Vous êtes sur la page 1sur 88

An introduction

to the theory of

rotating relativistic stars


arXiv:1003.5015v1 [gr-qc] 25 Mar 2010

Lectures at the COMPSTAR 2010 School


GANIL, Caen, France, 8-16 February 2010

ric Gourgoulhon

Laboratoire Univers et Thories,


UMR 8102 du C.N.R.S., Observatoire de Paris,
Universit Paris Diderot - Paris 7
92190 Meudon, France
eric.gourgoulhon@obspm.fr

25 March 2010
2
Contents

Preface 5

1 General relativity in brief 7


1.1 Geometrical framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1.1 The spacetime of general relativity . . . . . . . . . . . . . . . . . . . . . . 7
1.1.2 Linear forms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
1.1.3 Tensors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.1.4 Metric tensor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.1.5 Covariant derivative . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
1.2 Einstein equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3 3+1 formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
1.3.1 Foliation of spacetime . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.3.2 Eulerian observer or ZAMO . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.3.3 Adapted coordinates and shift vector . . . . . . . . . . . . . . . . . . . . . 17
1.3.4 Extrinsic curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
1.3.5 3+1 Einstein equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2 Stationary and axisymmetric spacetimes 21


2.1 Stationary and axisymmetric spacetimes . . . . . . . . . . . . . . . . . . . . . . . 21
2.1.1 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.1.2 Stationarity and axisymmetry . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.2 Circular stationary and axisymmetric spacetimes . . . . . . . . . . . . . . . . . . 27
2.2.1 Orthogonal transitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.2.2 Quasi-isotropic coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . 29
2.2.3 Link with the 3+1 formalism . . . . . . . . . . . . . . . . . . . . . . . . . 31

3 Einstein equations for rotating stars 33


3.1 General framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
3.2 Einstein equations in QI coordinates . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2.1 Derivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.2.2 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.2.3 Case of a perfect uid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
3.2.4 Newtonian limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.2.5 Historical note . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
3.3 Spherical symmetry limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.3.1 Taking the limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
4 CONTENTS

3.3.2 Link with the Tolman-Oppenheimer-Volko system . . . . . . . . . . . . . 41


3.3.3 Solution outside the star . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
3.4 Fluid motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
3.4.1 Equation of motion at zero temperature . . . . . . . . . . . . . . . . . . . 43
3.4.2 Bernoulli theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
3.4.3 First integral of motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
3.4.4 Stellar surface and maximum rotation velocity . . . . . . . . . . . . . . . 49
3.5 Numerical resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
3.5.1 The self-consistent-eld method . . . . . . . . . . . . . . . . . . . . . . . . 49
3.5.2 Numerical codes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
3.5.3 An illustrative solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

4 Global properties of rotating stars 57


4.1 Total baryon number . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.2 Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2.1 Komar mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.2.2 ADM mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2.3 Binding energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.3 Virial identities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.3.1 GRV3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.3.2 GRV2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62
4.4 Angular momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.5 Circumferential radius . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.6 Redshifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.6.1 General expression of the redshift . . . . . . . . . . . . . . . . . . . . . . . 65
4.6.2 Polar redshift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
4.6.3 Equatorial redshifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
4.7 Orbits around the star . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.7.1 General properties of orbits . . . . . . . . . . . . . . . . . . . . . . . . . . 68
4.7.2 Orbits in the equatorial plane . . . . . . . . . . . . . . . . . . . . . . . . . 69
4.7.3 Circular orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.7.4 Innermost stable circular orbit (ISCO) . . . . . . . . . . . . . . . . . . . . 72

A Lie derivative 75
A.1 Lie derivative of a vector eld . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
A.1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
A.1.2 Denition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
A.2 Generalization to any tensor eld . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

B Lorene/nrotstar code 79

Bibliography 81
Preface

These notes are the written version of lectures given at the Compstar 2010 School1 held in Caen
on 8-16 February 2010. They are intended to introduce the theory of rotating stars in general
relativity. Whereas the application is clearly towards neutron stars (and associated objects
like strange quark stars), these last ones are not discussed here, except for illustrative purposes.
Instead the focus is put on the theoretical foundations, with a detailed discussion of the spacetime
symmetries, the choice of coordinates and the derivation of the equations of structure from the
Einstein equation. The global properties of rotating stars (mass, angular momentum, redshifts,
orbits, etc.) are also introduced. Good introductions to rotating neutron star models computed
by the method exposed here are the review article by N. Stergioulas [87] and the textbook by
P. Haensel, A. Y. Potekhin & D. G. Yakovlev [50].
The present notes are limited to perfect uid stars. For more complicated physics (super-
uidity, dissipation, etc.), see N. Andersson & G. L. Comers review [3]. Magnetized stars are
treated in J. A. Pons lectures at the Compstar 2010 School. Moreover, the stability of rotating
stars is not investigated here. For a review of this rich topic, see N. Stergioulas article [87] or
L. Villains one [95].
To make the exposure rather self-contained, a brief overview of general relativity and the
3+1 formalism is given in Chap. 1. For the experienced reader, this can be considered as a mere
presentation of the notations used in the text. Chap. 2 is entirely devoted to the spacetime
symmetries relevant for rotating stars, namely stationarity and axisymmetry. The choice of co-
ordinates is also discussed in this chapter. The system of partial dierential equations governing
the structure of rotating stars is derived from the Einstein equation in Chap. 3, where its numer-
ical resolution is also discussed. Finally, Chap. 4 presents the global properties of rotating stars,
some of them being directly measurable by a distant observer, like the redshifts. The circular
orbits around the star are also discussed in this chapter.

Acknowledgements
I would like to thank warmly Jrme Margueron for the perfect organization of the Compstar
2010 School, as well as Micaela Oertel and Jrme Novak for their help in the preparation of
the nrotstar code. Some of the studies reported here are the fruit of a long and so pleasant
collaboration with Silvano Bonazzola, Pawe Haensel, Julian Leszek Zdunik, Dorota Gondek-
Rosiska and, more recently, Micha Bejger. To all of them, I express my deep gratitude.
Corrections and suggestions for improvement are welcome at eric.gourgoulhon@obspm.fr.

1
http://ipnweb.in2p3.fr/compstar2010/
6 CONTENTS
Chapter 1

General relativity in brief

Contents
1.1 Geometrical framework . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Einstein equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
1.3 3+1 formalism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

1.1 Geometrical framework


1.1.1 The spacetime of general relativity
Relativity has performed the fusion of space and time, two notions which were completely distinct
in Newtonian mechanics. This gave rise to the concept of spacetime, on which both the special
and general theory of relativity are based. Although this is not particularly fruitful (except for
contrasting with the relativistic case), one may also speak of spacetime in the Newtonian frame-
work: the Newtonian spacetime M is nothing but the ane space R4 , foliated by the hyperplanes
t of constant absolute time t: these hyperplanes represent the ordinary 3-dimensional space at
successive instants. The foliation (t )tR is a basic structure of the Newtonian spacetime and
does not depend upon any observer. The worldline L of a particle is the curve in M generated
by the successive positions of the particle. At any point A L, the time read on a clock moving
along L is simply the parameter t of the hyperplane t that intersects L at A.
The spacetime M of special relativity is the same mathematical space as the Newtonian one,
i.e. the ane space R4 . The major dierence with the Newtonian case is that there does not
exist any privileged foliation (t )tR . Physically this means that the notion of absolute time
is absent in special relativity. However M is still endowed with some absolute structure: the
metric tensor g and the associated light cones. The metric tensor is a symmetric bilinear
form g on M , which denes the scalar product of vectors. The null (isotropic) directions of g
give the worldlines of photons (the light cones). Therefore these worldlines depend only on the
absolute structure g and not, for instance, on the observer who emits the photon.
The spacetime M of general relativity diers from both Newtonian and special relativistic
spacetimes, in so far as it is no longer the ane space R4 but a more general mathematical
8 General relativity in brief

Figure 1.1: The vectors at two points p and q on the spacetime manifold M belong to two different
vector spaces: the tangent spaces Tp (M ) and Tq (M ).

structure, namely a manifold. A manifold of dimension 4 is a topological space such that


around each point there exists a neighbourhood which is homeomorphic to an open subset of R4 .
This simply means that, locally, one can label the points of M in a continuous way by 4 real
numbers (x ){0,1,2,3} (which are called coordinates). To cover the full M , several dierent
coordinate patches (charts in mathematical jargon) can be required.
Within the manifold structure the denition of vectors is not as trivial as within the ane
structure of the Newtonian and special relativistic spacetimes. Indeed, only innitesimal vectors
connecting two innitely close points can be dened a priori on a manifold. At a given point
p M , the set of such vectors generates a 4-dimensional vector space, which is called the
tangent space at the point p and is denoted by Tp (M ). The situation is therefore dierent
from the Newtonian or special relativistic one, for which the very denition of an ane space
provides a unique global vector space. On a manifold there are as many vector spaces as points
p (cf. Fig. 1.1).
Given a coordinate system (x ) around some point p M , one can dene a vector basis
~
( ){0,1,2,3} of the tangent space Tp (M ) by asking that the components of the innitesimal
vector
that connect neighbouring points p (coordinates (x )) and p (coordinates (x + dx ))
pp
are the innitesimal increments dx in the coordinates when moving from p to p :

3
X


pp = ~ = dx
dx ~ . (1.1)
=0

For the second equality in the above equation, we have used Einstein summation convention
on repeated indices, which amounts to omitting the sign. We shall make a systematic use of
this convention in these notes. The vector basis ( ~ ) is called the natural basis associated
~
with the coordinates (x ). The notation is inspired by /x and is a reminiscence of the

intrinsic denition of vectors on a manifold as dierential operators acting on scalar elds (see
e.g. [96]).
For any vector ~
v Tp (M ) (not necessarily an innitesimal one), one denes the components
1.1 Geometrical framework 9

~ ) according to the expansion


(v ) of ~v in the basis (
~ .
~v = v (1.2)

1.1.2 Linear forms


Besides vectors, important objects on the manifold M are linear forms. At each point p M ,
a linear form is a an application1
: Tp (M ) R
(1.3)
~v 7 h, v
~i

that is linear: h, ~
v+u ~ i = h, v ~ i. The set of all linear forms at p constitutes a 4-
~ i + h, u
dimensional vector space, which is called the dual space of Tp (M ) and denoted by Tp (M ) . A
eld of linear forms on M is called a 1-form. Given the natural basis ( ~ ) of Tp (M ) associated
with the coordinates (x ), there is a unique basis of Tp (M ) , denoted by (dx ), such that

~ i = ,
hdx , (1.4)

where is the Kronecker symbol : = 1 if = and 0 otherwise. The basis (dx )


is called the dual basis of the basis ( ~ ). The notation (dx ) stems from the fact that if we

apply the linear form dx to the innitesimal displacement vector (1.1), we get nothing but the
number dx :
hdx ,
i = hdx , dx
pp ~ i = dx hdx ,
~ i = dx = dx .
(1.5)
The dual basis can be used to expand any linear form , thereby dening its components :

= dx . (1.6)

In terms of components, the action of a linear form on a vector takes then a very simple form:

~ i = v .
h, v (1.7)

Given a smooth scalar field on M , namely a smooth application f : M R, there exists a


1-form canonically associated with it, called the gradient of f and denoted df . It is the unique
linear form which, once applied to the innitesimal displacement vector
, gives the change in
pp
f between points p and p :
df := f (p ) f (p) = df (
).
pp (1.8)
Since df = f /x dx , formula (1.5) implies that the components of the gradient with respect
to the dual basis are nothing but the partial derivatives of f with respect to the coordinates (x )
:
f
df = dx . (1.9)
x
Remark : In non-relativistic physics, the gradient is very often considered as a vector and not as a
1-form. This is so because one associates implicitly a vector ~ to any 1-form via the Euclidean
~
scalar product of R3 : ~v R3 , h, ~v i = ~ ~v . Accordingly, formula (1.8) is rewritten as df = f .
pp
But one shall keep in mind that, fundamentally, the gradient is a linear form and not a vector.
1
We are using the same bra-ket notation as in quantum mechanics to denote the action of a linear form on a
vector.
10 General relativity in brief

1.1.3 Tensors
In relativistic physics, an abundant use is made of linear forms and their generalizations: the
tensors. A tensor of type (k, ), also called tensor k times contravariant and times
covariant, is an application

T : Tp (M ) Tp (M ) Tp (M ) Tp (M ) R
| {z } | {z }
k times times
~1, . . . , v
(1 , . . . , k , v ~ ) 7 T (1 , . . . , k , v
~ 1, . . . , v
~ )
(1.10)
that is linear with respect to each of its arguments. The integer k + is called the valence of the
tensor. Let us recall the canonical duality Tp (M ) = Tp (M ), which means that every vector
~ can be considered as a linear form on the space Tp (M ) , via v
v ~ : Tp (M ) R, 7 h, v ~ i.
Accordingly a vector is a tensor of type (1, 0). A linear form is a tensor of type (0, 1). A tensor
of type (0, 2) is called a bilinear form. It maps couples of vectors to real numbers, in a linear
way for each vector.
Given the natural basis ( ~ ) associated with some coordinates (x ) and the corresponding
dual basis (dx ), we can expand any tensor T of type (k, ) as

T = T 1 ...k1 ...
~ . . .
1
~ dx1 . . . dx ,
k
(1.11)

where the tensor product ~ . . . ~ dx1 . . . dx is the tensor of type (k, ) for
1 k
which the image of (1 , . . . , k , v ~ ) as in (1.10) is the real number
~ 1, . . . , v

k
Y
Y
~ i
hi , hdxj , v
~ j i.
i
i=1 j=1

Notice that all the products in the above formula are simply products in R. The 4k+ scalar
coecients T 1 ...k1 ... in (1.11) are called the components of the tensor T with respect
the coordinates (x ). These components are unique and fully characterize the tensor T .

Remark : The notation v and already introduced for the components of a vector v ~ or a linear form
are of course the particular cases (k, ) = (1, 0) and (k, ) = (0, 1) of the general definition given
above.

1.1.4 Metric tensor


As for special relativity, the fundamental structure given on the spacetime manifold M of general
relativity is the metric tensor g. It is now a eld on M : at each point p M , g(p) is a bilinear
form acting on vectors in the tangent space Tp (M ):

g(p) : Tp (M ) Tp (M ) R
(1.12)
(~ ~)
u, v 7 g(~ ~ ).
u, v

It is demanded that

g is symmetric: g(~ ~ ) = g(~


u, v ~ );
v, u
1.1 Geometrical framework 11

g is non-degenerate: any vector u ~ that is orthogonal to all vectors (~


v Tp (M ), ~) =
u, v
g(~
0) is necessarily the null vector;

g has a signature (, +, +, +): in any basis of Tp (M ) where the components (g ) form


a diagonal matrix, there is necessarily one negative component (g00 say) and three positive
ones (g11 , g22 and g33 say).

Remark : An alternative convention for the signature of g exists in the literature (mostly in particle
physics): (+, , , ). It is of course equivalent to the convention adopted here (via a mere
replacement of g by g). The reader is warned that the two signatures may lead to different signs
in some formulas (not in the values of measurable quantities !).

