. von (1986) Differential Forms in Mathematical Physics
(Amsterdam: North-Holland).
Wheeler, J.A. (1962) Geometrodynamics (New York: Academic Press).
Yano, K. (1955) The Theory of Lie Derivatives and Its Applications
(Amsterdam: North-Holland).

27
4 Deduction of the Schwarzschild solution
First to use the ideas of symmetry to determine the general form of metric
to describe the spherically symmetric gravitational field (spacetime) without
initially excluding dependence of metric on the t variable. The variable r is
introduced from the spherical symmetry (in the angular sense) like this was
done by Synge.
ds2 = e2α(r,t) dt2 − e2β(r,t) dr2 − r2 dϑ2 + sin2 ϑdφ2 ;


θ(0) = eα dt, θ(1) = eβ dr, θ(2) = rdϑ, θ(3) = r sin ϑdφ;

dθ(0) = α0 e−β θ(1) ∧ θ(0) , dθ(1) = β̇e−α θ(0) ∧ θ(1) , dθ(2) = 1r e−β θ(1) ∧ θ(2) ,
dθ(3) = 1r e−β θ(1) ∧ θ(3) + cotr ϑ θ(2) ∧ θ(3) ;

ω (0) (1) = α0 e−β θ(0) + β̇e−α θ(1) , ω (2) (1) = 1r e−β θ(2) , ω (3) (1) = 1r e−β θ(3) ,
ω (3) (2) = cotr ϑ θ(3) ;
 .
0 α−β 0 −α−β (1)
e−α−β θ(0) ∧ θ(1) ,
(0)
 (0) β−α
Ω (1) = α e e θ ∧ θ + β̇e
−2β
Ω(0) (2) = ω (0) (1) ∧ ω (1) (2) = −α0 e r θ(0) ∧ θ(2) ,
−2β
Ω(0) (3) = −α0 e r θ(0) ∧ θ(3) ,
−2β −α−β
Ω(2) (1) = −β 0 e r θ(1) ∧ θ(2) − β̇ e r θ(0) ∧ θ(2) ,
−2β −α−β
Ω(3) (1) = −β 0 e r θ(1) ∧ θ(3) − β̇ e r θ(0) ∧ θ(3) ,
Ω(3) (2) = r12 1 − e−2β θ(3) ∧ θ(2) .


Please check the following properties: ω (0) (i) = +ω (i) (0) , ω (i) (j) = −ω (j) (i) , and
the same for Ω’s. These properties for ω’s in fact follow from the relations
(3.11) given in the Topics of Riemannian Geometry, and they then apply
only in the case of constant components of g(α)(β) , while for Ω’s they come
from the general properties of the curvature tensor and do apply without any
restrictions.

Thus h   β−α . i −α−β


(0) 0 α−β 0
R (1)(1)(0) = α e − β̇e e ,
−2β
R(0) (2)(0)(2) = R(0) (3)(0)(3) = −α0 e r ,
−2β
R(2) (1)(2)(1) = R(3) (1)(3)(1) = β 0 e r ,
−α−β
R(2) (1)(2)(0) = R(3) (1)(3)(0) = β̇ e r ,
R(3) (2)(3)(2) = r12 1 − e−2β ;

28
h  . i
0 α−β 0 −2β
e−α−β − 2α0 e r ,
 β−α
R(0)(0) = − α e − β̇e
h 0  . i −2β
R(1)(1) = α0 eα−β − β̇eβ−α e−α−β − 2β 0 e r ,
−α−β
R(0)(1) = −2β̇ e r ,
−2β
R(2)(2) = R(3)(3) = (α0 − β 0 ) e r − 1
1 − e−2β .

r2

In a vacuum, the scalar curvature R = 0, thus Einstein’s equations re-


duce to R(α)(β) = 0. When we take (α)(β) ⇒ (0)(1), we see that β = β(r).
Einstein’s equations now take the form
0 −2β
R(0)(0) = − α0 eα−β e−α−β − 2α0 e r = 0,
0 −2β
R(1)(1) = α0 eα−β e−α−β − 2β 0 e r = 0,

−2β
R(2)(2) = R(3)(3) = (α0 − β 0 ) e r − r12 1 − e−2β = 0.