The properties of being symmetric and non-degenerate are typical of a scalar product,
thereby justifying the notation

~ v
u ~ := g(~ ~ ) = g u v
u, v (1.13)

that we shall employ throughout. The last equality in the above equation involves the components
g of g, which according to denition (1.11), are given by

g = g dx dx . (1.14)

The isotropic directions of g give the local light cones: a vector v ~ Tp (M ) satisfying
~ = 0 is tangent to a light cone and called a null or lightlike vector. If ~v ~v < 0, the vector
~ v
v
~ is said timelike and if v
v ~ ~ ~ is said spacelike.
v > 0, v

1.1.5 Covariant derivative


Given a vector eld ~ v on M , it is natural to consider the variation of v
~ between two neighbouring

points p and p . But one faces the following problem: from the manifold structure alone, the
variation of v
~ cannot be dened by a formula analogous to that used for a scalar eld [Eq. (1.8)],
namely d~ v = v~ (p ) ~v(p), because ~ v(p ) and v
~ (p) belong to dierent vector spaces: Tp (M )
and Tp (M ) respectively (cf. Fig. 1.1). Accordingly the subtraction v ~ (p ) ~v (p) is not well
dened ! The resolution of this problem is provided by introducing an affine connection on
the manifold, i.e. of a operator that, at any point p M , associates to any innitesimal
displacement vector
pp Tp (M ) a vector of Tp (M ) denoted ~v that one denes as the
pp
variation of ~ v between p and p :

d~v = ~.
v
pp (1.15)

The complete denition of allows for any vector u ~ Tp (M ) instead of and includes all
pp
the properties of a derivative operator (Leibniz rule, etc...), which we shall not list here (see e.g.
Walds textbook [96]). One says that v ~ is transported parallelly to itself between p and p
i d~v = 0.
The name connection stems from the fact that, by providing the denition of the variation d~ v,
the operator connects the adjacent tangent spaces Tp (M ) and Tp (M ). From the manifold
structure alone, there exists an innite number of possible connections and none is preferred.
Taking account the metric tensor g changes the situation: there exists a unique connection,
called the Levi-Civita connection, such that the tangent vectors to the geodesics with respect
12 General relativity in brief

to g are transported parallelly to themselves along the geodesics. In what follows, we will make
use only of the Levi-Civita connection.
Given a vector eld ~
v and a point p M , we can consider the type (1, 1) tensor at p denoted
by ~v (p) and dened by

v (p) : Tp (M ) Tp (M )
~ R
. (1.16)
(, u~) 7 (u~ ~v )

By varying p we get a type (1, 1) tensor eld denoted ~ v and called the covariant derivative
of v
~.
The covariant derivative is extended to any tensor eld by (i) demanding that for a scalar
eld f := df and (ii) using the Leibniz rule. As a result, the covariant derivative of a tensor
eld T of type (k, ) is a tensor eld T of type (k, + 1). Its components with respect a given
coordinate system (x ) are denoted

T 1 ...k1 ... := (T )1 ...k 1 ... (1.17)

(note the position of the index !) and are given by


i
k
1 ...k 1 ...k X
i 1 ... ...k
T 1 ... = T 1 ... + T 1 ...
x
i=1

X
i T 1 ...k1 ... ... , (1.18)

i=1 i

where the coecients are the Christoffel symbols of the metric g with respect to the
coordinates (x ). They are expressible in terms of the partial derivatives of the components of
the metric tensor, via  
1 g g g
:= g + , (1.19)
2 x x x
where g stands for the components of the inverse matrix2 of (g ):

g g = . (1.20)

A distinctive feature of the Levi-Civita connection is that

g = 0 . (1.21)

Given a vector eld u


~ and a tensor eld T of type (k, ), we dene the covariant derivative
of T along u~ as
~ ).
u~ T := T ( ., . . . , . , u (1.22)
| {z }
k+ slots

Note that u~ T is a tensor of the same type (k, ) as T and that its components are

(u~ T )1 ...k 1 ... = u T 1 ...k1 ... . (1.23)


2
Since g is non-degenerate, its matrix (g ) is always invertible, whatever the coordinate system.
1.2 Einstein equation 13

1.2 Einstein equation


General relativity is ruled by the Einstein equation:

1 8G
R Rg = 4 T , (1.24)
2 c

where R is the Ricci tensor associated with the Levi-Civita connection (to be dened below),
R := trg R = g R is the trace of R with respect to g and T is the energy-momentum tensor of
matter and electromagnetic eld3 (to be dened below). The constants G and c are respectively
Newtons gravitational constant and the speed of light. In the rest of this lecture, we choose
units such that
G=1 and c = 1. (1.25)
The Ricci tensor R is a part of the Riemann tensor R, a type (1, 3) tensor which
describes the curvature of , i.e. the tendency of two initially parallel geodesics to deviate
from each other. Equivalently the Riemann tensor measures the lack of commutativity of two
successive covariant derivatives:

v v = R v . (1.26)

The Ricci tensor R is dened as the following trace of the Riemann tensor:

R := R . (1.27)

It is expressible in terms of the derivatives of the components of the metric tensor with respect
to some coordinate system (x ) as


R = + , (1.28)
x x
where the have to be replaced by their expression (1.19). Although this not obvious on the
above expression, the Ricci tensor is symmetric : R = R .
Having dened the left-hand side of Einstein equation (1.24), let us focus on the right-hand
side, namely the energy-momentum tensor T . To dene it, let us introduce a generic observer
O in spacetime. It is described by its worldline L , which is a timelike curve in M , and the
future-directed unit vector u
~ tangent to L (the so-called observer 4-velocity ):

~ u
u ~ = 1. (1.29)

An important operator is the projector onto the 3-dimensional vector space Eu orthogonal to
~ , which can be considered as the local rest space of the observer O. is a type (1, 1) tensor,
u
the components of which are
= + u u . (1.30)

In particular, u = u + u u u = u + u (1) = 0. The energy-momentum tensor T is


the bilinear form dened by the following properties which must be valid for any observer O:
3
Or any kind of field which are present in spacetime.
14 General relativity in brief

the energy density measured by O is

E = T (~ ~ ) = T u u ;
u, u (1.31)

the momentum density measured by O is the 1-form p the components of which are

p = T u ; (1.32)

the energy flux vector 4 measured by O is the 1-form the components of which are

= T u ; (1.33)

the stress tensor measured by O is the tensor S the components of which are

S = T . (1.34)

A basic property of the energy-momentum tensor is to be symmetric. Anyway, it could not be


otherwise since the left-hand side of Einstein equation (1.24) is symmetric. From Eqs. (1.32) and
(1.33), the symmetry of T implies the equality of the momentum density and the energy ux
form5 : p = . This reects the equivalence of mass and energy in relativity.
Another property of T follows from the following geometric property, known as contracted
Bianchi identity :  
1
R R g = 0, (1.35)
2
where := g . The Einstein equation (1.24) implies then the vanishing of the divergence
of T :
T = 0 . (1.36)
This equation is a local expression of the conservation of energy and momentum.
An astrophysically very important case of energy-momentum tensor is that of a perfect
fluid :
T = ( + p)u u + p g , (1.37)
where u is the 1-form associated by metric duality6 to a unit timelike vector eld u ~ representing
the uid 4-velocity, and p are two scalar elds, representing respectively the energy density and
the pressure, both in the uid frame. Indeed, plugging Eq. (1.37) into Eqs. (1.31)-(1.34) leads
respectively to the following measured quantities by an observer comoving with the uid:

E = , p = 0, , = 0, S = p .

1.3 3+1 formalism


The 3+1 decomposition of Einstein equation is the standard formulation used in numerical
relativity [2, 13, 42]. It is also very useful in the study of rotating stars. Therefore we introduce
it briey here (see e.g. [42] for an extended presentation).
4
For an electromagnetic field, this is the Poynting vector.
5
Restoring units where c 6= 1, this relation must read p = c2 .
6
The components of u are u = g u .
1.3 3+1 formalism 15

Figure 1.2: Foliation of the spacetime M by a family of spacelike hypersurfaces (t )tR .

1.3.1 Foliation of spacetime


The 3+1 formalism relies on a slicing of spacetime by a family of spacelike hypersurfaces (t )tR .
Hypersurface means that t is a 3-dimensional submanifold of M , spacelike means that every
vector tangent to t is spacelike and slicing means (cf. Fig. 1.2)
[
M = t . (1.38)
tR

with t t = for t 6= t . Not all spacetimes allow for a global foliation as in Eq. (1.38), but
only those belonging to the class of the so-called globally hyperbolic spacetimes 7 . However
this class is large enough to encompass spacetimes generated by rotating stars.
At this stage, t is a real parameter labelling the hypersurfaces t , to be identied later with
some coordinate time. We may consider t as a scalar eld on M . It gradient t = df is then
a 1-form that satises ht, v ~ i = 0 for any vector v
~ tangent to t . This is direct consequence of t
~
being constant on t . Equivalently the vector t associated to the 1-form t by metric duality
(i.e. the vector whose components are t = g t) is normal to t : t ~ ~v = 0 for any vector
~ tangent to t . Since t is spacelike, it possesses at each point a unique unit timelike normal
v
vector n~ , which is future-oriented (cf. Fig. 1.2):

~ n
n ~ = 1. (1.39)
~ and n
The two normal vectors t ~ are necessarily colinear:

~ .
~ = N t
n (1.40)
7
Cf. [33] or 3.2.1 of [42] for the precise definition.
16 General relativity in brief

The proportionality coecient N is called the lapse function. The minus sign is chosen so that
N 0 if the scalar eld t is increasing towards the future.
Since each hypersurface t is assumed to be spacelike, the metric induced8 by g onto t
is definite positive, i.e.
~
v Tp (t ), ~v 6= 0 (~ ~ ) > 0.
v, v (1.41)
Considered as a tensor eld on M , the components of are given in terms of the components
of the normal via
= g + n n . (1.42)
Note that if we raise the rst index via g , we get the components of the orthogonal projector
onto t [compare with (1.30)] :
= + n n , (1.43)
which we will denote by ~
.
Let D be the Levi-Civita connection in t associated with the metric . It is expressible in
terms of the spacetime connection and the orthogonal projector ~ as
1 ...p p q 1 ...p
D T 1 ...q = 11 p 1 1 q T 1 ...q . (1.44)

1.3.2 Eulerian observer or ZAMO


Since n
~ is a unit timelike vector, we may consider the family of observers whose 4-velocity is n
~.
Their worldlines are then the eld lines of n
~ and are everywhere orthogonal to the hypersurfaces
t . These observers are called the Eulerian observers. In the context of rotating stars or
black holes, they are also called the locally non-rotating observers [10] or zero-angular-
momentum observers (ZAMO) [94] . Thanks to relation (1.40), the proper time of a
ZAMO is related to t by
d = N dt, (1.45)
hence the name lapse function given to N . The 4-acceleration of a ZAMO is the vector ~a := n~n~.
This vector is tangent to t and is related to the spatial gradient of the lapse function by (see
e.g. 3.3.3 of [42] for details)
~ ln N = ~ (
~a = D ~ ln N ). (1.46)
In general ~a 6= 0, so that the ZAMOs are accelerated observers (they feel the gravitational
force). On the other side, they are non-rotating: being orthogonal to the hypersurfaces t ,
their worldlines form a congruence without any twist. Physically this means that the ZAMOs
do not feel any centrifugal nor Coriolis force.
Let us express the energy-momentum tensor T in terms of the energy E density, the momen-
tum density p and the stress tensor S, these three quantities being measured by the ZAMO:
setting u~ =n~ and = ~ in Eqs. (1.31), (1.32) and (1.34), we get
E = T n n (1.47)
p = T u (1.48)
S = T . (1.49)
The associated 3+1 decomposition of the energy-momentum tensor is then:
T = E n n + p n + n p + S . (1.50)
8
is nothing but the restriction of g to t .
1.3 3+1 formalism 17

Figure 1.3: Coordinates (xi ) on the hypersurfaces t : each line xi = const cuts across the foliation
~t and the shift vector of the spacetime coordinate system (x ) =
(t )tR and defines the time vector
i
(t, x ).

1.3.3 Adapted coordinates and shift vector


A coordinate system (x ) is said to be adapted to the foliation (t )tR i x0 = t. Then the
triplet9 (xi ) = (x1 , x2 , x3 ) constitutes a coordinate system on each hypersurface t , which we
may call spatial coordinates. Given such a coordinate system, we may decompose the natural
basis vector ~t into a part along n ~ and a part tangent to t :

~t = N n
~
~ + with n ~ = 0.
~ (1.51)

The spacelike vector ~ is called the shift vector. Indeed it measures the shift of the lines of
constant spatial coordinates with respect to the normal to the hypersurfaces t (cf. Fig. 1.3). The
fact that the coecient of n ~ in Eq. (1.51) is the lapse function N is an immediate consequence
~
of hdt, t i = 1 [Eq. (1.4)] and relation (1.40). As for any vector tangent to t , the shift vector
has no component along ~t :
~ = i
~i . (1.52)
Equation (1.51) leads then to the following expression for the components of n
~:
 
1 1 2 3
n = , , , . (1.53)
N N N N
The covariant components (i.e. the components of the 1-form associated with n
~ by metric
duality) are given by Eq. (1.40) and (1.6):

n = (N, 0, 0, 0). (1.54)


~ is a timelike unit vector: n
Remark : It is immediate to check on (1.53)-(1.54) that n ~ = n n =
~ n
(N )(1/N ) = 1.
9
Latin indices (i, j, ...) run in {1, 2, 3}, whereas Greek indices (, , ...) run in {0, 1, 2, 3}.
18 General relativity in brief

The components (g ) of the spacetime metric are expressible in terms of the components
(ij ) of the induced metric in t , the components of the shift vector and the lapse function:

g dx dx = N 2 dt2 + ij (dxi + i dt)(dxj + j dt) . (1.55)

1.3.4 Extrinsic curvature


The intrinsic curvature of the hypersurface t equipped with the induced metric is given
by the Riemann tensor of the Levi-Civita connection D associated with (cf. 1.2). On the
other side, the extrinsic curvature describes the way t is embedded into the spacetime (M , g).
It is measurable by the variation of the normal unit vector n
~ as one moves on t . More precisely,
the extrinsic curvature tensor K is the bilinear form dened on t by

(~ ~ ) Tp (t ) Tp (t ),
u, v ~ ) := ~
u, v
K(~ u v~ n
~. (1.56)

It can be shown (see e.g. 2.3.4 of [42]) that, as a consequence of n ~ being hypersurface-
orthogonal, the bilinear form K is symmetric (this result is known as the Weingarten prop-
erty ).
The components of K with respect to the coordinates (xi ) in t are expressible in terms of
the time derivative of the induced metric according to
   
1 ij 1 ij ij k k
Kij = L~ ij = + k k + kj i + ik j . (1.57)
2N t 2N t x x x

~ (cf. Ap-
L~ ij stands for the components of the Lie derivative of along the vector eld
pendix A) and the second equality results from the 3-dimensional version of Eq. (A.8). The
trace of K with respect to the metric is connected to the covariant divergence of the unit
normal to t :
K := ij Kij = n . (1.58)

1.3.5 3+1 Einstein equations


Projecting the Einstein equation (1.24) (i) twice onto t , (ii) twice along n
~ and (iii) once on t
and once along n~ , one gets respectively the following equations [2, 13, 42] :

Kij n o
L~ Kij = Di Dj N + N 3 Rij + KKij 2Kik K kj + 4 [(S E)ij 2Sij ]
t
(1.59)
3
R + K 2 Kij K ij = 16E (1.60)

Dj K j i Di K = 8pi . (1.61)

In this system, E, pi and Sij are the matter quantities relative to the ZAMO and are dened
respectively by Eqs. (1.47)-(1.49). The scalar S is the trace of S with respect to the metric :
S = ij Sij . The covariant derivatives Di can be expressed in terms of partial derivatives with
1.3 3+1 formalism 19

respect to the spatial coordinates (xi ) by means of the Christoel symbols 3 i jk of D associated
with (xi ):

2N N
Di Dj N = i j
3 k ij k , (1.62)
x x x
j
K
Dj K j i = i
+ 3 j jk K ki 3 k ji K j k , (1.63)
xj
K
Di K = . (1.64)
xi

The L~ Kij are the components of the Lie derivative of the tensor K along the vector ~ (cf.
Appendix A); according to the 3-dimensional version of formula (A.8), they can be expressed in
terms of partial derivatives with respect to the spatial coordinates (xi ) :

Kij k k
L~ Kij = k + Kkj + Kik . (1.65)
xk xi xj
The Ricci tensor and scalar curvature of are expressible according to the 3-dimensional analog
of formula (1.28):

3
3 k ij 3 k ik 3 k 3 l
Rij = + ij kl 3 l ik 3 k lj (1.66)
xk xj
3 ij 3
R = Rij . (1.67)

Finally, let us recall that Kij is related to the derivatives of ij by Eq. (1.57).
Equations (1.60) and (1.61), which do not contain any second order time derivative of ij ,
are called respectively the Hamiltonian constraint and the momentum constraint.
The 3+1 Einstein equations are at the basis of the formulation of general relativity as a
Cauchy problem and are much employed in numerical relativity (for reviews see e.g. [33, 42, 2]).
20 General relativity in brief
Chapter 2

Stationary and axisymmetric


spacetimes

Contents
2.1 Stationary and axisymmetric spacetimes . . . . . . . . . . . . . . . . 21
2.2 Circular stationary and axisymmetric spacetimes . . . . . . . . . . . 27

Having reviewed general relativity in Chap. 1, we focus now on spacetimes that possess two
symmetries: stationarity and axisymmetry, in view of considering rotating stars in Chap. 3.

2.1 Stationary and axisymmetric spacetimes


2.1.1 Definitions
Group action on spacetime
Symmetries of spacetime are described in a coordinate-independent way by means of a (sym-
metry) group acting on the spacetime manifold M . Through this action, each transformation
belonging to the group displaces points within M and one demands that the metric g is invariant
under such displacement. More precisely, given a group G, a group action of G on M is an
application
: G M M
(2.1)
(g, p) 7 (g, p) =: g(p)
such that

(g, h) G2 , p M , g(h(p)) = gh(p), where gh denotes the product of g by h according


to the group law of G (cf. Fig. 2.1);

if e is the identity element of G, then p M , e(p) = p .

The orbit of a point p M is the set {g(p), g G} M , i.e. the set of points which are
connected to p by some group transformation.
22 Stationary and axisymmetric spacetimes

Figure 2.1: Group action on the spacetime manifold M .

An important class of group actions are those for which G is a one-dimensional Lie group
(i.e. a continuous group). Then around e, the elements of G can be labelled by a parameter
t R, such that gt=0 = e. The orbit of a given point p M under the group action is then
either {p} (case where p is xed point of the group action) or a one-dimensional curve of M . In
the latter case, t is then a natural parameter along the curve (cf. Fig. 2.2). The tangent vector
corresponding to that parameter is called the generator of the symmetry group associated
with the t parametrization. It is given by
d~
x
~ = , (2.2)
dt
where d~x is the innitesimal vector connecting the point p to the point gdt (p) (cf. 1.1.1 and
Fig. 2.2). The action of G on M in any innitesimal neighbourhood of p amounts then to
~
translations along the innitesimal vector dt .

Stationarity
A spacetime (M , g) is said to be stationary i there exists a group action on M with the
following properties:
1. the group G is isomorphic to (R, +), i.e. the group of unidimensional translations;
2. the orbits are timelike curves in M ;
3. the metric is invariant under the group action, which is translated by

L~ g = 0 , (2.3)

where ~ is the generator of G associated with some parameter t of G and L~ g denotes


the Lie derivative of g along the vector eld .~ The Lie derivative measures the variation
~
of g along the eld lines of , i.e. the variation under the group action, and is dened in
Appendix A.
2.1 Stationary and axisymmetric spacetimes 23

Figure 2.2: Orbit of a point p under the action of a one-dimensional Lie group, the elements gt of which
being labelled by the parameter t R. The vector ~ = d~x/dt is the group generator associated with this
parameter.

Remark 1: The property (2.3) expresses the fact that if two vectors u ~ and v~ are invariant under trans-
~ ~ ~v is also invariant. Indeed, thanks to
port along the field lines of , then their scalar product u
the Leibniz rule:

~ (~ v ) = L~ (~
u~ v ) = L~ [g(~
u~ u, L~ v
~ , ~v ) + g(~
u, ~v ) + g(L~ u
u, ~v )] = (L~ g)(~ ~ ),

so that if L~ u
~ = 0 and L~ v
~ = 0 hold, L~ g = 0 implies ~ (~
uv
~ ) = 0.

~ i.e. of the parametrization


Remark 2: The property (2.3) is independent of the choice of the generator ,
t of G. Indeed, under a change of parametrization t 7 t , inducing a change of generator ~ 7 ~ ,

the following scaling law holds:


dt
L~ g = L~ g.
dt

Expressing the Lie derivative L~ g according to Eq. (A.11) and using g = 0 [property (1.21)]
as well as = g shows that the condition (2.3) is equivalent to

+ = 0 . (2.4)

This equation is known as Killing equation 1 . Accordingly, the symmetry generator ~ is called
a Killing vector.
For an asymptotically at spacetime, as we are considering here, we can determine the Killing
eld uniquely ~ by demanding that far from the central object ~ ~ 1. Consequently, the
associated parameter t coincides with the proper time of the asymptotically inertial observer at
rest with respect to the central source. In the following we shall employ only that Killing vector.
1
Named after the German mathematician Wilhelm Killing (1847-1923).
24 Stationary and axisymmetric spacetimes

Staticity
A spacetime (M , g) is said to be static i

1. it is stationary;

2. the Killing vector eld ~ is orthogonal to a family of hypersurfaces.

Remark : Broadly speaking, condition 1 means that nothing depends on time and condition 2 that
there is no motion in spacetime. This will be made clear below for the specific case of rotating
stars: we will show that condition 2 implies that the star is not rotating.

Axisymmetry
A spacetime (M , g) is said to be axisymmetric i there exists a group action on M with the
following properties:

1. the group G is isomorphic to SO(2), i.e. the group of rotations in the plane;

2. the metric is invariant under the group action:

L~ g = 0 , (2.5)

~ being the generator of G associated with some parameter of G and L~ g denoting the

Lie derivative of g along the vector eld
~;

3. the set of xed points under the action of G is a 2-dimensional surface of M , which we
will denote by .

Carter [29] has shown that for asymptotically at spacetimes, which are our main concern here,
the rst and second properties in the above denition imply the third one. Moreover, he has
also shown that is necessarily a timelike 2-surface, i.e. the metric induced on it by g has the
signature (, +). is called the rotation axis.
Remark : The 2-dimensional and timelike characters of the rotation axis can be understood by con-
sidering as the time development (history) of the standard one-dimensional rotation axis in a
3-dimensional space.

From the very denition of a generator of a symmetry group, the vector eld ~ must vanish on
the rotation axis (otherwise, the latter would not be a set of xed points):

~ | = 0 .
(2.6)

~ the condition (2.5) is equivalent to demanding that


In addition, as for , ~ be a Killing vector:

+ = 0 . (2.7)

Given a SO(2) group action, we can determine uniquely the Killing vector ~ by demanding
that the associated parameter takes its values in [0, 2[. In the following, we shall always
employ that Killing vector.
2.1 Stationary and axisymmetric spacetimes 25

Figure 2.3: Commutativity of the stationary and axisymmetric actions.

2.1.2 Stationarity and axisymmetry


We consider a spacetime (M , g) that is both stationary and axisymmetric. Carter has shown in
1970 [29] that no generality is lost by considering that the stationary and axisymmetric actions
commute. In other words, the spacetime (M , g) is invariant under the action of the Abelian group
R SO(2), and not only under the actions of R and SO(2) separately. Saying that the stationary
and axisymmetric actions commute means that starting from any point p M , moving to the
point q = Tt (p) under the action of an element Tt of the group (R, +) and then displacing q
via an element R of SO(2) yields the same point as when performing the displacements in the
reverse order (cf. Fig. 2.3):
p M , t R, [0, 2[, R (Tt (p)) = Tt (R (p)). (2.8)

This property can be translated in terms of the Killing vectors ~ and


~ associated respectively
with the parameter t of (R, +) and the parameter of SO(2). Indeed (2.8) is equivalent to the
vanishing of their commutator:
~
[, ~] = 0 , (2.9)
the commutator being the vector whose components are

~

[, ~ ] = = . (2.10)
x x
An important consequence of the above commutation property is that the parameters t and
labelling the two symmetry groups (R, +) and SO(2) can be chosen as coordinates on the
spacetime manifold M . Indeed, having set the coordinates (t, ) = (0, 0) to a given point p,
(2.8) implies that we can unambiguously attribute the coordinates (t, ) to the point obtained
from p by a time translation of parameter t and a rotation of parameter , whatever the order
of these two transformations (cf. Fig. 2.3).
Let us complete (t, ) by two other coordinates (x1 , x2 ) = (r, ) to get a full coordinate
system on M :
(x ) = (t, r, , ) . (2.11)
26 Stationary and axisymmetric spacetimes

Figure 2.4: Spherical coordinates (r, , ) and the associated natural basis (~r , ~ , ~ ).

(r, ) are chosen so that the spatial coordinates (r, , ) are of spherical type: r [0, +[,
[0, ] and = 0 or = on the rotation axis (cf. Fig. 2.4).
Remark : An alternative choice would have been (x1 , x2 ) = (, z) such that the spatial coordinates
(, z, ) are of cylindrical type: [0, +[, z R and = 0 on . The relation between the two
types of coordinates is = r sin and z = r cos .

By construction, the coordinate system (2.11) is such that the rst and fourth vectors of the
associated natural basis (cf. 1.1.1) are the two Killing vectors:
~t = ~
and ~ =
~. (2.12)

Consequently, in terms of tensor components in the coordinates (x ), the Lie derivatives with
respect to ~ (resp.
~ ) reduce to partial derivatives with respect to t (resp. ) [cf. Eq. (A.3)]. In
particular, the stationary and axisymmetry conditions (2.3) and (2.5) are respectively equivalent
to
g g
=0 and =0. (2.13)
t
For this reason, t and are called ignorable coordinates. The four coordinates (x ) are said
to be adapted coordinates to the spacetime symmetries.
There is some freedom in choosing the coordinates (x ) with the above properties. Indeed,
any change of coordinates of the form


t = t + T (r, )

r = R(r, )
(2.14)
= (r, )


= + (r, ) ,

where T , R, and are arbitrary (smooth) functions2 of (r, ), lead to another adapted coor-
2
Of course, to preserve the spherical type of the spatial coordinates, the functions R and have to fulfill
certain properties.
2.2 Circular stationary and axisymmetric spacetimes 27

Figure 2.5: Change of coordinate t 7 t according to (2.14). The solid lines represent the hypersurfaces
of constant t (t ) and the dashed lines those of constant t (t ). The dotted lines are the orbits of the
~
stationarity action (generator ).

dinate system:
~t = ~
and ~ =
~. (2.15)

It is important to realize that in the above equation, ~ and ~ are the same vectors than in

Eq. (2.12). The change of coordinate t 7 t according to (2.14) is merely a reparametrization of
each orbit of the stationarity action reparametrization which leaves invariant the associated
tangent vector ~ (cf. Fig. 2.5). A similar thing can be said for 7 .

Remark : As illustrated in Fig. 2.5, stationarity by itself does not introduce a privileged 3+1 slicing of
spacetime (cf. 1.3), for the change of t coordinate according to (2.14) changes the hypersurfaces
of constant t.

2.2 Circular stationary and axisymmetric spacetimes


2.2.1 Orthogonal transitivity
Regarding the components of the metric tensor, the properties (2.13) are a priori the only ones
implied by the spacetime symmetries (stationarity and axisymmetry). In particular, we are not
allowed to set to zero some components g , which would of course simplify further the expres-
sion of Einstein equations. There is however a wide subclass of stationary and axisymmetric
spacetimes in which we are allowed to do so for 5 components: the so-called circular spacetimes.
Furthermore, these spacetimes are much relevant for astrophysics.
Let us rst remark that having g = 0 for some value of (, ) is an orthogonality condition:
~ and
that of the vectors ~ , i.e. of the lines {x = const, 6= } and {x = const, 6= }.
Besides, we note that, in the present context, there are privileged 2-dimensional surfaces in
spacetime: the orbits of the RSO(2) group action. They are called the surfaces of transitivity
and denoted SA , where A is some label. The two Killing vectors ~ and ~ are everywhere tangent
to these surfaces and, except on the rotation axis, they form a basis of the tangent space to SA
28 Stationary and axisymmetric spacetimes

at each point. Once adapted coordinates (x ) = (t, r, , ) are chosen, the surfaces of transitivity
can be labelled by the value of (r, ) since both r and are xed on each of these surfaces:

SA = Sr = {p M / x1 (p) = r, x2 (p) = }. (2.16)


Remark : r is a mere label for the surfaces Sr ; the latter depend only on the symmetry group
R SO(2) and not on the choice of the (r, ) coordinates. A coordinate change (r, ) 7 (r , )
according to (2.14) will simply result in a relabelling of the surfaces Sr .

Given the family of 2-surfaces (Sr ), one may ask if there exists another family of 2-surfaces,
(MB ) say, which are everywhere orthogonal to Sr . If this is the case, the R SO(2) group
action is said to be orthogonally transitive [28] and the spacetime (M , g) to be circular.
One may choose the coordinates (r, ) to lie in MB , i.e. the label B to be t since both t and
are then xed on each of these surfaces:

MB = Mt = {p M / x0 (p) = t, x3 (p) = }. (2.17)

The 2-surfaces MB are called the meridional surfaces. In case of orthogonal transitivity, the
following metric components are identically zero, to reect the orthogonality between Sr and
Mt :
gtr = 0, gt = 0, gr = 0, g = 0. (2.18)
The following theorem states under which conditions this is guaranteed:
Generalized Papapetrou theorem: a stationary (Killing vector ) ~ and axisym-
metric (Killing vector
~ ) spacetime ruled by the Einstein equation is circular i the
energy-momentum tensor T obeys to

T [ ] = 0 (2.19)
T [ ] = 0, (2.20)

where the square bracket denote a full antisymmetrization.


This theorem has been demonstrated in the case of vacuum solutions (T = 0) by Papapetrou
(1966) [78] and extended to the non-vacuum case by Kundt & Trmper (1966) [63] and Carter
(1969) [28, 30] (see also 7.1 of Walds textbook [96], 7.2.1 of Straumanns one [90], or 19.2
of [86]). Note that Eq. (2.19)-(2.20) are equivalent to

V ~
~ Vect(, ~) (2.21)
W ~
~ Vect(, ~ ), (2.22)

where V~ (resp. W~ ) is the vector whose components are V = T (resp. W = T ) and



~
Vect(, ~ ) denotes the vector plane generated by ~ and
~.
In the important case of a perfect uid source, the energy-momentum tensor takes the form
(1.37), so that

T [ ] = ( + p) u u[ ] + p [ ]
= ( + p) u u[ ] + p [ ]
| {z }
0
[ ]
= ( + p) u u .
2.2 Circular stationary and axisymmetric spacetimes 29

Similarly,
T [ ] = ( + p) u u[ ] .

Since ~ and u
~ are both timelike, we have u 6= 0. The circularity conditions (2.19)-(2.20) are
therefore equivalent to
u[ ] = 0, (2.23)
i.e. to
~
~ Vect(,
u ~) . (2.24)

Taking into account that ~ =


~t and
~ =~ , the above condition is equivalent to ur = 0 and

u = 0, or
 
~ = ut ~ +
u ~ , (2.25)

with
u d
:= t
= . (2.26)
u dt
The 4-velocity (2.25) describes a pure circular motion of the uid around the rotation axis,
hence the qualier circular given to spacetimes that obeys the orthogonal transitivity property.
In such case, there is no uid motion in the meridional surfaces, i.e. no convection. Stationary
and axisymmetric spacetimes that are not circular have been studied in Ref. [43].

In all what follows, we limit ourselves to circular spacetimes.

2.2.2 Quasi-isotropic coordinates


In a stationary and axisymmetric spacetime (M , g) that is circular, we may use adapted coor-
dinates (x ) = (t, r, , ) such that (r, ) span the 2-surfaces Mt which are orthogonal to the
surfaces of transitivity Sr , leading to the vanishing of the metric components listed in (2.18).
Moreover, we can always choose the coordinates (r, ) in each 2-surface Mt such that the metric
induced by g takes the form3

gab dxa dxb = A2 (dr 2 + r 2 d 2 ), (2.27)

where A = A(r, ). Indeed, dr 2 + r 2 d 2 is the line element of the 2-dimensional at metric in


polar coordinates and, in dimension 2, all the metrics are conformally related, meaning that
they dier only by a scalar factor A2 as in (2.27).
Remark : This is no longer true in dimension 3 or higher: in general one cannot write the metric line
element as a conformal factor times the flat one by a mere choice of coordinates.

Note that the choice of coordinates (r, ) leading to (2.27) is equivalent to the two conditions:

gr = 0 and g = r 2 grr . (2.28)

The coordinates (t, r, , ) with the above choice for (r, ) are called quasi-isotropic coordi-
nates (QI). The related cylindrical coordinates (t, , z, ), with := r sin and z := r cos ,
3
Indices a and b take their values in {1, 2}.
30 Stationary and axisymmetric spacetimes

are called Lewis-Papapetrou coordinates, from the work of Lewis (1932) [67] and Papapetrou
(1966) [78].
Let us dene the scalar function = (r, ) as minus the scalar product of the two Killing
vectors ~ and
~ , normalized by the scalar square of
~:

~
~
:= . (2.29)
~
~
As we will see below, the minus sign ensures that for a rotating star, 0 (cf. Fig. 3.3). Since
gt = ~
~ and g = ~ , we may write
~
gt = g . (2.30)
Besides, let us introduce the following function of (r, ):
g
B 2 := 2 2 (2.31)
r sin
Collecting relations (2.18), (2.27), (2.30) and (2.31), we may write the components of the metric
tensor in the form
g dx dx = N 2 dt2 + A2 (dr 2 + r 2 d 2 ) + B 2 r 2 sin2 (d dt)2 , (2.32)

where N , A, B and are four functions of (r, ):

N = N (r, ) , A = A(r, ) , B = B(r, ) , = (r, ) . (2.33)


Equivalently, in matrix form:

N 2 + B 2 2 r 2 sin2 0 0 B 2 r 2 sin2
0 A2 0 0
g = . (2.34)
0 0 A2 r 2 0
B 2 r 2 sin2 0 0 B 2 r 2 sin2
The inverse of this matrix is

N 2 0 0 /N 2
0 A2 0 0
g = 2 2
. (2.35)
0 0 A r 0
/N 2 0 0 (Br sin )2 2 /N 2
To prove it, it suces to check that relation (1.20) holds.
The local atness of spacetime implies that the metric functions A and B must coincide on
the rotation axis:
A| = B| . (2.36)
Indeed, let us consider a small circle around the rotation axis at a xed value of both t and ,
with 1, and centered on the point of coordinates (x ) = (t, r, 0, 0) (cf. Fig. 2.6). According
to the line element (2.32), the metric length of its radius is R A(r, 0) r, whereas its metric
circumference is C B(r/ cos , ) r sin 2 2B(r, 0) r. The local atness hypothesis
implies that C = 2R. Would this relation not hold, a conical singularity would be present on
the rotation axis. From the above values of C and R, we get A(r, 0) = B(r, 0), i.e. the property
(2.36) for the half part of corresponding to = 0. The demonstration for the second part
( = ) is similar.
2.2 Circular stationary and axisymmetric spacetimes 31

Figure 2.6: Radius R and circumference C of a small circle around the rotation axis ( 1).

2.2.3 Link with the 3+1 formalism


In terms of the 3+1 formalism introduced in 1.3, the comparison of (2.32) with (1.55) leads
immediately to (i) the identication of N as the lapse function, (ii) the following components of
the shift vector:
i = (0, 0, ) , (2.37)

and (iii) the following expression of the induced metric in the hypersurfaces t :

ij dxi dxj = A2 (dr 2 + r 2 d 2 ) + B 2 r 2 sin2 d2 . (2.38)

Therefore 
ij = diag A2 , A2 r 2 , B 2 r 2 sin2 (2.39)

and  
ij 1 1 1
= diag 2
, 2 2, 2 2 2 . (2.40)
A A r B r sin

Remark : Comparing with (2.34)-(2.35), note that gij = ij but g ij 6= ij .

~t = ~ and
Moreover, since in the present case ~ =
~ =
~ , relation (1.51) becomes

~ = N n
~
~ . (2.41)

In particular, if = 0, then ~ is colinear to n


~ and hence orthogonal to the hypersurfaces t .
According to the denition given in 2.1.1, this means that for = 0, the spacetime is static.
The reverse is true, assuming that in case of staticity, the axisymmetric action takes place in the
~ We may then state
surfaces orthogonal to .