From R(0)(0) + R(1)(1) = 0 we see that α + β = f (t) (we already know


that β(r), i.e. β does not depend on t. Now, from ds2 we see that the still
not determined function f (t) can be simply eliminated by merger with the
variable t of which now depends nothing. This is equivalent to put f (t) = 0
in our final results. Thus we simply put α + β = 0. Substituting β = −α
into the equation R(2)(2) = 0, we come to the simple equation only for α:

(e2α )0 = 1−er , which yields e2α = 1 − Cr , C being the only survived integra-
tion constant. It is now easy to show that the full set of Einstein’s equations
in this case is completely solved with this α, and the solution in the form of
ds2 is
2M −1 2
ds2 = (1 − 2M 2 2 2 2 2

r
)dt − (1 − r
) dr − r dϑ + sin ϑdφ . M easily can be
shown to be the central mass (in the units involving the Newtonian gravita-
tional constant) creating this gravitational field. The very solution bears the
name “Schwarzschild solution” having been obtained by Karl Schwarzschild
after few days since Einstein’s paper containing the final form of his gravita-
tion theory appeared in December of 1915.

5 Interpretation of the Schwarzschild metric:


the first steps
The further steps: 1. In the region where r  M , consider the geodesic
equation in non-relativistic limit and weak field approximation. A compari-

29
son with the Newtonian theory then gives the interpretation of M .
2. Beginning with r > 2M , consider exact application of the geodesic equa-
tion to the radial free fall of a test particle on the Schwarzschild centre. Using
s as the proper time parameter, calculate duration of the fall till r = 2M
and till r = 0 (the field singularity). Then change to the variable t as a
“time” parameter and show that the fall duration from r > 2M to r = 2M
is infinite. A discussion of these conclusions.
d β σ dxτ
Let us take the geodesic equation in the form dλ (gαβ dx

) = 12 gστ,α dx
dλ dλ
.
Obviously, two conservation laws follow from it for α = 0 and α = 3 (there is
one more independent conservation law with a mixture of φ and ϑ completing
the picture of absence of angular momentum as one of the cases of conserva-
tion, but we use the “angular” laws only as a justification of the radial fall
treatment). The law corresponding to α = 0 has its exact expression as

dt
gtt = E, (5.1)
ds
being the first integral of motion named “kinetic energy of the test particle
per unit of its rest mass” (please don’t confuse E with !).
The first step in our consideration is here to obtain the physical interpre-
tation of the constant M in ds2 . To this end we take the radial component
of the geodesic equation, that is
"    2 #
  2
d dr 1 0 dt 0 dr
grr = gtt + grr
ds ds 2 ds ds

(in the case of a radial fall) which identically reads

d2 r 1
2
= − gtt0 . (5.2)
ds 2
All this is done exactly, without approximations.
It is easy to see that in the non-relativistic limit ds ⇒ dt; at the same
time, we explicitly differentiate gtt with respect to r. This finally gives
d2 r
dt2
= −M r2
, being precisely the 2nd Newtonian law when the mass M is
attracting a test mass falling onto it (in fact, M = Gm where G is the
Newtonian gravitational constant, and m is value of the central mass in its
usual units). You see: the approximative calculation was used at the final
step, and only once.

30
The second step of our consideration of the Schwarzschild metric is to
reveal the main property of its horizon, r = 2M . We here use for another
dt
time the (α = 0)-conservation law, g00 ds = E. Writing down ds2 , dividing it
by ds2, and taking into account the energy conservation law, we come exactly
2
to dr
ds
= E 2 − gtt , so that for a particle falling downward in the sense of r,

dr
ds = − p , (5.3)
E 2 − gtt

which can be considered as another first integral of motion, tohether with


(5.1). To simplify the further calculation, let us take (only in this second
step) E = 1 which means that the falling particle starts falling from spatial
infinity with zero velocity (the state of rest), though we, of course, may follow
its fall beginning with any final point outside r0 > 2M where this particle
already had non-zero velocity towards the origin. The interval of the proper
time s between the particle R rwas passing the point r0 and then came to some
1/2 dr
r√
minor value of r, is ∆s = r0 2M . The result of integration is obvious, and
it is finite both for r = 2M and for r = 0. But this is true only for the proper
time of the falling particle. If we would try to describe the time of fall in
terms of the variable t, the result will be infinite already for r = 2M . Then
what occurs with this r = 2M ?!
Recall that the physical sense of coordinates is determined by their sig-
nature. Due to its definition used in these notes, the sign + marks the
timelike dimension and the sign −, the other three spatial dimensions. But
in Schwarzschild’s ds2 t is timelike and r, spacelike, only outside the hori-
zon r0 = 2M , while inside the horizon t becomes spacelike and r, timelike;
moreover, on the horizon these two coordinated completely lose their math-
ematical meaning, and there is nothing in the Schwarzschild metric what
could indicate in which direction is the future and in which is the past of this
timelike r inside the horizon, even we would somehow define what mean past
and future for the variable t outside it. However there is a proper way to
do this introducing Novikov’s coordinates simultaneously in whole spacetime
while beginning with the usual Schwarzschildean coordinates. We shall not
realize this program here in its complete form, but we shall now come to
a sufficiently comprehensive hint how to do this, beginning with somewhat
simpler synchronous coordinates.

31
6 The synchronous coordinates
Definition: Synchronous coordinates are such in which the time coordinate
T lines represent a non-rotating timelike geodesic congruence parameterized
by proper time along the lines, while the spatial coordinates lines belong to
spacelike hypersurfaces perpendicular to the T -congruence. (Such congru-
ences are also called “normal congruences” since absence of rotation is the
necessary and sufficient condition of existence of global hypersurfaces orthog-
onal — “normal” — to these congruences.) Synchronous coordinates do not
need any special kind of geometry of the manifolds in which they are built, so
that one can construct synchronous coordinates in every pseudo-Riemannian
spacetime.
Let us first make it clear what mean the concepts of coordinates and of
coordinates’ lines. The first corresponds to independent variables existing
in the manifold under consideration. If we choose one of these variables
(one coordinate) and consider it running through all values it can take, while
fixing values of all other variables, we describe thus a line of this coordinate
marked by all constant values which take the other coordinates on this line.
If this line goes through the origin (where all coordinates, including that to
which belongs this line, are equal to zero; this concept of the origin is in fact
not so simple as it could seem initially), it is marked (labeled) by n − 1 zeros
(n is the number of all independent variables on this manifold, thus it is
n-dimensional), and this special line is then called the axis of the coordinate
which we did not fix in this paragraph. It is interesting that when we consider
a coordinates’ transformation, the lines of coordinates become automatically
transformed in a way which deserves to be named “complementary” to the
way of transformation of the very coordinates. An example seems to me here
to be appropriate. Let us consider a two-dimensional spacetime {t, x} and
apply to these coordinates the usual Galilean transformation (first written,
by the way, by Newton): t0 = t, x0 = x − vt. The question is: how will then
change the axes of coordinates? By the definition of the axes, we have to take
for the axis of x0 the condition t0 = 0, but since t0 = t, this is at the same time
the condition by which the axis of x is determined, consequently, the axes
of x and of x0 coincide. As to the axis of t0 , we have to put x0 = 0; clearly,
this gives the following equation to find this axis: x = vt, thus the axis of t0
differs from that of t, and we did really come to a certain complementarity
(about which I never heard or read before). It remains to be added only
that the t0 -lines here exactly coincide with the new time coordinate’s lines

32
appearing after the usual Lorentz transformation with the same velocity v
as in the Galilean transformation several lines above.
We now turn to of synchronous coordinates in the Schwarzschild space-
time. We already know that the timelike (outside the horizon, though there
is a theorem asserting that the timelike, spacelike or null behaviour of a
geodesic cannot change) radial geodesic of a particle freely falling towards
the origin, is determined by the first integrals (5.1) and (5.3). They can be
combined (to exclude ds) as
Edr
dt + p = 0, (6.1)
gtt E 2 − gtt
a relation which equivalently describes radial geodesics of the free fall, like
we considered them in the beginning of section 2, though in another manner.
But we need these geodesics to the end of introducing the coordinate lines of
T in the synchronous system which now is under construction. How we come
to these coordinate lines? Since the relations between coordinates T , R, on
the one hand, and t, r, on the other, have to be of the type of T = T (t, r)
and R = R(t, r), and recalling the general definition of coordinate lines in
the second paragraph of this section (that coordinate lines of the coordinate
T are determined by fixation of the alternative coordinate, here R: dR = 0),
we have to consider the equation dR = ∂R ∂t
dt + ∂R
∂r
dr = 0 as an analogue of
(6.1). Therefore let the definition of the new coordinate R be13
E 2 dr
dR = Edt + p , (6.2)
gtt E 2 − gtt
an equation having as its left- and right-hand sides only total differentials
(which is of great importance), and fixation of R, i.e. dR = 0, then really
leads to determination of timelike geodesic lines which we need as coordinate
lines of T . Why not construct dT (t, r) in a similar manner, that is dT =
dt + F (r)dr (though we now are not concerned in geodesic properties of
coordinate lines of R), and then determine the unknown function F (r)? The
way to determine it is, moreover, clear: the mutual orthogonality of dT and
dR and the unit property of the T T -component of the new synchronous
metric. Look: dR · dT = GRT = 0, and then dT · dT = GT T = 1 (under G we
13
We introduced here and in (6.3) an extra constant factor E in dR and dT which is
necessary for obtaining GT T = 1 and for achieving more symmetry in these expressions,
as well as for having a simple form of (6.4).