= 0 (M , g) static . (2.42)
32 Stationary and axisymmetric spacetimes

Let us evaluate the extrinsic curvature tensor K of the hypersurfaces t (cf. 1.3.4). Since
ij /t = 0 and i = 0 except for = , Eq. (1.57) leads to
 
1 ij
Kij = j i i j .
2N x x
|{z}
0

Since (ij ) is diagonal [Eq. (2.39)] and / = 0, we conclude that all the components of K
are vanishing, except for
B 2 r 2 sin2
Kr = Kr = , (2.43)
2N r

B 2 r 2 sin2
K = K = . (2.44)
2N
In particular, the trace of K dened by (1.58) vanishes identically:
1 1 1
K := ij Kij = 2
Krr + 2 2 K + 2 2 2 K = 0. (2.45)
A |{z} A r |{z} B r sin |{z}
0 0 0

This is equivalent to the vanishing of the divergence of the unit normal n


~ to t [cf. Eq. (1.58)].
t is then of maximal volume with respect to nearby hypersurfaces (see e.g. 9.2.2 of Ref. [42] for
more details). This is fully analogous to the concept of minimal surfaces in a 3-dimensional Eu-
clidean space (the change from minimal to maximal being due to the change of metric signature).
For this reason, the foliation (t ) with K = 0 is called a maximal slicing.
The quadratic term Kij K ij which appears in the Hamiltonian constraint (1.60) is

Kij K ij = Kr K r + Kr K r + K K + K K = 2Kr K r + 2K K
= 2 rr (Kr )2 + 2 (K )2 ,

where to get the last line we used the fact that ij is diagonal. We thus get

B 2 r 2 sin2
Kij K ij = , (2.46)
2A2 N 2
with the following short-hand notation for any scalar elds u and v :

u v 1 u v
uv := + 2 . (2.47)
r r r
Chapter 3

Einstein equations for rotating stars

Contents
3.1 General framework . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

3.2 Einstein equations in QI coordinates . . . . . . . . . . . . . . . . . . . 34

3.3 Spherical symmetry limit . . . . . . . . . . . . . . . . . . . . . . . . . . 40

3.4 Fluid motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43

3.5 Numerical resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

3.1 General framework

We focus at present on the case of a single rotating star in equilibrium. The corresponding
spacetime (M , g) is then stationary and asymptotically at. In addition, it is reasonable to
assume that the spacetime is axisymmetric. This has been shown to be necessary for viscous
and heat-conducting uids by Lindblom (1976) [72]. For perfect uids, the general argument is
that a rotating non-axisymmetric body will emit gravitational radiation, and therefore cannot
be stationary. In this respect the situation is dierent from the Newtonian one, where stationary
non-axisymmetric rotating congurations do exist, the best known example being that of the
Jacobi ellipsoids.
In addition of being axisymmetric, we will assume that the spacetime is circular (cf. 2.2.1),
since this is relevant for a perfect uid rotating about the axis of symmetry [cf. Eq. (2.25)].
We may then use the QI coordinates (x ) = (t, r, , ) introduced in 2.2.2. Consequently, the
metric tensor is fully specied by the four functions N , A, B and listed in (2.33). In this
chapter we derive the partial dierential equations that these functions have to fulll in order
for the metric to obey the Einstein equation (1.24), as well as the equations of equilibrium for a
perfect uid. We shall also discuss the numerical method of resolution.
34 Einstein equations for rotating stars

3.2 Einstein equations in QI coordinates


3.2.1 Derivation
We shall derive the equations within the framework of the 3+1 formalism presented in 1.3.
Let us rst consider the trace of the 3+1 Einstein equation (1.59) in which we make use of the
Hamiltonian constraint (1.60) to replace 3 R + K 2 by 16E + Kij K ij , to get

K K  
i i = Di D i N + N 4(E + S) + Kij K ij . (3.1)
t x

In the present context, K = 0 and Kij K ij is given by Eq. (2.46). Besides, the Laplacian of N
can be evaluated via the standard formula:
 
i 1 ij N
Di D N = , (3.2)
xi xj

where is the square root of the determinant of the components ij of the metric . From
expression (2.39), we get

= A2 Br 2 sin . (3.3)

Taking into account expression (2.40) for ij , we conclude that (3.1) is equivalent to

2N 2 N 1 2N 1 N 2 B 2 r 2 sin2
+ + + = 4A N (E + S) + ln BN. (3.4)
r 2 r r r 2 2 r 2 tan 2N

Let us now consider the Hamiltonian constraint (1.60). On the right-hand side, we have
K2 = 0 and Kij K ij given by Eq. (2.46). There remains to compute the Ricci scalar 3 R. This
can be done by means of formulas (1.66)-(1.67). One gets
( "     #
3 2 1 2 A 1 A 1 A 2 1 2A 1 A 2
R = 2 + + 2 2 2
A A r 2 r r A r r r A
 2  )
1 B 3 B 1 2B 2 B
+ + + 2 + 2 . (3.5)
B r 2 r r r 2 r tan

The Hamiltonian constraint (1.60) thus becomes


"     #
1 2 A 1 A 1 A 2 1 2A 1 A 2 B 2 r 2 sin2
+ + +
A r 2 r r A r r 2 2 r 2 A 4N 2
 
1 2 B 3 B 1 2B 2 B
+ + + + = 8A2 E. (3.6)
B r 2 r r r 2 2 r 2 tan

Moving to the momentum constraint (1.61), we rst notice that Di K = 0. We may use the
standard formula to evaluate the divergence of a symmetric tensor (such as K):

1  j  1 jk jk
Dj K j i = K i K . (3.7)
xj 2 xi
3.2 Einstein equations in QI coordinates 35

In the present case, the last term always vanishes since K jk is non-zero only for non-diagonal
terms (jk=r, r, or ) and jk is diagonal [Eq. (2.39)]. Since in addition / = 0, we
are left with  
j 1 r  
Dj K i = ( K i ) + K i . (3.8)
r
Now, K r r = K r = 0 and K r = K = 0. We thus conclude that the rst two components of
the momentum constraint (1.61) reduce to
pr = 0 and p = 0. (3.9)
For the third component, we use Eq. (3.8) for i = , taking into account that K r = rj Kj =
rr Kr = A2 Kr and K = j Kj = K = A2 r 2 K with the values (2.43)-(2.44) for
Kr and K . We thus obtain
  3
  3

N sin 4 B 1 3 B N A2 p
r + sin = 16 . (3.10)
B 3 r 3 r N r r sin2 N B 2 r sin
Finally, let us consider the component of the 3+1 Einstein equation (1.59). We have
K = 0 and, according to Eq. (1.65),
k k
L~ K = k K +Kk +Kk = 0.
xk |{z}
0 |{z} |{z}
0 0

Thus the left-hand side of the component of Eq. (1.59) vanishes identically. On the right-hand
side, we have Dj N = N /xj ,
N 3 k N N 3 N
D D N = k = 3 r
x r
|{z}
0
2  1 B    
2 2 B 1 N 1 1 B 1 N
= r sin 2 + + 2 + ,
A B r r r r B tan
 2 
3 2 2 B B 3 B 1 2B 2 B
R = r sin 2 + + 2 + 2 .
A r 2 r r r 2 r tan
and
1 1 B 4 r 4 sin4
Kk K k = Kr K r + K K = (Kr )2
+ (K )2
= .
A2 A2 r 2 4A2 N 2
In addition, the matter term is
(S E) 2S = (S r r + S + S E) 2 S = r 2 sin2 B 2 (S r r + S S E).
Accordingly, after division by r 2 sin2 N B 2 /A2 , the component of the 3+1 Einstein equation
(1.59) writes
 
1 2 B 3 B 1 2B 2 B 1 B 2 r 2 sin2
+ + + + BN +
B r 2 r r r 2 2 r 2 tan BN 2N 2
   
1 N 1 N
+ + = 4A2 S r r + S S E . (3.11)
rN r r tan
36 Einstein equations for rotating stars

At this stage, we have four equations, (3.4), (3.6) (3.10) and (3.11), for the four unknowns
N , A, B and . It is forth to perform some linear combinations of these equations to let appear
classical elliptic operators. Firstly, multiplying Eq. (3.4) by B, Eq. (3.11) by N B and adding
the two yields

2 3 1 2 2
2
(N B) + (N B) + 2 2
(N B) + 2 (N B) = 8N A2 B(S r r + S ). (3.12)
r r r r r tan
Secondly, dividing Eq. (3.4) by N , adding Eq. (3.6) and subtracting Eq. (3.11) yields

1 2N 1 N 1 2N 1 2A 1 A 1 2A
+ + + + +
N r 2 rN r r 2 N 2 A r 2 rA r r 2 A 2
2 2 2
1 3B r sin
= 8A2 S + 2 AA + . (3.13)
A 4N 2
Reorganizing slightly Eqs. (3.4), (3.10), (3.12) and (3.13), we obtain the nal system:

B 2 r 2 sin2
3 = 4A2 (E + S) + ( + ln B) (3.14)
2N 2

2
3 (r sin ) = 16 N A p + r sin ( 3 ln B)
(3.15)
B 2 r sin

2 [(N B 1)r sin ] = 8N A2 Br sin (S r r + S ) (3.16)

3B 2 r 2 sin2
2 (ln A + ) = 8A2 S + , (3.17)
4N 2
where the following abbreviations have been introduced:

:= ln N (3.18)
2 1 1 2
2 := 2 + + 2 2 (3.19)
r r r r
2 2 1 2 1
3 := 2 + + 2 2+ 2 (3.20)
r r r r r tan
3 := 3 1
. (3.21)
r 2 sin2
Besides, let us recall that terms such that have been dened by (2.47).
The operator 2 is nothing but the Laplacian in a 2-dimensional at space spanned by the
polar coordinates (r, ), whereas 3 is the Laplacian in a 3-dimensional at space, taking into
account the axisymmetry (/ = 0). A deeper understanding of why these operators naturally
occur in the problem is provided by the (2+1)+1 formalism developed in Ref. [43].

3.2.2 Boundary conditions


Equations (3.14)-(3.17) forms a system of elliptic partial dierential equations. It must be
supplemented by a set of boundary conditions. Those are provided by the asymptotic atness
3.2 Einstein equations in QI coordinates 37

hypothesis: for r +, the metric tensor tends towards Minkowski metric , whose components
in spherical coordinates are

dx dx = dt2 + dr 2 + r 2 d 2 + r 2 sin2 d2 . (3.22)

Comparing with (2.32), we get immediately the boundary conditions:




N 1

A 1
when r +. (3.23)

B 1

0

Remark : Spatial infinity is the only place where exact boundary conditions can be set, because the
right-hand sides of Eqs. (3.14)-(3.17) have a non-compact support, except for (3.16). This con-
trasts with the Newtonian case, where the basic equation is Poisson equation for the gravitational
potential [Eq. (3.43) below]:
3 = 4G. (3.24)
For a star, the right-hand side, involving the mass density , has clearly a compact support. The
general solution outside the star is then known in advance: it is the (axisymmetric) harmonic
function
X P (cos )
(r, ) = , (3.25)
r+1
=0
where P is the Legendre polynomial of degree . In this case, one can set the boundary conditions
for a finite value of r and perform some matching of and /r to determine the coefficients .

From the properties of the operators 3 and 3 and the boundary conditions (3.23), one can
infer the following asymptotic behavior of the functions N and :
 
K0 1
N =1+ +O , (3.26)
r r2
 
K1 1
= 3 +O , (3.27)
r r4
where K0 and K1 are two constants. In Chap. 4, we will show that

K0 = M and K1 = 2J, (3.28)

M and J being respectively the mass of the star and its angular momentum [cf. Eqs. (4.16) and
(4.41)].

3.2.3 Case of a perfect fluid


Since we are interested in rotating stars, let us write the matter source terms in the system
(3.14)-(3.17) for the case of a perfect uid [cf. Eq. (1.37)]. Due to the circularity hypothesis,
the uid 4-velocity is necessarily of the form (2.25): u~ = ut (~ +
~ ). We may use Eq. (2.41) to
~
express in terms of the unit timelike vector n ~ (the 4-velocity of the ZAMO) and ~ . We thus
get  
t 1
~ = (N u ) n
u ~ + ( )~ . (3.29)
N
38 Einstein equations for rotating stars

Now, according to Eq. (1.54), n


~ ~ = n = n = 0. Therefore, Eq. (3.29) constitutes an
orthogonal decomposition of the uid 4-velocity with respect to the ZAMO 4-velocity n
~:
 
~ = n
u ~ +U~ , (3.30)

with
= N ut and ~ = 1 ( )~
U . (3.31)
N
is the Lorentz factor of the uid with respect to the ZAMO and the vector U ~ is the uid
~
velocity (3-velocity) with respect to the ZAMO (see e.g. Ref. [41] for details). U is a spacelike
vector, tangent to the hypersurfaces t . Let us dene
B
U := ( )r sin . (3.32)
N
Since ~ = B 2 r 2 sin2 , we have
~
~ U
U ~ = U 2. (3.33)
From n ~ = 1 and n
~ n ~ = 0, Eq. (3.30) yields to the geometrical interpretation of the Lorentz
~ U
factor as (minus) the scalar product of the ZAMO 4-velocity with the uid one:
= ~
nu
~ . (3.34)
Besides, the normalization condition u ~ = 1 expressed in terms of (3.30) leads to the standard
~ u
relation
1/2
= 1 U2 . (3.35)
The quantities E, p , S r r , S and S which appear in the right-hand of Eqs. (3.14)-(3.17)
are computed by means of Eqs. (1.47)-(1.49), using the form (1.37) of the energy momentum
tensor. We get, using (3.30),
E = T n n = ( + p) u n u n +p g n n = 2 ( + p) p,
| {z } | {z } | {z }
1

p = T n = ( + p) u u n +p g n = 2 ( + p)U ,
| {z } | {z } | {z }
U 0

S = T = ( + p) u u +p g = 2 ( + p)U U + p .
| {z } | {z } | {z }
U U

Now, from Eq. (3.31), U = (0, 0, 0, U ), and from Eq. (3.33), U U = U 2 and U = Br sin U .
Besides r r = = = 1. Accordingly, the above results can be expressed as

E = 2 ( + p) p , (3.36)

p = B(E + p)U r sin , (3.37)

Srr = p , S = p , S = p + (E + p)U 2 , S = 3p + (E + p)U 2 . (3.38)


From the above expressions, the term E + S in the right-hand side of Eq. (3.14) can be written
1 + U2
E+S = ( + p) + 2p. (3.39)
1 U2
3.2 Einstein equations in QI coordinates 39

3.2.4 Newtonian limit


In the non-relativistic limit, tends to the Newtonian gravitational potential (divided by c2 ):


(3.40)
c2 Newt.

Moreover, Eq. (3.32) reduces to


U = r sin Newt. (3.41)
and U 2 1, p and c2 , where is the mass density. Accordingly, (3.39) gives

E + S c2 . (3.42)

Since in addition A 1 and all the quadratic terms involving gradient of the metric potentials
tend to zero, the rst equation of the Einstein system (3.14)-(3.17) reduces to (after restoring
the G and cs)
3 = 4G (3.43)
Newt.
The other three equations become trivial, of the type 0 = 0. We conclude that, at the non-
relativistic limit, the Einstein equations (3.14)-(3.17) reduce to the Poisson equation (3.43), which
is the basic equation for Newtonian gravity.

3.2.5 Historical note


The Einstein equations for stationary axisymmetric bodies have been written quite early. In
1917, Weyl [97] treated the special case of static bodies, i.e. with orthogonal Killing vectors
~ and ~ , or equivalently with = 0 [cf. Eq. (2.42)]. The eect of rotation was rst taken
into account by Lense and Thirring in 1918 [66], who derived the rst order correction induced
by rotation in the metric outside a spherical body. The rst full formulation of the problem
of axisymmetric rotating matter has been performed by Lanczos in 1924 [64] and van Stockum
in 1937 [89]. They wrote the Einstein equations in Lewis-Papapetrou coordinates (cf. 2.2.2)
and obtained an exact solution describing dust (i.e. pressureless uid) rotating rigidly about
some axis. However this solution does not correspond to a star since the dust lls the entire
spacetime, with an increasing density away from the rotation axis. In particular the solution is
not asymptotically at. Rotating stars have been rst considered by Hartle and Sharp (1967)
[55], who formulated a variational principle for rotating uid bodies in general relativity. Hartle
(1967) [52] and Sedrakian & Chubaryan (1968) [82] wrote Einstein equations for a slowly rotating
star, in coordinates dierent from the QI ones1 . For rapidly rotating stars, a system of elliptic
equations similar to (3.14)-(3.17) has been presented (in integral form) by Bonazzola & Machio
(1971) [20]. It diers from (3.14)-(3.17) by the use of ln gtt instead of ln N = . Bardeen
& Wagoner (1971) [12] presented a system composed of three elliptic equations equivalent to
Eqs. (3.14), (3.15) and (3.16), supplemented by a rst order equation of the form

(ln A + ) = S(r, ), (3.44)

where S(r, ) is a complicated expression containing rst and second derivatives of , A, B
and and no matter term [see Eq. (4d) of Ref. [27] (9 lines !)]. Equation (3.44) is fact not
1
Namely coordinates for which g = g / sin2 instead of (2.28).
40 Einstein equations for rotating stars

derived from the Einstein equation (1.24) but is an identity stating that the Ricci tensor satises
Rrr = R /r 2 and Rr = 0. Note that these identities are veried only for an isotropic energy-
momentum tensor, such as the perfect uid one [Eq. (1.37)]. For anisotropic ones (like that of an
electromagnetic eld), Eq. (3.44) should be modied so as to include the anisotropic part of the
energy-momentum tensor on its right-hand side. The full system of the four elliptic equations
(3.14)-(3.17) has been rst written by Bonazzola, Gourgoulhon, Salgado & Marck (1993)2 [19].