33
now understand the new metric in T , R, ϑ, and φ coordinates; the angular
coordinates remain without any change in the synchronous coordinates since
they were orthogonal to t and r, hence also to T and to R). The calculations
are extremely simple:
EF (r)
(1) dR · dT = dt · dt + √ 2
dr · dr where dt · dt = g tt = 1/gtt , dr · dr =
gtt E −gtt
g rr = −gtt , while dt · dr = 0. Therefore in order to have dR · dT = 0, we take
p
E 2 − gtt
dT = Edt + dr. (6.3)
gtt
(2) dT · dT = 1.
(3) A non-trivial fact is that it is easy to express the old radial variable r in
terms of the combination of new coordinates T and R,
dr
d(R − T ) = p . (6.4)
E 2 − gtt
This relation is immediately integrable and gives the concrete dependence
r(R − T ), at least implicitly. As to the last two variables ϑ and φ which
did not change, the metric coefficients corresponding to them now involve
r2 (R − T ). Finally, the Schwarzschild spacetime is described in synchronous
coordinates by the metric
E 2 dR2
ds2 = dT 2 − 2 2 2 2

2M
− r (R − T ) dϑ + sin ϑdφ . (6.5)
E 2 − 1 + r(R−T )

From these expressions we see that in the synchronous coordinates there re-
main no traces of the singularity earlier detected on the horizon in Schwarz-
schildean coordinates. This singularity only marked the inadmissible charac-
ter of the usual Schwarzschildean coordinates on the horizon, without being
in any sense related to the physical properties of the Schwarzschild solu-
tion; in the latter, there exists only the singularity at r = 0. Naturally,
one might choose the constant E in such a way that there would appear
zero in the denominator at some r, but this would simply be an inadequate
restricted choice of the T -congruence in constructing the synchronous coor-
dinates, nothing more. As an exercise, the case of E = 1 is proposed in which
you have to perform all calculations up to the final results which will then
be quite simple; then to draw a spacetime diagram in the T R-plane giving
there some lines of constant r, including r = 0 (singularity) and r = 2M

34
(horizon). Look to which types pertain these two lines (timelike, spacelike,
or null). Using as a definition of null (lightlike) lines ds2 = 0, draw in this
diagram (approximately) from some convenient point on the horizon taken
as a vertex, both generatrices of the light cone (to the future as well as to
the past). Try to imagine how can go a timelike world line of a test parti-
cle from −∞ in T to the future, and determine the region from where this
particle (say, a spaceship) can escape hitting the singularity, and from where
it cannot. How could you relate this behaviour of a particle to the gravita-
tional collapse phenomenon? Finally, answer the question: if we would like
to compare the “point-like” object at the origin (r = 0) of the Schwarzschild
spacetime with some type of a particle, which type would be acceptable —
that of a usual massive particle, of a photon, or of a tachyon? — Defini-
tively, the source of the Schwarzschild field is a tachyon with the
mass M , moving with infinite 3-velocity, while always “remaining”
at r = 0 (especially clear when we use a Penrose diagram).14

7 The Kerr solution “deduced”


Here we shall introduce the famous Kerr black hole solution using our gener-
alization of the so-called “falling box” method first published by Sommerfeld
(with a reference to Lenz) where the Schwarzschild solution was “deduced”.
From 1949 till 1984 that was the only metric “deduced” by this extrav-
agant method. Then I have invented similar, though more sophisticated
approaches to semi-intuitively come first to the Kerr solution, then to the
gravitational-electric Reissner–Nordström, and finally to the most compli-
cated three-parameter Kerr–Newman solution [see Vladimirov, Mitskievich
and Horský (1987)]. Now, with the example of the Kerr spacetime, we shall
make acquaintance with this Sommerfeld–Lenz method severely criticized by
quite serious physicists, but after the “deduction” with its help of all four
black hole metrics, safely survived at least for being used in the popular
literature on gravitation.
We begin with the Minkowski spacetime supposedly applicable far from
14
An addition of as small as you like charge (or angular momentum, or both) to
Schwarzschild’s centre immediately results in the change of the spacelike (r = 0)-line
(of the “point-like” singularity) to timelike and, consequently, of the tachyonic particle to
a usual massive one, with the corresponding complete instantaneous restructuring of the
Penrose or synchronous coordinates’ diagram.

35
localized sources of the gravitational field. Thus, taking for that “far” the
mark of infinity ∞, we may write (in fact, using the spherical coordinates)
(0) 2 (1) 2 (2) 2 (3) 2
ds2∞ = dt2 − dr2 − r2 dϑ2 + sin2 ϑdφ2 = θ∞

− θ∞ − θ∞ − θ∞ (7.1)
(α)
from where the structure of all four basis covectors θ∞ becomes obvious.
A transition to a non-uniformly rotating reference frame is done by locally
applying Lorentz transformations so that every point has its own speed of
motion directed towards an increasing angle φ. The absolute value of this
velocity is a function V which depends, generally speaking, on both coor-
dinates r and ϑ. Such a Lorentz transformation is not equivalent to the
transformation of coordinates in the domain being studied (in practice, this
domain is the whole of space), but is limited only to the transformation of
the basis at each point. Thus, we have
  √ 
(0) (3) (1) (2)
θ̃(0) = θ∞ − V θ∞ / 1 − V 2 , θ̃(1) = θ∞ , θ̃(2) = θ∞ , 

(3) (0)
 √ (7.2)
θ̃(3) = θ∞ − V θ∞ / 1 − V 2 . 

Now let the box with the observer be released from infinity. In this case
we can write a new basis in which time has slowed down, and the lengths
(α)
in radial direction have shortened. This is equivalent to substitution of θ∞
first in (7.2) and then the resulting expressions into
√ √
θ0(0) = θ̃(0) 1 − v 2 , θ0(1) = θ̃(1) / 1 − v 2 , θ0(2) = θ̃(2) , θ0(3) = θ̃(3) . (7.3)
We have thus assumed that the observer makes his measurements in the
rotating frame and notices the relativistic changes in his observations. Now
let us do the reverse local Lorentz transformations with respect to (7.2) this
time applied to the basis (7.3):
 √
θ(0) = θ0(0) + V θ0(3)  /√1 − V 2 , θ(1) = θ0(1) , θ(2) = θ0(2) ,

(7.4)
θ(3) = θ0(3) + V θ0(0) / 1 − V 2 .
We now insert into(7.4) the θ0(α) basis which is expressed in terms of the
θ̃(α) from (7.3), and then write this expression in terms of θ̃(α) from (7.2).
We postulate that the resulting basis (7.4) remains orthonormalized. A few
manipulations yield
(0) 2 (1) 2 (2) 2
 
v2
ds = 1 − 1−V 2 θ∞ − (1 − v 2 )−1 θ∞ − θ∞
2
 
(3) 2
(7.5)
v2 V 2 2V v 2 (0) (3)
− 1 + 1−V 2 θ ∞ + 1−V 2
θ ∞ θ ∞ .

36
The principle of correspondence with Newton’s theory presupposes, as
this was assumed by Sommerfeld in his falling-box deduction of the Schwarz-
schild solution, that
2M
gtt = 1 − ≡ 1 + 2ΦN , (7.6)
r
ΦN being the Newtonian gravitational potential considered there as a spher-
ically symmetric solution of the Laplace equation. Here this should be also
a solution of the Laplace equation, but now under the rotational symmetry
and not spherical. Therefore it is now worth considering oblique spheroidal
coordinates in flat space. These coordinates, r, ϑ, and φ, are defined as

x2 + y 2 z 2
x + iy = (r + ia)eiφ sin ϑ, z = r cos ϑ, + 2 = 1. (7.7)
r 2 + a2 r
We know that in the spherical coordinates ∆(1/r) = 0 when r 6= 0, and this
equality holds under any translation of coordinates. Let this translation now
be purely imaginary and directed along the pz axis, i.e. x → x, y → y, and
z → z −ia. Then it is easy to find that r = x2 + y 2 + z 2 → (r2 −a2 cos2 ϑ−
2iar cos ϑ)1/2 = r − ia cos ϑ. From here the expression for Newton’s potential
in oblique spheroidal coordinates follows,
M M −M r
ΦN = → ΦN = −Re = 2 , (7.8)
r r − ia cos ϑ r + a2 cos2 ϑ
since the Laplace equation is satisfied simultaneously by both the real and
the imaginary parts of the translated potential. Hence we can get with the
help of (7.6) (the abbreviation ρ2 = r2 + a2 cos2 ϑ is henceforth used)

v2 2M r
2
= 2 . (7.9)
1−V ρ
We determine the velocity V using a model of rotating ring of some ra-
dius r0 for the source of the Kerr field, this ring being stationary relative to
1/2
the rotating reference frame (7.3). On the one hand, V = Ω (x2 + y 2 ) =
1/2
(r2 + a2 ) Ω sin ϑ, corresponds to the well-known relationship between an-
gular and linear velocities in spherical coordinates. On the other hand, it is
clear that the reference frame cannot rotate as a rigid body in all the world,
otherwise the frame wouldn’t be extendible beyond the “light cylinder” as
we dropped our box from infinity. Therefore the angular velocity Ω has also