3.3 Spherical symmetry limit


3.3.1 Taking the limit
In spherical symmetry, the metric components g simplify signicantly. Indeed, we have seen
that A and B must be equal on the rotation axis [Eq. (2.36)]. But, if the star is spherically
symmetric, there is no privileged axis. Therefore, we can extend (2.36) to all space:

A = B spher. sym. (3.45)

Besides, to comply with spherical symmetry, the parameter in the decomposition (2.25) of the
uid 4-velocity must vanish. Otherwise, there would exist a privileged direction around which
the uid is rotating. Hence we have

= 0 spher. sym. (3.46)

and u ~ From Eqs. (3.37) and (3.32) with = 0, we have


~ is colinear to the Killing vector .
p . Equation (3.15) is then a linear equation in . It admits a unique solution which is
well-behaved in all space:
= 0 spher. sym. (3.47)
According to (2.42), this implies that the spacetime is static.
Properties (3.46) and (3.47), once inserted in Eq. (3.32) leads to the vanishing of the uid
velocity with respect to the ZAMO:

U = 0 spher. sym. (3.48)

Thanks to (3.45) and (3.47), the spacetime metric (2.32) becomes

g dx dx = N (r)2 dt2 + A(r)2 (dr 2 + r 2 d 2 + r 2 sin2 d2 ) . (3.49)


spher. sym.

The coordinates (t, r, , ) are in this case truly isotropic, the spatial metric being conformal to
a at one. This justify the name quasi-isotropic given to these coordinates in the general case.
Since U = 0 implies = 1, the Einstein equations (3.14)-(3.17) with a perfect uid source
and spherical symmetry reduce to

d2 2 d d d
2
+ = 4A2 ( + 3p) ( + ln A) (3.50)
dr r dr dr dr spher. sym.
2 = (AB)1/4 and B
In the original article [19], the functions A = (B/A)1/2 were employed instead of A and B.
The equations with the present values of A and B are given in Ref. [45].
3.3 Spherical symmetry limit 41

d2 3 d
(N A) + (N A) = 16N A3 p (3.51)
dr 2 r dr spher. sym.
 2
d2 1 d d
(ln A + ) + (ln A + ) = 8A2 p (3.52)
dr 2 r dr dr
spher. sym.

We have omitted Eq. (3.15) since, as seen above, it is trivially satised in spherical symmetry.
Besides, we have used the equivalent form (3.12) of Eq. (3.16).

3.3.2 Link with the Tolman-Oppenheimer-Volkoff system


The reader familiar with static models of neutron stars might have been surprised that the spheri-
cal limit of the Einstein equations (3.14)-(3.17) does not give the familiar Tolman-Oppenheimer-
Volkoff system (TOV). Indeed, the TOV system is (see e.g. [50, 53]):

dm
= 4r2 (3.53)
d
r
   
d 2m( r ) 1 m(r)
= 1 + 4
rp (3.54)
d
r r r2
dp d
= ( + p) , (3.55)
d
r dr
which is pretty dierent from the system (3.50)-(3.52). In particular, the latter is second order,
whereas (3.53)-(3.55) is clearly rst order. However, the reason of the discrepancy is already
apparent in the notations used in (3.53)-(3.55) : the radial coordinate is not the same in the
two system: in (3.50)-(3.52), it the isotropic coordinate r, whereas in the TOV system, it is the
areal radius r. The latter is such that the metric components write
 1
2 22m( r)
g d x = N (
x d r) dt + 1 r 2 + r2 d 2 + r2 sin2 d2 .
d (3.56)
r

This form is dierent from the isotropic one, given by (3.49). Note however that the coordinates
= (t, r, , ) dier from the isotropic coordinates x = (t, r, , ) only in the choice of r. From
x
(3.56), it is clear that the area of a 2-sphere dened by t = const and r = const is given by

r2,
A = 4 (3.57)

hence the qualier areal given to the coordinate r. On the contrary, we see from (3.49) that the
area of a 2-sphere dened by t = const and r = const is

A = 4A(r)2 r 2 , (3.58)

so that there is no simple relation between the area and r.


Outside the star, the solution is Schwarzschild metric (see 3.3.3) and the coordinates (x )
are nothing than the standard Schwarzschild coordinates. In particular, outside the star m( r) =
M = const, where M is the gravitational mass of the star and N ( r)2 = (1 2M/r ), so that
we recognize in (3.56) the standard expression of Schwarzschild solution. The expression of this
42 Einstein equations for rotating stars

solution the isotropic coordinates (x ) used here is derived in 3.3.3. The relation between the
two radial coordinates is given by
 2
M
r = r 1 + (outside the star). (3.59)
2r

Far from the star, in the weak eld region (M/r 1), we have of course r r.
There exists two coordinate systems which are some extension of the TOV areal coordinates
to the rotating case:

the coordinates employed by Hartle & Thorne [52, 56] and Sedrakyan & Chubaryan [82] to
compute stellar models in the slow rotation approximation (metric expanded to the order
2 );

the so-called radial gauge introduced by Bardeen & Piran [11]; however the entire hypersur-
face t cannot be covered in a regular way by these coordinates, as shown in Appendix A
of Ref. [19].

3.3.3 Solution outside the star


Outside the star, = 0 and p = 0. Equation (3.51) reduces then to
 
d2 3 d 1 d 3 d
(N A) + (N A) = 3 r (N A) = 0,
dr 2 r dr r dr dr

from which we deduce immediately that

d 2a
(N A) = 3 , with a = const.
dr r
The factor 2 is chosen for latter convenience. The above equation is easily integrated, taken into
account the boundary conditions (3.23):
a
NA = 1 . (3.60)
r2
Besides, subtracting Eq. (3.50) from Eq. (3.52), both with = 0 and p = 0, we get the following
equation outside the star:

d2 1 d 1 d d d
2
ln A + ln A = ln A. (3.61)
dr r dr r dr dr dr
Now Eq. (3.60) gives, since = ln N ,

d d 2a
= ln A + .
dr dr r(r 2 a)

Inserting this relation into Eq. (3.61), we get

d2 A 2(r 2 2a) dA 2a
2
+ 2
2 2 A = 0. (3.62)
dr r(r a) dr r (r a)
3.4 Fluid motion 43

The reader can check easily that the unique solution of this linear ordinary dierential equation
which satises to the boundary condition (3.23) requires a 0 and is
 2
a
A= 1+ . (3.63)
r

Relation (3.60) yields then


  1
a a
N= 1 1+ . (3.64)
r r

The constant a is actually half the gravitational mass of the star. Indeed the asymptotic
expansion for r of the above expression is
   
a a 2 a
N 1 1 1 .
r r r

Comparing with Eqs. (3.26) and (3.28), we get


M
a= . (3.65)
2
Replacing a, A and N by the above values in (3.49), we obtain the following form of the metric
outside the star:
!2  
1 M 2 M 4 2 
g dx dx = 2r
M
dt + 1 + dr + r 2 (d 2 + sin2 d2 ) . (3.66)
1 + 2r 2r

We recognize the Schwarzschild metric expressed in isotropic coordinates. Actually, we have


recovered the Birkhoff theorem 3 : in spherical symmetry, the only solution of Einstein equation
outside the central body is the Schwarzschild solution.

Remark : There is no equivalent of the Birkhoff theorem for axisymmetric rotating bodies: for black
holes, the generalization of the Schwarzschild metric beyond spherical symmetry is the Kerr metric,
and the generic metric outside a rotating star is not the Kerr metric. Moreover, no fluid star has
been found to be a source for the Kerr metric (see e.g. [74]). The only matter source for the Kerr
metric found so far is made of two counterrotating thin disks of collisionless particles [15].

3.4 Fluid motion


Having discussed the gravitational eld equations (Einstein equation), let us now turn to the
equation governing the equilibrium of the uid, i.e. Eq. (1.36).

3.4.1 Equation of motion at zero temperature


We consider a perfect uid at zero temperature, which is a very good approximation for a
neutron star, except immediately after its birth. The case of nite temperature has been treated
3
More precisely, we have established Birkhoff theorem under the hypothesis of stationarity; however this
hypothesis is not necessary.
44 Einstein equations for rotating stars

by Goussard, Haensel & Zdunik (1997) [46]. The energy-momentum tensor has the form (1.37)
and, thanks to the zero temperature hypothesis, the equation of state (EOS) can be written as
= (nb ) (3.67)
p = p(nb ), (3.68)
where nb is the baryon number density in the uid frame.
The equations of motion are the energy-momentum conservation law (1.36) :
T = 0 (3.69)
and the baryon number conservation law:
(nb u ) = 0. (3.70)
Inserting the perfect uid form (1.37) into Eq. (3.69), expanding and projecting orthogonally to
~ [via the projector given by (1.30)], we get the relativistic Euler equation:
the uid 4-velocity u
( + p)u u + ( + u u ) p = 0. (3.71)
Now the Gibbs-Duhem relation at zero temperature states that
dp = nb d, (3.72)
where is the baryon chemical potential, := d/dnb . Moreover, thanks to the rst law of
Thermodynamics at zero temperature (see e.g. Ref. [41] for details), is equal to the enthalpy
per baryon h dened by
+p
h := . (3.73)
nb
Thus we may rewrite (3.72) as dp = nb dh, hence
p = nb h.
Accordingly, Eq. (3.71) becomes, after division by nb ,
hu u + ( + u u ) h = 0,
which can be written in the compact form

u (hu ) + h = 0 . (3.74)

This equation is more useful than the original Euler equation (3.71). It leads easily to the
relativistic generalization of the classical Bernoulli theorem ( 3.4.2) and to a rst integral of
motion for a rotating star ( 3.4.3).
Remark : Thanks to the properties u u = 1 and u u = 0 (the latter being a consequence of the
former), Eq. (3.74) can be rewritten in the equivalent form
u [ (hu ) (hu )] = 0. (3.75)
It makes appear the antisymmetric bilinear form := (hu ) (hu ), called the vorticity
2-form. The fluid equation of motion (3.75) has been popularized by Lichnerowicz [70, 71] and
Carter [31], when treating relativistic hydrodynamics by means of Cartans exterior calculus (see
e.g. Ref. [41] for an introduction).
3.4 Fluid motion 45

3.4.2 Bernoulli theorem


Equation (3.74) is an identity between 1-forms. Applying these 1-forms to the stationarity Killing
vector ~ (i.e. contracting (3.74) with ), we get successively

u (hu ) + h = 0,
| {z }
0
u (hu ) hu u = 0,
u (hu ) h u u = 0,
| {z }
0

u (hu ) = 0.

Note that h = 0 stems from the stationarity of the uid ow and u u = 0 from
Killing equation (2.4). We thus have

~ =0.
u )
u~ (h~ (3.76)

In other words, the scalar quantity h~ u ~ is constant along any given uid line. This results
constitutes a relativistic generalization of the classical Bernoulli theorem. To show it, let us rst
recast Eq. (3.76) in an alternative form. Combining Eq. (3.30) with Eq. (2.41), we get (using
~ n
n ~ = 1, n ~ ~ = 0 and n ~ = 0)
~ U
 
u ~ = h(~
h~ n+U ~ ) (N n
~ ~ ) = hN 1 + ~ .
~ U
N
Accordingly, by taking the logarithm of h~ ~ we obtain that property (3.76) is equivalent to
u ,
 
H + + ln + ln 1 + ~ = const along a uid line ,
~ U (3.77)
N
where we have let appear the log-enthalpy
 
h
H := ln , (3.78)
mb

mb being the mean baryon mass : mb 1.66 1027 kg. Dening the uid internal energy
density by
int := mb nb (3.79)
we have [cf. (3.73)]  
int + p
H = ln 1 + . (3.80)
mb n b
At the Newtonian limit,
int + p
1,
mb n b
and mb nb (the mass density), so that H tends towards the (non-relativistic) specic enthalpy:

int + p
H . (3.81)

Newt.
46 Einstein equations for rotating stars

Besides, thanks to (3.35) and U 2 1, we have ln U 2 /2. Moreover we have already seen
that tends towards the Newtonian gravitational potential [Eq. (3.40)]. In addition, at the
Newtonian limit, = 0. Consequently, the Newtonian limit of (3.77) is

U2
H ++ = const along a uid line , (3.82)
2 Newt.
which is nothing but the classical Bernoulli theorem.
The general relativistic Bernoulli theorem (3.76) has been rst derived by Lichnerowicz in
1940 [68, 69]. The special relativistic version had been obtained previously by Synge in 1937
[91].
~ gives rise to a conserved quantity
Remark : That a symmetry (here stationarity, via the Killing vector )
[Eq. (3.77)] is of course a manifestation of the Noether theorem. We will encounter another one in
4.6.1.

To establish (3.76) we have used nothing but the fact that the ow obeys the symmetry and
~ is a Killing vector. For a ow that is axisymmetric, we thus have the analogous property
u
u~ (h~ ~) = 0 . (3.83)
The quantity h~ ~ is thus conserved along the uid lines. Thanks to (3.30) and n
u ~ = 0, we
~
may write it as
h~u~ = h~ U~ = hB 2 r 2 sin2 U . (3.84)
At the Newtonian limit, h 1, 1, B 1 and U , so that the conserved quantity
along each uid line is the specic angular momentum about the rotation axis, r 2 sin2 .

3.4.3 First integral of motion


In the derivation of the Bernoulli theorem, we have assumed a general stationary uid. Let us
now focus to the case of a rotating star, namely of a circular uid motion around the rotation
axis. The uid 4-velocity takes then the form (2.25):

~ = ut~k ,
u (3.85)
with
~k := ~ +
~ . (3.86)
Note that if is a constant, ~k is a Killing vector: it trivially satises Killing equation since ~
and ~ both do. u~ is then colinear to a Killing vector and one says that the star is undergoing
rigid rotation. If is not constant, the star is said to be in differential rotation.
Let us evaluate the Lie derivative of the 1-form hu along the vector eld ~k. According to
Eq. (A.10), we have
L~k hu = k (hu ) + hu k
= k (hu ) + (hu k ) k (hu )
= (ut )1 [u (hu ) u (hu )] + (hu k )
= (ut )1 [u (hu ) u u h h u u ] + (hu k )
| {z } | {z }
1 0
t 1
= (u ) [u (hu ) + h] + (hu k ). (3.87)
3.4 Fluid motion 47

Besides, from Eqs. (3.86) and (A.10), we have


L~k hu = L~ hu + L~ hu +hu = hu , (3.88)
| {z } | {z }
0 0

where the properties L~ hu = 0 and L~ hu = 0 reect respectively the stationarity and the
axisymmetry of the ow. Combining Eqs. (3.87) and (3.88), we get
(hu k ) hu = (ut )1 [u (hu ) + h] ,

u ~k,
i.e., after division by hu k = h~
~
u ~ 1
u ~k)
ln(h~ = [u~ (hu) + h] . (3.89)
~ ~k
u u ~k
ut h~
~ and ~k are
Remark : The minus in the argument of the logarithm is justified by the fact, that since u
~
~ k < 0.
both future directed timelike vectors, u
Comparing the right-hand side with the equation of motion (3.74), we conclude that a necessary
and sucient condition for the relativistic Euler equation to hold is

~
u ~
u ~k)
ln(h~ = 0 . (3.90)
~
~ k
u
The integrability condition of this equation is either (i) = const (rigid rotation) or (ii) the
factor in front of is a function of :
~
u ~
= F (). (3.91)
~
~ k
u
Let us discuss these two cases successively.

Rigid rotation
If = const, then = 0 and Eq. (3.90) leads immediately to the rst integral of motion

u ~k = const .
h~ (3.92)
Thanks to (3.85) and (3.31), we have
1 N
~ ~k = (ut )
u u
~ {z
|u ~} = , (3.93)
|{z}
/N 1

so that taking the logarithm of the rst integral (3.92) leads to

H + ln = const . (3.94)
This rst integral of motion has been obtained rst in 1965 by Boyer [22] for incompressible uids4
and generalized to compressible uids by Hartle & Sharp (1965) [54] and Boyer & Lindquist
(1966) [24]. Note the dierence of sign in front of ln with respect to the Bernoulli theorem
(3.77).
4
More precisely, the first integral obtained by Boyer is (within our notations) N 2 (1 U 2 ) = const/h2 . Taking
the square root and the logarithm, we obtain (3.94).
48 Einstein equations for rotating stars

Remark : Another important difference is that (3.77) provides a constant along each fluid line, but
this constant may vary from one fluid line to another one, whereas (3.94) involves a constant
throughout the entire star. However, (3.94) requires the fluid to be in a pure circular motion, while
the Bernoulli theorem (3.77) is valid for any stationary flow.

The Newtonian limit of (3.94) is readily obtained since and ln U 2 /2 (cf. 3.4.2);
it reads
U2
H + = const . (3.95)
2 Newt.
Again notice the change of sign in front of U 2 /2 with respect to the Bernoulli expression (3.82).
Notice also that U 2 = 2 r 2 sin2 [Eq. (3.41)] and that
1
tot := 2 r 2 sin2 (3.96)
2
is the total potential (i.e. that generating the gravitational force and the centrifugal one) in the
frame rotating at the angular velocity about the rotation axis.

Differential rotation
Equation (3.90) gives, thanks to (3.78), (3.91) and (3.93),

(H + ln ) + F () = 0.

We deduce immediately the rst integral of motion :


Z
H + ln + F ( )d = const . (3.97)
0

Remark : The lower boundary in the above integral has been set to zero but it can be changed to any
value: this will simply change the value of the constant in the right-hand side.