37
to be a function of position. The ring lies naturally in the equatorial plane,
so that the angular momentum along the z axis is
q
L = mV r02 + a2 , (r = r0 , ϑ = π/2). (7.10)

We now introduce an important hypothesis which establishes a connection


between the angular momentum and the Kerr parameter a, which is also
characteristic for spheroidal coordinates (7.7), namely we put a = L/m These
last three statements yield Ω(r = r0 , ϑ = π/2) = a (r02 + a2 ). If we now add a
second hypothesis, that the field is independent of the choice of the ring radius
(depending only on its angular momentum), then we get Ω = a/(r2 + a2 ),
and finally
V = a(r2 + a2 )−1/2 sin ϑ. (7.11)
(α)
It only remains for us to choose the expression for a basis θ∞ which would
correspond to the assumed rotational symmetry (i.e., to the oblique spheroidal
coordinates). The reader may substitute differentials of the coordinates x, y,
and z from (7.7) him/herself into the Minkowski squared interval, hence get-
ting a quadratic form with a non-diagonal term. This term, which contains
the product drdφ, can be excluded by a simple change of the azimuth angle:
dφ → dφ + a(r2 + a2 )−1 dr (a total differential), thus leading to a diagonal
quadratic form. If now the square roots of the separate summands are taken,
(α)
we get the final form of the initial basis θ∞ :
q )
(0) (1) ρ2
θ∞ = dt, θ∞ = r2 +a2 dr,
(2)
p (3) √ (7.12)
θ∞ = ρ2 dϑ, θ∞ = r2 + a2 sin ϑdφ.

A mere substitution of these expressions into (7.5) yields the standard


form of the Kerr metric in terms of the Boyer–Lindquist coordinates,
  2
ds = 1 − ρ2 dt2 − ρ∆ dr2 − ρ2 dϑ2
2 2M r
 2 2
 2
(7.13)
− r2 + a2 + 2M raρ2sin ϑ sin2 ϑdφ2 + 2 2M raρ2sin ϑ dtdφ.

We remind that
ρ2 = r2 + a2 cos2 ϑ; (7.14)
the notations
∆ = r2 − 2M r + a2 (7.15)

38
and
Σ2 = (r2 + a2 )2 − a2 ∆ sin2 ϑ (7.16)
are also frequently used (in the Kerr–Newman case ∆ = r2 − 2M r + a2 + Q2
where Q is electric charge of the Kerr–Newman centre given in the units of
length, like this is done with M as we have seen in the deduction of the
Schwarzschild metric).
We give now a transparent hint to three anagrams of the Kerr metric
with different combinations of three terms containing dt and/or dφ. First,
these terms can be combined in two constructions giving a part (for two di-
mensions) of the complete four-dimensional orthonormal basis with rotation
in the Kerr spacetime,

2M a2 r sin2 ϑ
   
2M r 2 2 2
1− 2 (dt + ωdφ) − r + a + 2 sin2 ϑdφ2 (7.17)
ρ ρ − 2M r

with
2M ar sin2 ϑ
ω= . (7.18)
ρ2 − 2M r
Second, there is another combination of the same terms into a part of the
pseudo-rotational orthonormal basis,

ρ2 ∆ 2 Σ2 sin2
dt − (dφ − wdt)2 (7.19)
Σ2 ρ2

where15
2M ar
w=. (7.20)
Σ
And finally, dt-dφ terms can be combined into a very simple construction

∆ 2
2 sin2 ϑ  2 2
r + a2 dφ − adt

2
dt − a sin ϑdφ − 2
(7.21)
ρ ρ

including both rotation and pseudo-rotation [see Mitskievich and Vargas


Rodrı́guez (2005) about them]. Similar situation occurs in the general (three-
parametric) Kerr–Newman case too.
15
Please don’t confuse two very different letters from very different alphabets, ω and w!

39
8 Checking the Kerr solution
Now we have to verify if the metric (7.13) really satisfies Einstein’s equations
in a vacuum. To this end we have first to find the connection 1-forms using
the 1st Cartan structure equations. As the simplest one, we choose the
orthonormal basis from (7.13) and (7.21), so that

θ(0) = ρ∆ (dt − a sin2 ϑdφ), θ(1) = √ρ∆ dr,
(8.1)
θ(2) = ρdϑ, θ(3) = sinρ ϑ [(r2 + a2 )dφ − adt];

consequently,

a 1 r 2 + a2 a sin ϑ (3)
dφ = √ θ(0) + θ(3) , dt = √ θ(0) + θ . (8.2)
ρ ∆ ρ sin ϑ ρ ∆ ρ

We now write down the direct d-differentiation of the basis covectors,


and, in parallel, the respective 1st Cartan equations:
M (r2 −a2 cos2 ϑ)−a2 r sin2 ϑ (1)
dθ(0) = √
ρ3 ∆
θ ∧ θ(0)
2

− aρ3 sin ϑ cos ϑθ(2) ∧ θ(0) − 2aρ3 ∆ cos ϑθ(2) ∧ θ(3) , (8.3)
dθ(0) = −ω (0) (1) ∧ θ(1) − ω (0) (2) ∧ θ(2) − ω (0) (3) ∧ θ(3) ;
2
dθ(1) = − aρ3 sin ϑ cos ϑθ(2) ∧ θ(1) ,
(8.4)
dθ(1) = −ω (1) (0) ∧ θ(0) − ω (1) (2) ∧ θ(2) − ω (1) (3) ∧ θ(3) ;

dθ(2) = r ρ3∆ θ(1) ∧ θ(2) ,
(8.5)
dθ(2) = −ω (2) (0) ∧ θ(0) − ω (2) (1) ∧ θ(1) − ω (2) (3) ∧ θ(3) ;
√ 2 2
dθ(3) = r ρ3∆ θ(1) ∧ θ(3) + r ρ+a
3 cot ϑθ(2) ∧ θ(3) + 2 ar ρsin
3
ϑ (1)
θ ∧ θ(0) ,
(8.6)
dθ(3) = −ω (3) (0) ∧ θ(0) − ω (3) (1) ∧ θ(1) − ω (3) (2) ∧ θ(2) .
We see that here the problem of algebraically solving the system of equations
is somewhat more complicated than in the deduction of the Schwarzschild
solution. √Of course, there are present similar obvious terms of the type
dθ(2) = r ρ3∆ θ(1) ∧ θ(2) where it is clear that θ(2) can pertain only to the
connection 1-form and not to the basis covector following it in the exte-
rior product since otherwise we should encounter ω (2) (2) which is identically
equal to zero. Hence there should be ω (2) (1) (multiplied by θ(1) which has no
other choice), and this ω (2) (1) should be proportional to θ(2) , the coefficient

40
of proportionality being quite obvious. However even here we run into an
additional term in ω (2) (1) , that containing θ(1) and giving no contribution
to the equation (8.5); this term being important in (8.4) where the term
which was the only one being present in (8.5), loses its significance (simply
disappears). To this collection of terms in our problem here two new trios
of terms now are added: one trio being marked with the indices 0, 2, 3 and
another, by 0, 1, 3 (in all mutually non-reducible successions). This could
probably be not quite obvious, but such a situation (with each single trio) is
typical in calculations of the connection 1-forms, and checking this rule via
the practical use easily shows that we deal with them correctly. We now find
that in all but one (ω (1) (2) ) connection 1-forms there now appear new terms
(for the first trio, in the connection 1-forms where two of the indices 0, 2,
and 3 are involved; similarly for the second trio with two indices of 0, 1, and
3). Let us see how they have to be treated. In the first trio, we write in
the corresponding terms still not determined constant coefficients A, B, C
(and we work similarly with the second trio introducing other three constant
coefficients, say, P, Q, R):
2 2 2 ϑ)−a2 r sin2 ϑ

ω (0) (1) = M (r −a cos √
ρ3 ∆
θ (0)
+ P ar sin ϑ (3) 
ρ 3 θ ,



a2 a ∆
(0) (0) (3)

ω (2) = − ρ3 sin ϑ cos ϑθ + A ρ3 cos ϑθ , 




(0) a ∆ (2) ar sin ϑ (1)

ω (3) = B ρ3 cos ϑθ + Q ρ3 θ , 
2
√ (8.7)
ω (1) (2) = − aρ3 sin ϑ cos ϑθ(1) − r ρ3∆ θ(2) , 



ω (1) (3) = − r ρ3∆ θ(3) + R ar ρsin ϑ (0)

3 θ , 




(2) r 2 +a2 (3) a ∆ (0)

ω (3) = − ρ3 cot ϑθ + C ρ3 cos ϑθ 

(we gave here the complete set of the connection 1-forms, though with the
already mentioned terms containing still not determined coefficients). These
constant coefficients are now determined by substitution of these connection
1-forms into the 1st structure equations written in (8.3)-(8.6). A comparison
with dθ(α) found there in parallel lines yields two sets of algebraic equations
(one for each trio) whose solutions are A = −B = C = 1, P = Q = R = −1.