Let us make explicit Eq. (3.91), by means of respectively Eqs. (3.93), (3.30), (3.31), (3.35) and
(3.32):
~
u ~ ~) 2 ~ 2
F () = = (~ n+U ~= U ~ = 2 ( )~ ~
u~ ~k N N N
1 2 2 2 B 2 ( )r 2 sin2
= ( )B r sin = .
N 2 (1 U 2 ) N 2 [1 (B/N )2 ( )2 r 2 sin2 ]
Hence
B 2 ( )r 2 sin2
F () = . (3.98)
N 2 B 2 ( )2 r 2 sin2
At the Newtonian limit, this relation reduces to

F () = r 2 sin2 (3.99)
Newt.
and (3.97) to
Z
U2
H + + F ( )d = const . (3.100)
2 0 Newt.
3.5 Numerical resolution 49

Equation (3.99) implies that must be a function of the distance from the rotation axis:

= (r sin ) . (3.101)
Newt.

We thus recover a well known result for Newtonian dierentially rotating stars, the so-called
Poincar-Wavre theorem (see e.g. 3.2.1 of Ref. [93]). In the general relativistic case, once
some choice of the function F is made, one must solve Eq. (3.98) in to get the value of at
each point. The rst integral (3.97) has been rst written5 by Boyer (1966) [23]. The relation
(3.98) has been presented by Bonazzola & Machio (1971) [20].

3.4.4 Stellar surface and maximum rotation velocity


Let us consider a rigidly rotating star and rewrite the rst integral of motion (3.94) as

H = Hc + c + ln , (3.102)

where Hc and c are the values of H and at r = 0 [we have ln c = 0 since U = 0 at r = 0 by


Eq. (3.32)]. The surface of the star is dened by the vanishing of the pressure:

stellar surface: p=0, (3.103)

which expresses the local equilibrium with respect to the ambient vacuum. Let H0 be the value
of H corresponding to p = 0 (usually, H0 = 0). As a solution of the Poisson-like equation (3.14),
is an increasing function of r, from c < 0 to = 0 at r = +. For instance, at the Newtonian
limit and in spherical symmetry, = M/r outside the star. Therefore is a decreasing
function of r and if ln = 0 (no rotation), (3.102) implies that H is a decreasing function of r as
well. Provided that Hc > H0 , it always reaches H0 at some point, thereby dening the surface
of the star. If the star is rotating, ln is an increasing function of r. It takes its maximum value
in the equatorial plane = /2. For instance, for a Newtonian star, ln = (2 /2)r 2 sin2 . If
is too large, it could be that ln compensate the decay of in such a way that H remains
always larger than H0 . Therefore, for a given value of Hc , there exists a maximum rotation
speed. Physically this corresponds to the uid particles reaching the orbital Keplerian velocity
at the stellar equator. For this reason, the maximal value of is called the Keplerian limit
and denoted K .

3.5 Numerical resolution


Having established all the relevant equations in the preceding sections, let us now discuss briey
their numerical resolution to get a model of a rotating star.

3.5.1 The self-consistent-field method


To be specic, we consider rigidly rotating stars: = const, but the following discussion can be
adapted to take into account dierential rotation. A standard algorithm to numerically construct
a rotating stellar model is the following one. Choose
5
in the equivalent form N 2 (1 U 2 ) = [1 + ()]2 /h2 , the link between the functions () and F () being
R
ln[1 + ()] = 0 F ( )d
50 Einstein equations for rotating stars

a barotropic EOS, of the form


= (H) and p = p(H), (3.104)
assuming that H = 0 corresponds to p = 0;
some central value Hc of the log-enthalpy H;
some value for the constant angular velocity .
In addition assume some values for the four metric functions N , , A and B. This may be crude
values, such as the at spacetime ones: N = A = B = 1 and = 0. Besides, initialize U to zero
and the enthalpy to a crude prole, like H = Hc (1 r 2 /R2 ) where R is some radius chosen, also
crudely, around the expected radius of the nal model. Then
1. Via the EOS (3.104), evaluate and p. The surface of the star is then dened by p = 0.
2. From U , , and p, compute via Eqs. (3.36)-(3.38) the matter sources E, p , S, S r r , S ,
S that appear in the right-hand side of Einstein equations.
3. Solve the Einstein equations (3.14)-(3.17) as at Poisson-like equations, using the current
value of , , A and B on their right-hand sides, thereby getting new values of N = exp ,
, A and B.
4. Compute the velocity eld from Eq. (3.32): U = (B/N )( )r sin ; evaluate the Lorentz
factor via Eq. (3.35): = (1 U 2 )1/2 .
5. Use the rst integral of uid motion in the form (3.102):
H = Hc + c + ln
to compute H in all space. Go to step 1.
In practise, this method leads to a unique solution for a given value of the input parameters
(Hc , ) and a xed EOS.
The above method is called the self-consistent-field method (SCF) and has been intro-
duced for computing Newtonian rotating stars by Ostriker & Mark (1968) [77] (see Ref. [80] for
some historical account). The rst application of this method to compute rapidly rotating stellar
models in full general relativity, along the lines sketched above is due to Bonazzola & Maschio
(1971) [20].
Remark : The procedure described above is only a sketch of the actual algorithm implemented in a
numerical code; the latter is generally more complicated, involving some relaxation, progressive
increase of from null value, convergence to a given mass, etc.

The mathematical analysis of the self-consistent-eld method applied to the Einstein equa-
tions (3.14)-(3.17) has been performed by Schaudt & Pster [84, 83, 79]. They proved the
exponential convergence of the method and that the exterior and interior Dirichlet problem for
relativistic rotating stars is solvable, under some conditions on the boundary data. The full
mathematical demonstration of existence of rotating uid stars in general relativity has been
given by Heilig in 1995 [57], by a method dierent from the self-consistent-eld, and only for
suciently small angular velocities . Let us stress that, up to now, we do not know any exact
solution describing both the interior and the exterior of a rotating star in general relativity (see
e.g. [73, 85] for a discussion).
3.5 Numerical resolution 51

3.5.2 Numerical codes


The rst numerical models of rotating relativistic stars have been computed by Hartle & Thorne
(1968) [56] in the approximation of slow rotation (cf. 3.2.5). The rst rapidly rotating models
have been obtained by Bonazzola & Maschio (1971) [20] using a formulation very close to that
presented here, namely based on Lewis-Papapetrou coordinates (the cylindrical version of QI
coordinates, cf. 2.2.2) and applying the self-consistent-eld method to a system of equations

similar to (3.14)-(3.17) (basically using the variable gtt instead of N ). All the numerical
codes developed since then have been based on the self-consistent-eld method. In 1972, Wilson
[98] used the Bardeen-Wagoner formulation [12] (cf. 3.2.5) to compute models of dierentially
rotating stars without an explicit EOS. In 1974 Bonazzola & Schneider (1974) [21] improved
Bonazzola & Maschios code and computed the rst models models of rapidly rotating relativistic
stars constructed on an explicit EOS (an ideal Fermi gas of neutron). In 1975, Butterworth
& Ipser [26, 27] developed a code based on the Bardeen-Wagoner formulation and computed
homogeneous rotating bodies [26, 27] as well as polytropes [25]. Friedman, Ipser & Parker (1986,
1989) [39, 40] applied their method to some realistic (i.e. based onto detailed microphysics
calculations) EOS. This study has been extended by Lattimer et al. (1990) [65] to include more
EOS.
In 1989, Komatsu, Eriguchi & Hachisu (KEH) [61] developed a new code, also based on the
Bardeen-Wagoner formulation, and applied it to dierentially rotating polytropes [62]. Their
method has been improved by Cook, Shapiro & Teukolsky (CST) (1992,1994) [34, 35], by intro-
ducing the radial variable s = r/(r + req ), where req is the r coordinate of the stellar equator,
in order that the computational domain extends to spatial innity the only place where the
boundary conditions are known in advance [Eq. (3.23)]. In 1995, Stergioulas & Friedman [88]
developed their own code based on the CST scheme, improving the accuracy, leading to the
famous public domain code rns [105].
In 1993, Bonazzola, Gourgoulhon, Salgado & Marck (BGSM) [19] developed a new code based
a formulation very close to that presented here (cf. 3.2.5), using spectral methods [48], whereas
all previous codes employed nite dierences. This code has been improved in 1998 [18, 45] (by
introducing numerical domains adapted to the surface of the star) and incorporated into the
Lorene library, to become the public domain code nrotstar [104]. It is presented briey in
Appendix B.
In 1998, Nozawa et al. [76] have performed a detailed comparison of the KEH code, the rns
code and the BGSM code, showing that the relative dierence between rns and BGSM is of the
order of 104 (at most 103 in some extreme cases). The relative dierence between KEH and
BGSM is only of the order 103 to 102 . This is due to the nite computational domain used
by KEH, implying approximate boundary conditions for some nite value of r. On the contrary,
thanks to some compactication, both rns and BGSM are based on a computational domain
that extends to r = + where the exact boundary conditions (3.23) can be imposed.
A new multi-domain spectral code has been developed by Ansorg, Kleinwaechter & Meinel
(AKM) in 2002 [6, 4]. It is extremely accurate, leading to rotating stellar models to machine
accuracy (i. e. 1015 for 16-digits computations). It is has been used recently to compute
sequences of dierentially rotating stars, with a very high degree of dierential rotating, leading
to toroidal stellar shapes [5].
For a more extended discussion of the numerical codes, see the review article by Stergioulas
[87]. In particular, a comparison between the AKM, KEH, BGSM, rns and Lorene/rotstar
52 Einstein equations for rotating stars

Figure 3.1: Isocontours of the proper energy density in the meridional plane = 0 for the rotating
neutron star model discussed in 3.5.3. The coordinates (x, z) are defined by x := r sin and z := r cos .
The thick solid line marks the surface of the star.

codes can be found in Table 2 of that article.

3.5.3 An illustrative solution


In order to provide some view of the various elds in a relativistic rotating star, we present
here some plots from a numerical solution obtained by the Lorene/nrotstar code described in
Appendix B. The solution corresponds to a neutron star of mass M = 1.4 M rigidly rotating
at the frequency /(2) = 716 Hz. This value has been chosen for it is the highest one among
observed neutron stars: it is achieved by the pulsar PSR J1748-2446ad discovered in 2006 [58].
The EOS is the following one (see [50] for details):

for the core: the model A18+v+UIX* of Akmal, Pandharipande & Ravenhall (1998) [1],
describing a matter of neutrons, protons, electrons and muons via a Hamiltonian including
two-body and three-body interactions, as well as relativistic corrections;

for the inner crust: the SLy4 model of Douchin & Haensel [38];

for the outer crust: the Haensel & Pichon (1994) model [49], which is based on the exper-
imental masses of neutron rich nuclei.

A global (coordinate-based) view of the star is provided in Fig. 3.1 and some of its global
properties are listed in Table 3.1. All these quantities are properly dened in Chap. 4. The
metric functions N , and A are displayed in Figs. 3.2, 3.3 and 3.4 respectively. It appears that
the deviations from the at space values (N = 1, = 0 and A = 1) are maximal at the stellar
center. One may also check on these gures the tendency to fulll the boundary conditions (3.23)
when r +. Moreover, the faster decay of [in 1/r 3 , cf. Eq. (3.27)] with respect to that of
A [in 1/r, cf. (3.63)] is clearly apparent.
3.5 Numerical resolution 53

Gravitational mass M 1.400 M


Baryon mass Mb 1.542 M
Rotation frequency /(2) 716 Hz
Central log-enthalpy Hc 0.226 c2
Central proper baryon density nb,c 0.5301 fm3
Central proper energy density c 5.7838 nuc c2
Central pressure pc 0.8628 nuc c2
Coordinate equatorial radius req 9.867 km
Coordinate polar radius rp 8.649 km
Axis ratio rp /req 0.8763
Circumferential equat. radius Rcirc 12.08 km
Compactness GM/(c2 Rcirc ) 0.1711
Angular momentum J 0.7238 GM2 /c
Kerr parameter cJ/(GM 2 ) 0.3693
Moment of inertia I 1.417 1038 kg m2
Kinetic energy ratio T /W 0.0348
Velocity at the equator Ueq 0.1967 c
f
Redshift from equator, forward zeq 0.01362
Redshift from equator, backward zeq b 0.5529
Redshift from pole zp 0.2618

Table 3.1: Properties of the rotating neutron star model presented in 3.5.3;
M := 1.989 1030 kg, nuc := 1.66 1017 kg m3 .

1
=0
= /2

0.9
N

0.8

0.7

0.6
0 5 10 15 20 25 30
r [km]

Figure 3.2: Profile of the lapse function N in two different directions: = 0 (rotation axis, dashed line)
and = /2 (equatorial plane, solid line). The vertical solid line (resp. dashed line) marks the location
of the stellar surface in the equatorial plane (resp. along the rotation axis).
54 Einstein equations for rotating stars

0.5
=0
= /2

0.4

/ 0.3

0.2

0.1

0
0 5 10 15 20 25 30
r [km]

Figure 3.3: Same as Fig. 3.2 but for the shift vector component = .

1.6
=0
= /2
1.5

1.4
A

1.3

1.2

1.1

1
0 5 10 15 20 25 30
r [km]

Figure 3.4: Same as Fig. 3.2 but for the metric coefficient A.
3.5 Numerical resolution 55

0.0025

0.002

B-A 0.0015

0.001

0.0005

0
0 5 10 15 20 25 30
r [km]

Figure 3.5: Difference between the metric coefficients B and A in the equatorial plane. The vertical
solid line marks the location of the stellar surface in that plane.

Remark : Figures 3.2 to 3.4 have been truncated at r = 30 km for graphical convenience, but the
computational domain of the Lorene/nrotstar code extends to r = +.

The metric function B turns out to be very close to A: on the rotation axis, it is exactly equal
to A [cf. Eq. (2.36)] and, in the equatorial plane, the dierence B A is at most 2.3 103 , as
shown in Fig. 3.5.
Two uid quantities are displayed in Figs. 3.6 and 3.7: the proper energy density and the
uid velocity U with respect to the ZAMO [cf. Eqs. (3.32) and (3.33)]. In the rst gure, note
the presence of the neutron star crust, of width 0.4 km for = 0 and 0.6 km for = /2.
The prole of U shown in Fig. 3.7 is deceptively simple: it looks very close to the Newtonian
prole r [Eq. (3.41) with = /2], with one dierence however: the mean slope is not but
1.3 . This slope is approximately constant due to some numerical compensation of the two
factors B/N and (1 /) that appear in Eq. (3.32) once rewritten as U = (B/N )(1 /) r.
Indeed B/N is a decreasing function of r, taking the values 2.36 at r = 0 and 1.52 at r = req .
On the contrary, 1 / is an increasing function, taking the values 0.54 at r = 0 and 0.87 at
r = req . This results in (B/N )(1 /) taking almost the same values r = 0 and r = req (1.27
and 1.32 respectively), hence the straight behavior of the plot of U (r). Notice that at the stellar
surface, U is a non-negligible fraction of the velocity of light: U 0.2 c.
56 Einstein equations for rotating stars

6
=0

Proper energy density [nuc]


=/2
5

0
0 2 4 6 8 10
r [km]

Figure 3.6: Profile of the fluid proper energy density in two different directions: = 0 (rotation
axis, dashed line) and = /2 (equatorial plane, solid line). The small vertical solid line (resp. dashed
line) marks the location of the stellar surface in the equatorial plane (resp. along the rotation axis).
nuc := 1.66 1017 kg m3 .

0.2

0.15
U/c

0.1

0.05

0
0 2 4 6 8 10
r [km]

Figure 3.7: Profile of the norm U of the fluid velocity with respect to the ZAMO, in the equatorial
plane.
Chapter 4

Global properties of rotating stars

Contents
4.1 Total baryon number . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
4.2 Mass . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
4.3 Virial identities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
4.4 Angular momentum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63
4.5 Circumferential radius . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
4.6 Redshifts . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
4.7 Orbits around the star . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

4.1 Total baryon number


The baryon 4-current is
~b = nb u
~, (4.1)
where nb is the baryon number density in the uid frame and u
~ is the uid 4-velocity. For a stellar
model computed along the method sketched in 3.5.1, nb is obtained from the log-enthalpy H
and the EOS (3.104) via the formula

+p
nb = exp(H), (4.2)
mb

which is easily deduced from Eqs. (3.78) and (3.73). The baryon 4-current obeys the conservation
equation (3.70):
jb = 0. (4.3)
Thanks to this property, it can be shown, via the Gauss-Ostrogradski theorem, that the integral
Z

N := ~ dx1 dx2 dx3
~b n (4.4)
t
58 Global properties of rotating stars

is independent of the hypersurface t , i.e. of the slicing (t )tR of spacetime. It gives the total
number of baryons in the star. Note that, by denition of the baryon 4-current, the integrand,
nb = ~b n ~ , is nothing but the baryon number density as measured by the ZAMO, whose 4-
velocity is n ~ . Thanks to the relation = ~ nu ~ [Eq. (3.34)], we have nb = nb . Choosing

(x1 , x2 , x3 ) = (r, , ), we have in addition = A2 Br 2 sin [Eq. (3.3)], so that expression (4.4)
can be explicited as
Z
N = nb A2 Br 2 sin dr d d . (4.5)
t

We may convert N to a mass, by multiplying it by the mean baryon mass mb 1.66 1027 kg :

Mb := mb N . (4.6)

Mb is called the baryon mass of the star.

4.2 Mass
Due to the mass-energy equivalence, the mass of the star must involve all forms of energy,
including the gravitational eld energy. Now the latter is not localizable (as a consequence of
the Equivalence Principle): there is no energy density of the gravitational eld. This implies
that the mass of the star cannot be dened as a integral over t of some mass density as in
(4.4), which gives the N as the integral of the baryon density nb = ~b n
~ over t .
The are two major concepts of mass in general relativity (see Refs. [42], [92], [59] for more
details) : the ADM mass which applies to any asymptotically at spacetime and the Komar
mass which holds only for stationary spacetimes. In the present case, both coincide. We shall
examine rst the Komar mass denition.

4.2.1 Komar mass


Given the Killing vector ~ associated with stationarity, the Komar mass is
I
1
M := dS , (4.7)
8 S

where S is any closed 2-surface (sphere) surrounding the star and dS is the area-element
2-form normal to S . In the case where S is entirely contained in the hypersurface t and is
dened by r = const, it is spanned by the coordinates (, ) and

dS = (s n s n ) q d d, (4.8)

where the n s are given by (1.54), the s s are the covariant components of the unit normal
~s of S in (t , ) and q is the determinant of the components qab of the metric induced by
(or equivalently g) on S . From the diagonal form (2.39) of , we have ~s = sr ~r and sr is
determined by the normalization condition ~s ~s = 1. Hence

s = (0, A1 , 0, 0) and s = (0, A, 0, 0). (4.9)


4.2 Mass 59

Since S is dened by {t = const, r = const}, we read on (2.38) the following components of the
metric q on S : 
qab dxa dxb = r 2 A2 d 2 + B 2 sin2 d2 . (4.10)
Hence

q = ABr 2 sin . (4.11)
Remark : The Komar mass as defined by (4.7) is the flux integral through S of the type (2, 0) tensor
~ In dimension 3, there is locally only one direction ~s normal to a given sphere, so that instead
~ .

of the normal 2-form (4.8), one considers the normal 1-form dSi = si q d d; flux integrals are
then defined for vectors, i.e. type (1, 0) tensors. In the 4-dimensional spacetime (M , g), there
exists a whole plane of directions normal to the sphere and the vectors (~n, ~s) which appear in (4.8)
form a basis of that plane.

A priori the quantity M dened by (4.7) should depend on the choice of the surface S .
However, owing to the fact that ~ is a Killing vector, this is not the case, as long as S is located
outside the star. This was shown by Komar in 1959 [60]. To prove it, one may use the Gauss-
Ostrogradski theorem, the Killing equation (2.4) and the Einstein equation (1.24) to rewrite the
integral (4.7) as1 Z  
~ 1 ~
M =2 n, ) T n
T (~ ~ dr d d, (4.12)
t 2
where T is the energy-momentum tensor and T := g T its trace with respect to g. From
(4.12), it is clear that outside the support of T , i.e. outside the star, there is no contribution to
M , hence the independency of M upon the choice of S .
Expressing (4.7) in terms of 3+1 quantities, we get (cf. 7.6.2 of [42] for details),
I
1 
M= si Di N Kij si j q d d. (4.13)
4 S
In the present case, substituting (4.9) for si , (2.43)-(2.44) for Kij , (2.37) for j and (4.11) for

q, we have
I  
1 N B 2 r 2 sin2
M= Br 2 sin d d. (4.14)
4 S r 2N r
If we let the radius of the sphere extend to +, then the term containing does no longer
contribute to the integral, since (3.27) implies
 
2 1
r =O .
r r5
Taking into account B 1 when r +, we may then write
I
1 N 2
M= lim r sin d d , (4.15)
4 S S r

where the limit means that the integral is to be taken on a sphere S of radius r +. Plugging
the asymptotic behavior (3.26) of N into the above formula yields K0 = M , hence
 
M 1
N =1 +O . (4.16)
r r2
1
See e.g. 7.6.1 of [42] for the detailed demonstration.
60 Global properties of rotating stars

Let us now turn to the volume expression of M provided by formula (4.12). Introducing in it
the 3+1 decomposition (1.50) of the energy-momentum tensor, and making use of ~ = N n ~ + ~
~ ~
[Eq. (1.51)] with t = , we get
Z
 
M= N (E + S) 2pi i dr d d. (4.17)
t

In the present case, taking into account expression (3.37) for pi , (2.37) for i and (3.3) for ,
we obtain Z
M= [N (E + S) + 2B(E + p)U r sin ] A2 Br 2 sin dr d d . (4.18)
t

At the Newtonian limit, only the term in E remains in this expression:


Z
M= r 2 sin dr d d . (4.19)
t
Newt.

Of course, we recognize the expression of the mass as the integral of the mass density .

4.2.2 ADM mass


The ADM mass has been introduced by Arnowitt, Deser & Misner in 1962 [7] for any asymp-
totically at spacelike hypersurface t . It is expressible as [42, 59]
I h i
1
MADM = lim D j ij Di (f kl kl ) si q d d , (4.20)
16 S S

where (i) f is a ducial at 3-metric on t : fij = diag(1, r 2 , r 2 sin2 ) and f ij = diag(1, r 2 , r 2 sin2 )
and (ii) D stands for the connection associated with f .
Remark : Contrary to the Komar mass definition (4.7), the surface integral in (4.20) has to be taken
at infinity. Note also that the ADM mass does not require stationarity to be defined, contrary to
the Komar mass.

We leave as an exercise to the reader to show that with the form (2.39) of ij , the formula
(4.20) can be explicited as
I  
1 B 2 A2 2
MADM = lim (A2 + B 2 ) + r sin d d . (4.21)
16 S S r r

Although the two denitions look very dierent, Beig (1978) [14] and Ashtekar & Magnon-
Ashtekar (1979) [8] have shown that for a stationary spacetime, the ADM mass is always equal
to the Komar mass, provided that t is orthogonal to the Killing vector ~ at spatial innity,
which is our case here:
MADM = M . (4.22)
Therefore, for a stationary rotating star, the Komar mass and ADM mass coincide. This
common value, denoted here by M , is sometimes called gravitational mass, to distinguish it
from the baryon mass dened by (4.6).
4.3 Virial identities 61

4.2.3 Binding energy


The binding energy of the star is dened by as the dierence between the gravitational mass
and baryon mass:
Ebind := M Mb . (4.23)

Physically, |Ebind | is the energy one should spend to disperse all the particles (baryons) consti-
tuting the star to innity. The star is bound i Ebind < 0. If Ebind 0, the star would be
unstable and would explode, releasing the energy Ebind .

4.3 Virial identities


4.3.1 GRV3
Taking the identity of ADM and Komar mass as a starting point, Gourgoulhon & Bonazzola
(1994) [44] have derived a formula that generalizes the Newtonian virial theorem to general
relativity for stationary spacetimes. It is called GRV3, for General Relativistic Virial theorem,
the 3 recalling that the integral is a 3-dimensional one. For a rotating star, it reads [Eq. (43)
of [44]]
Z       
1 1 1 1 1 1 A 1 A
4S 2 AB + +
t A 2AB 2r A2 B 2 A r r tan
  2 2 
1 B 1 B 3B sin
+ + 2 2 2 A2 Br 2 sin dr d d = 0, (4.24)
2B r r tan 8r A N

with S given by (3.38) :


S = 3p + (E + p)U 2 .

At the Newtonian limit, S 3p + U 2 , /c2 , ln A /c2 , ln B /c2 , A 1, B 1


and = 0, so that, after division by 4, (4.24) becomes

2Ekin + 3P + Egrav = 0 , (4.25)


Newt.

where
Z
1
Ekin = U 2 r 2 sin dr d d (4.26)
2 t
Z
P = p r 2 sin dr d d (4.27)
t
Z
1 ~ 2 r 2 sin dr d d.
Egrav = () (4.28)
8G t

Ekin and Egrav are respectively the kinetic energy and the gravitational potential energy of the
star, and we recognize in (4.25) the classical virial theorem for stationary congurations (see e.g.
Eq. (2.180) of Ref. [93]).
The GRV3 identity proved to be very useful to check numerical solutions [76, 87].
62 Global properties of rotating stars

4.3.2 GRV2
Another virial identity has been obtained by Bonazzola (1973) [16] (see also Ref. [19, 17]).
Following [17], it is called GRV2, because contrary to GRV3, it involves a 2-dimensional integral.
The derivation of GRV2 is quite straightforward once we consider the last equation of the Einstein
system (3.14)-(3.17). The operator on the left-hand side of (3.17) is the 2-dimensional at
Laplacian 2 . It operates in the plane R2 spanned by the polar coordinates (r, ) and its Green
function is
1
G(~x, ~x ) = ln k~x ~x k, (4.29)
2
where x~ denotes the vector of R2 of coordinates (xa ) = (r, ) and k~x ~x k is the norm of ~x ~x
with respect to the at Euclidean metric of R2 . Accordingly, the generic solution of (3.17),
assuming its right-hand side is given, is
 
I 1
ln A + = ln r + O , (4.30)
2 r

where I is the integral of the right-hand side of (3.17),

3B 2 r 2 sin2
:= 8A2 S + , (4.31)
4N 2
over R2 with respect to the at area element r dr d:
Z r=+ Z =2
I= r dr d. (4.32)
r=0 =0

In the above integral, the coordinate spans the interval [0, 2], whereas the QI coordinate
runs in [0, ] only. This is because the meridional surface Mt covers only half of R2 . To dene
(4.32), we therefore use the analytical continuation of from [0, ] to [, 2]. More precisely, since
the regularity conditions imply that can be expanded into a series of cos(k), the analytical
continuation satises
[, 2], (r, ) = (r, 2 ). (4.33)
Now, when r +, the asymptotic behavior of is O(1/r) [cf. Eq. (3.26)] and the same things
holds for ln A [see (3.63) for the spherically symmetric case]. To have a regular asymptotically
at solution, we must therefore have I = 0 in (4.30). In view of (4.31), (4.33) and expression
(3.38) for S , this implies
Z r=+ Z =  
2 2 3B 2 r 2 sin2
8A [p + (E + p)U ] + r dr d = 0 . (4.34)
r=0 =0 4N 2

This is the GRV2 identity. Note that the integration domain in has been reduced to [0, ]
thanks to the property (4.33), thanks to which I is twice the integral in the left-hand side of
(4.34). As GRV3, it is very useful to check the accuracy of numerical solutions [76, 87].
The Newtonian limit of GRV2 is
Z r=+ Z =  
2 1 ~ 2
p + U () r dr d = 0 . (4.35)
r=0 =0 8G
Newt.
4.4 Angular momentum 63

The identity (4.35) has been exhibited in 1939 by Chandrasekhar [32] in the special case of
spherical symmetry, but it seems that it has not been considered for axisymmetric Newtonian
bodies before the work [17].

Remark : Although it also involves a 2-dimensional Laplacian, Eq. (3.16) of the Einstein system does
not give birth to a virial identity because of the sin factor on its right-hand side, which leads to
(r, ) = (r, 2 ) instead of (4.33), implying that I is automatically zero.

4.4 Angular momentum


The Komar angular momentum (or angular momentum for brevety) is dened by a
integral similar to that used to dene the Komar mass in 4.2.1 but replacing the stationarity
Killing vector ~ by the axisymmetry Killing vector
~:
I
1
J := dS , (4.36)
16 S

where S is any closed 2-surface (sphere) surrounding the star and dS is the area-element
2-form normal to S given by (4.8). As for the Komar mass, J is independent of the choice of
S provided the star is entirely enclosed in S .
In terms of the 3+1 quantities J is expressible as (cf. [42] for details)
I
1
J= Kij si j q d d. (4.37)
8 S

By means of the Gauss-Ostrogradski theorem and Einstein equation, we can get a volume ex-
pression of J similar to (4.12), diering only by the replacement of ~ by
~ and the overall factor
2 by 1. Since ~ is tangent to t , we have T (~ ~ ) = ~
n, p~ and n~ ~ = 0, so that the nal
expression reduces to Z

J= pi i dr d d. (4.38)
t

For a perfect uid rotating star, pi i = p and we get, by combining Eqs. (3.37) and (3.3),
Z
J= (E + p)U A2 B 2 r 3 sin2 dr d d . (4.39)
t

At the Newtonian limit, this expression reduces to


Z
J= U r sin r 2 sin dr d d . (4.40)
t
Newt.

We recognize the z-component of the star total angular momentum.


Let us make explicit the surface integral expression of J, Eq. (4.37), in the present case. We
have from Eq. (2.43)-(2.44) and (4.9),

B 2 r 2 sin2
Kij si j = Kr sr = r 2 sin when r +.
2AN r r
64 Global properties of rotating stars

Therefore taking the limit S in Eq. (4.37) gives


I
1 4 3
J = lim r sin d d.
16 S S r
Now, from the asymptotic behavior of as given by (3.27), /r = 3K1 /r 4 + O(1/r 5 ). Hence
Z Z
3K1 3 3K1 K1
J= sin d = (1 cos2 ) sin d = .
8 0 8 0 2
We may therefore rewrite the asymptotic behavior of as
 
2J 1
= 3 +O . (4.41)
r r4

In particular, we recover the fact that if the star is not rotating (J = 0), then = 0 [cf. (2.42)].
Having dened both J and M , we can form the dimensionless ratio

J
j := . (4.42)
M2
j is called the Kerr parameter. For a Kerr black hole of parameters (a, M ), j = a/M and
varies between 0 (Schwarzschild black hole) and 1 (extreme Kerr black hole).
Besides, for a rigidly rotating star, one denes the moment of inertia by
J
I := . (4.43)

Remark : Due to the so-called supertranslation ambiguity, there is no equivalent of the ADM mass for
the angular momentum (see e.g. [59] for a discussion).

4.5 Circumferential radius


The stellar equator is the closed line at the surface of the star [p = 0, cf. (3.103)] dened by
t = const and = /2 (equatorial plane). It has a constant value of the coordinate r, that we
have already denoted by req . A coordinate-independent characterization of the stellar equator is
the circumferential radius:
C
Rcirc := , (4.44)
2
where C is the circumference of the star in the equatorial plane, i.e. the length of the equator
as given by the metric tensor. According to the line element (2.38), we have
I I Z
B(req , /2) req 2
q
1 1
Rcirc = ds = B 2 r 2 sin2 d2 = d,
2 r=req 2 r=req 2 0
=/2 =/2

hence
Rcirc = B(req , /2) req . (4.45)
Of course, at the Newtonian limit, B = 1 and Rcirc = req . For a relativistic star, we have
B > 1 (cf. Fig. 3.4, taking into account that B is very close to A, as shown in Fig. 3.5), so that
Rcirc > req .
4.6 Redshifts 65

4.6 Redshifts
4.6.1 General expression of the redshift
Consider a photon emitted by the uid at the surface of the star. The photon follows a null
~, which is a future directed null
geodesic L of spacetime and is described by its 4-momentum2 p
vector:
~p
p ~ = 0. (4.46)

~ is tangent to L in each point and obeys the geodesic equation


p

~ = 0.
p~ p (4.47)

For an observer who is receiving the photon, the relative shift in wavelength, called redshift,
is
rec em
z := , (4.48)
em

where rec is the photons wavelength as measured by the observer (the receiver) and em is the
photons wavelength in the uid frame (the emitter). In terms of the photon energy E = ~c/,
the above formula can be written
Eem
z= 1. (4.49)
Erec
Now, by the very denition of the photon 4-momentum, the photon energy measured by an
observer O is
EO = ~uO p
~, (4.50)

where u
~ O is the 4-velocity of O. In the present case, u ~ (the uid 4-velocity), whereas, for
~ em = u
an observer at innity and at rest with respect to the star u ~ Hence
~ rec = .

~ p
u ~|em
z= 1 , (4.51)
~ p

~
rec

where the indices em and rec indicate the location where the scalar products u ~ p~ and ~ p
~
are to be taken.
A fundamental property for the scalar product ~ p ~ is to be conserved along the photon
~ we have successively
geodesic worldline. To show it, let us take the scalar product of (4.47) by ;

p p = 0,
p [ ( p ) p ] = 0,
p ( p ) p p = 0,
| {z }
0

p ( p ) = 0, (4.52)
2
We are using the same letter p as for the matter momentum density with respect to the ZAMO inside the
star [cf. Eq. (1.32)], but no confusion should arise.
66 Global properties of rotating stars

where we have used the Killing equation (2.4) to set to zero the term p p :
1 1
p p = (p p + p p ) = p p ( + ) = 0.
2 2 | {z }
0

Equation (4.52) can be rewritten as

p~ (~ p
~) = 0 . (4.53)

Since p
~ is tangent to the photon worldline, this relation shows that

~ p
~ = const along the photon worldline . (4.54)

Hence the symmetry encoded by ~ (stationarity) gives birth to a conserved quantity. This is
another manifestation of the Noether theorem (cf. 3.4.2). Similarly, we have

~ = const along the photon worldline .


~ p
(4.55)
The conservation laws (4.54) and (4.55) are particularly useful when we specialize u ~ in (4.51)
to the circular form (2.25): u ~
~ = (/N )( + t
~ ) (we have used (3.31) to replace u by /N ). We
get then
~ 
u~ p
~= p~ + ~ p ~ ,
N
so that Eq. (4.51) becomes

~ ~
p

~ p~|em
z= em + 1. (4.56)
N em ~ p

~

~ p

~
rec rec

Thanks to (4.54), ~ p = ~ p

~ ~ , hence
em rec
 
~ p
~
z= 1 + 1 . (4.57)
N em ~ p
~

Note that we have suppressed the indices em or rec on ~ and ~ p


~ p ~ thanks to (4.54) and
(4.55).
Remark : For differentially rotating stars, is not a constant and the value to be used in Eq. (4.57) is
the value at the emission point.