41
Thus the final form of ωs reads
M (r2 −a2 cos2 ϑ)−a2 r sin2 ϑ (0)

ω (0) (1) = √
ρ3 ∆
θ − ar ρsin
3
ϑ (3)
θ , 



2
− aρ3 sin ϑ cos ϑθ(0) + a ρ3∆ cos ϑθ(3) ,

ω (0) (2) = 




ω (0) (3) − a ρ3∆ cos ϑθ(2) − ar ρsin ϑ (1)

= 3 θ , 
2
√ (8.8)
ω (1) (2) = − aρ3 sin ϑ cos ϑθ(1) − r ρ3∆ θ(2) , 



ω (1) (3) − r ρ3∆ θ(3) − ar ρsin ϑ (0)

= 3 θ , 




r 2 +a2
ω (2) (3) − ρ3 cot ϑθ(3) + a ρ3∆ cos ϑθ(0) .

= 

Further we have to calculate the components of the Riemann curvature


tensor using the second Cartan structure equations where most time and
persistence are needed while performing differentiation in the first right-hand
side term, but all these calculations are by no means problematic. Of course,
in this first term it is advisable first to express ω (α) (β) in terms of the coordi-
nated basis [although the very ω (α) (β) will continue to pertain to the tetrad
basis (8.1)]. Then

M (r2 − a2 cos2 ϑ) ar sin2 ϑ


ω (0) (1) = (dt − a sin 2
ϑdφ) + dφ, (8.9)
ρ4 ρ2
√ √
a2 ∆ a ∆
ω (0) (2) 2
= −2 4 sin ϑ cos ϑ(dt − a sin ϑdφ) + 2 sin ϑ cos ϑdφ, (8.10)
ρ ρ

a ∆ ar
ω (0) (3) = − 2 cos ϑdϑ − 2 dr, (8.11)
ρ ρ∆

(1) a2 r ∆
ω (2) = − √ sin ϑ cos ϑdr − 2 dϑ, (8.12)
ρ2 ∆ ρ

r ∆
ω (1) (3) = − 2 sin ϑdφ, (8.13)
ρ
r 2 + a2 a∆
ω (2) (3) = − 4
cos ϑ[(r 2
+ a 2
)dφ − adt] + 4
cos ϑ(dt − a sin2 ϑdφ) (8.14)
ρ ρ
[sometimes written with fragments of basis covectors for convenience, since
finally we have to bring results to their standard form with respect to the
basis (8.1)]. Some of these expressions seem to be far from elegance, but
what could be demanded from such a mixture of bases we used in their
representation?!

42
References
Hawking, S.W., and Ellis, G.F.R. (1973) The Large Scale Structure of
Space-Time (Cambridge: CUP).

Hobson, M.P., Estathiou, G.P., and Lasenby, A.N. (2006) General Rel-
ativity. An Introduction to Physicists. (Cambridge: CUP).

Islam, J.N. (1985) Rotating Fields in General Relativity (Cambridge:


CUP).

Misner, C.W., Thorne, K.S., and Wheeler, J.A. (1973) Gravitation (San
Francisco: W.H. Freeman).

Mitskievich, N.V. (2006) Relativistic Physics in Arbitrary Peference


Frames (New York: Nova Science Publishers).

Mitskievich, N.V., and Vargas Rodrı́guez, H. (2005) Gen. Relat. and


Grav. 37, 781.

Stephani, H., Kramer, D., MacCallum, M., Hoenselaers, C., and Herlt,
E. (2003) Exact Solutions of Einstein’s Field Equations, 2nd edition
(Cambridge: CUP).

Synge, J.L. (1960) Relativity: The General Theory (Amsterdam: North-


Holland).

Teukolsky, S.A. (1973) Astrophys. J. 185, 635.

Vladimirov, Yu.S., Mitskievich, N.V., and Horský, J. (1987) Space,


Time, Gravitation (Moscow: Mir).

43

Vous aimerez peut-être aussi