4.6.2 Polar redshift


Let us apply formula (4.57) to an emission taking place at one of the poles of the star, i.e. at
one of the two points of intersection of the stellar surface with the rotation axis: = 0 or = .
On the rotation axis ~ = 0 [Eq. (2.6)]. Moreover, = 1 since U = 0 on the rotation axis [cf.
Eq. (3.32) with sin = 0]. Accordingly, formula (4.57) simplies to
1
zp = 1 , (4.58)
Np
where the index p stands for polar values.
4.6 Redshifts 67

4.6.3 Equatorial redshifts


Let us now consider the emission from the equator and in a direction tangent to it, i.e. an
emission from the left or right edge of the star as seen by the distant observer. The photon
4-momentum at the emission is then
~ = pt ~ + p
p ~, (4.59)
with p > 0 (resp. p < 0) for an emission in the forward (resp. backward) direction with
respect to the uid motion. The term in factor of in formula (4.57) is then
~ p
~ ~ ~ +
~ ~ gt + g p
= = , with := . (4.60)
~
p~ ~ ~ ~
+ ~ gtt + gt pt
Note that since p~ is a future directed null vector, we have pt > 0; in particular pt 6= 0, so that
is always well dened.
Besides, Eq. (4.46) writes gtt (pt )2 + 2gt pt p + g (p )2 = 0, i.e.
g 2 + 2gt + gtt = 0. (4.61)
From this relation, we can write (4.60) as
~ p
~ 1
= . (4.62)
~ p
~
Finally, we compute by solving the second-order equation (4.61). We get two solutions:
q
gt gt 2 g g
tt Br N
= = ,
g Br
where the second equality follows from the expression (2.34) of the metric components, specialized
at = /2 (equatorial plane). Substituting the above value for into Eq. (4.62) leads to
~ p
~ Br
= . (4.63)
~ p
~ N Br

Since ~ and p~ are both future directed, we have always ~ p


~ < 0. For a emission in the forward
direction 2 2 t
~ = B r (p p ) 0, so the sign must selected in the above expression. On
~ p
the contrary, the + sign corresponds to an emission in the backward direction. Inserting this
result into Eq. (4.57), we thus get, for an emission in the forward direction
   
f Br Br 1 1U
zeq = 1 1= 1 ( ) 1 = 1,
N N + Br N + Br |N {z } N + Br 1 U 2
U
i.e. r
f 1 1U
zeq = 1 . (4.64)
N + Br 1+U
In the right-hand side, all the functions are to be taken at the emission point. Similarly choosing
the + sign in (4.63) leads to the redshift in the backward direction:
r
b 1 1+U
zeq = 1 . (4.65)
N Br 1 U
68 Global properties of rotating stars

Remark : As a check of the above expressions, we may consider the flat spacetime limit: N = 1 and
= 0. We then recover the standard Doppler formulas
r r
f 1U b 1+U
zeq = 1 and zeq = 1.
1+U 1U
f
They imply zeq 0, i.e. a blueshift (the emitter is moving towards the observer in the forward case)
b
and zeq 0, i.e. a true redshift (the emitter is moving away from the observer in the backward
case).

4.7 Orbits around the star


4.7.1 General properties of orbits
Let us consider a test particle P of mass m M orbiting the star. Let v
~ be the 4-velocity of
P; its components with respect to QI coordinates are
 
dt dr d d
v = , , , , (4.66)
d d d d
where is the proper time of P. The 4-momentum of the particle is

~ = m~
p v. (4.67)

The 4-velocity normalization relation v ~ = 1 is equivalent to


~ v

~ = m2 .
~p
p (4.68)

Assuming that P is submitted only to the gravitation of the star, its worldline is a timelike
geodesic of (M , g). In particular v~ ~v = 0 or equivalently

~ = 0.
p~ p (4.69)

Due to the symmetries of (M , g), we have the same conserved quantities as those given by
Eqs. (4.54)-(4.55) in the photon case:

E := ~ p
~ = pt = const (4.70)
~ = p = const .
~ p
L := (4.71)

Indeed the demonstration of (4.54)-(4.55) [cf. Eq. (4.52)] used only (i) the Killing character of ~
and ~ and (ii) the property (4.69), which holds for a photon as well as for a massive particle. If
the particle reaches spacelike innity, where ~ n
~ (the ZAMO 4-velocity), the constant E is the
particle energy as measured by the asymptotic ZAMO Similarly, L is the angular momentum of
the particle with respect to the rotation axis as measured by the same observer: when r +,
L mr 2 sin2 d/dt (assuming that asymptotically, the proper time of the P coincides with t,
i.e. that P is not relativistic).
Everywhere is space, the energy of P as measured by the ZAMO is [cf. Eq. (2.41)]
1 ~  1  ~ 
EZAMO = ~ np~= + ~ p ~= p
~ ~ p~ ,
N N
4.7 Orbits around the star 69

hence
1
EZAMO = (E L) . (4.72)
N
Remark : Since N 1 and 0 when r +, it is clear on this formula that E = EZAMO if the
particle reaches spatial infinity.

4.7.2 Orbits in the equatorial plane


We restrict ourselves to orbits in the equatorial plane: = /2. Then, according to Eq. (4.66)
and (4.67), p = md/d = 0. Given the form (2.32) of the metric coecients, this yields

p = g p = A2 r 2 p = 0.

Combining with (4.70)-(4.71), we conclude that we have three constants among the four compo-
nents p : pt , p and p . The fourth component, pr , is obtained from the normalization relation
(4.68):
g p p = m2 .
Substituting the values of g given by (2.35), we get
 
p2t pt p p2r p2 1 2
2 2 + 2+ 2 2+ p2 = m2 .
N N2 A r A B 2 r 2 sin2 N 2

Since = /2, pt = E, p = 0 and p = L, we get :

p2r 1 2 2 L2
= (E L) m . (4.73)
A2 N2 B 2r2
Now, from (2.34), (4.66) and (4.67),

dr
pr = gr p = A2 pr = A2 mv r = A2 m .
d
Thus, we can rewrite (4.73) as
 2
1 dr L)
=0,
+ V(r, E, (4.74)
2 d

and L
where E are the constants of motion dened by

:= E
E and := L ,
L (4.75)
m m
and 
1 1 2 2 
L

V(r, E, L) :=
1 2 E L + 2 2 . (4.76)
2A2 N B r

In view of Eq. (4.74), we conclude that the geodesic motion of a test particle in the equatorial
plane is equivalent to the Newtonian motion of a particle in a one-dimensional space (spanned
L).
by the coordinate r) under the eective potential V(r, E,
70 Global properties of rotating stars

4.7.3 Circular orbits


Circular orbits are dened by r = const, i.e. by
dr
=0 (4.77)
d
d2 r
= 0. (4.78)
d 2
Remark : Condition (4.77) is not sufficient to ensure a circular orbit, for it is satisfied at the periastron
or apoastron of non-circular orbits.

From Eq. (4.74), the two conditions (4.77)-(4.78) are equivalent to V = 0 and V/r = 0, i.e. to

1 2
2+ L =0

1 E L (4.79)
N2  B 2r2 

E L 2 
L 1


(E L) +L 2 2 + = 0, (4.80)
N2 r r B r r r

where
:= ln N and := ln B. (4.81)
As we did for the uid, let us introduce the particle angular velocity as seen by a distant
observer,
d v
P := = t, (4.82)
dt v
~ measured by the ZAMO. Then, similarly to Eqs. (2.25), (3.30)
as well as the particle velocity V
and (3.30) we have  
~ = v t ~ + P
v ~ , (4.83)
 
~ = P n
v ~ +V~ , (4.84)

P = N v t and ~ = 1 (P )~
V . (4.85)
N
~ is |V | with [cf. Eq. (3.32) with sin = 1]
As for (3.33), the norm of V

B
V := (P )r. (4.86)
N
The normalization relation v v = 1 is then equivalent to
~ ~
1/2
P = 1 V 2 . (4.87)

From Eq. (4.70) and (4.83), we have


 
= vt = gt v = (N 2 B 2 2 r 2 )v t + B 2 r 2 v t P = v t N 2 + B 2 r 2 (P ) .
E

Thanks to (4.85) and (4.86), we can write

= P (N + BrV ) .
E (4.88)
4.7 Orbits around the star 71

Besides, from Eq. (4.71) and (4.83), we have


= v = g v = B 2 r 2 v t + B 2 r 2 v t P = v t B 2 r 2 (P ),
L

i.e. thanks to (4.85) and (4.86),


= BrP V .
L (4.89)
Combining (4.88) and (4.89), we have

L
E = P N . (4.90)

This result leads to the following expression for the energy of P as measured by the ZAMO and
given by Eq. (4.72):
EZAMO = P m . (4.91)
The thus recover Einsteins mass-energy relation. This is not surprising since P is the Lorentz
factor of P with respect to the ZAMO.
In view of the above relations, let us go back to the system (4.79)-(4.80). Thanks to (4.90)
and (4.89), Eq. (4.79) is equivalent to 1 2P + 2P V 2 = 0. It is trivially satised since P and
V are related by (4.87). Inserting (4.90) and (4.89) into Eq. (4.80) gives rise to a second order
equation for V :  
1 Br
+ V2 V = 0. (4.92)
r r N r r
This equation has been rst written by Bardeen (1970) [9]. It admits two solutions:
r  
2
Br B2 r2

N r N2 r + 4
r r + 1
r
V =   . (4.93)
1
2 r + r

This formula gives the circular orbit velocity with respect to the ZAMO as a function of r. All
the metric coecients in the right-hand side are to be taken at = /2. The + (resp. ) sign
corresponds to a motion in the same sens (resp. opposite sens) than the stellar rotation: the
orbit is then called direct (resp. retrograde).
At the Newtonian limit, 0, |/r| 1/r and /c2 , so that (4.93) reduces to
q
c24r
r
V = c 2
r

i.e. r

V = r . (4.94)
r Newt.
We thus recover the standard formula obtained by equating the gravitational acceleration,
r ,
2
and the centrifugal one, V /r. In particular, in spherical symmetry, = GM/r and (4.94)
p
yields to the well known result : V = GM/r.

Remark : At the Newtonian limit, V = V+ , whereas in the relativistic case, if the star is rotating
( 6= 0), V 6= V+ .
72 Global properties of rotating stars

4.7.4 Innermost stable circular orbit (ISCO)


We have seen from Eq. (4.74) that a circular orbit corresponds to an extremum of the eective
potential: V/r = 0. It is stable if, and only if, this extremum is actually a minimum, i.e. if,
and only if,
2V
> 0. (4.95)
r 2
Taking into account that V = 0 and V/r = 0 on a circular orbit, the above criteria is equivalent
to
( "  2 #  2 )
E L 2
(E L)
2 +L 4
N2 r 2 r r 2 r r
"   #
2
L 2 4 2 3 L

2 2
+ 2 2 2 + +2 + 2 2 > 0. (4.96)
B r r r r r r N r

Substituting (4.90) for E L and (4.89) for L,


we get, after division by 2 ,
P
   
2 2 V Br 2
2
2 + 2
4
r r N r r r
"  2 #  
2
2
4 3 V 2 B 2 r 2 2
+V 2 + +2 + 2 > 0. (4.97)
r r r r r N2 r

The marginally stable orbit, if any, corresponds to the vanishing of the left-hand side in the above
inequality. This orbit is called the innermost stable circular orbit (ISCO). To compute its
location, one has to solve the equation formed by replacing the > sign in (4.97) by an equal sign:
 2  
2 V Br 2
2 + 4
r 2 r N r 2 r r
"   #  
2 2 4 2 3 V 2 B 2 r 2 2
+V 2 + +2 + 2 = 0. (4.98)
r r r r r N2 r

In this equation, V is considered as the function of the metric potentials N , B and given by
(4.93) and all the potentials are to be taken at = /2. So, once the metric is known, Eq. (4.98)
is an equation in r. If it admits a solution r > req (i.e. outside the star), then the ISCO exists.
Otherwise, all the circular orbits are stable down to the stellar surface.
At the Newtonian limit ( /c2 , || 1 and 0), Eq. (4.98) reduces to
2 3
+ = 0, (4.99)
r 2 r r
where we have used (4.94) to express V 2 .
There are basically two causes for the existence of the ISCO:
The star is highly relativistic: for instance in spherical symmetry the metric outside the
star is Schwarzschild metric (cf. 3.3.3); if the star is compact enough, its surface is located
under the famous r = 6M which denes the ISCO in Schwarzschild metric. Note that via
the relation (3.59) betweenthe areal coordinate r and the isotropic coordinate r, r = 6M
corresponds to r = (5/2 + 6)M .
4.7 Orbits around the star 73

The star deviates significantly from spherical symmetry: in Newtonian gravity, all the orbits
are stable around a spherically symmetric star. But if the gravitational potential deviates
signicantly from = GM/r, an ISCO may exist, even in Newtonian gravity. This was
shown explicitly for very rapidly rotating low-mass strange quark stars in Ref. [103]. For
instance, assuming that the deviation from spherical symmetry arises only from the mass
quadrupole moment of the star Q (this is the leading non-spherical term when r +),
we have3
GM GQ
= + 3 P2 (cos ), (4.100)
r r
where P2 (x) = (3x2 1)/2 is the Legendre polynomial of degree 2. Inserting the above
expression of in Eq. (4.99), setting = /2 and solving for r leads to
r
3Q
rISCO = . (4.101)
2M

3
Equation (4.100) is nothing but Eq. (3.25) truncated at = 2 and with equatorial symmetry, so that the term
= 1 vanishes.
74 Global properties of rotating stars
Appendix A

Lie derivative

Contents
A.1 Lie derivative of a vector field . . . . . . . . . . . . . . . . . . . . . . . 75
A.2 Generalization to any tensor field . . . . . . . . . . . . . . . . . . . . . 77

A.1 Lie derivative of a vector field


A.1.1 Introduction
We have seen in 1.1.5 that the denition of the derivative of a vector eld v ~ on M requires
some extra structure on the manifold M . This can be some connection (such as the Levi-
Civita connection associated with the metric tensor g), leading to the covariant derivative ~ v.
Another extra structure can be some reference vector eld u ~ , leading to the concept of derivative
of v ~ ; this is the Lie derivative presented in this Appendix. These two types of derivative
~ along u
generalize straightforwardly to any kind of tensor eld1 .

A.1.2 Definition
Consider a vector eld u ~ on M , called hereafter the flow. Let v ~ be another vector eld on M ,
the variation of which is to be studied. We can use the ow u ~ to transport the vector ~v from one
point p to a neighbouring one q and then dene rigorously the variation of ~v as the dierence
between the actual value of v~ at q and the transported value via u ~ . More precisely the denition
of the Lie derivative of v
~ with respect to u
~ is as follows (see Fig. A.1). We rst dene the image
(p) of the point p by the transport by an innitesimal distance along the eld lines of u ~


as (p) = q, where q is the point close to p such that pq = ~ u(p). Besides, if we multiply the
vector v~ (p) by some innitesimal parameter , it becomes an innitesimal vector at p. Then
there exists a unique point p close to p such that ~ v (p) =
. We may transport the point p
pp
1
For the specific kind of tensor fields constituted by antisymmetric multilinear forms, there exists a third type
of derivative, which does not require any extra structure on M : the exterior derivative (see the classical textbooks
[75, 96, 90] or Ref. [41] for an introduction).
76 Lie derivative

Figure A.1: Geometrical construction of the Lie derivative of a vector eld: given a small pa-
rameter , each extremity of the arrow ~ v is dragged by some small parameter along u ~ , to
form the vector denoted by (~v ). The latter is then compared with the actual value of ~
v at
the point q, the dierence (divided by ) dening the Lie derivative Lu~ ~v .

to a point q along the eld lines of u ~ by the same distance as that used to transport p to q:
q = (p ) (see Fig. A.1).

qq is then an innitesimal vector at q and we dene the transport by
the distance of the vector v ~ (p) along the eld lines of u
~ according to
1
(~
v (p)) := qq . (A.1)

v (p)) is vector in Tq (M ). We may then subtract it from the actual value of the eld v
(~ ~ at q
and dene the Lie derivative of v ~ along u
~ by
1
Lu~ ~v := lim v (q) (~
[~ v (p))] . (A.2)
0

If we consider a coordinate system (x ) adapted to the eld u ~ in the sense that u ~0 where
~ =
~
0 is the rst vector of the natural basis associated with the coordinates (x ) (cf. 1.1.1), then
the Lie derivative is simply given by the partial derivative of the vector components with respect
to x0 :
v
~ ) =
(Lu~ v . (A.3)
x0
In an arbitrary coordinate system, this formula is generalized to

v u

Lu~ v = u v , (A.4)
x x

where use has been made of the standard notation Lu~ v := (Lu~ v ~ ) . The above relation shows
that the Lie derivative of a vector with respect to another one is nothing but the commutator of
these two vectors:
Lu~ ~v = [~ ~] .
u, v (A.5)
A.2 Generalization to any tensor field 77

A.2 Generalization to any tensor field


The Lie derivative is extended to any tensor eld by (i) demanding that for a scalar eld f ,
u) and
Lu~ f = df (~  (ii) using the Leibniz rule. As a result, the Lie derivative Lu~ T of a tensor
eld T of type k is a tensor eld of the same type, the components of which with respect to
a given coordinate system (x ) are

k i
1 ...k 1 ...k


X 1 ... ...k ui X 1 ...k u
Lu~ T 1 ... =u T 1 ... T 1 ... x + T 1 ... ... xi .
x
i=1 i=1 i
(A.6)
In particular, for a 1-form,
u
Lu~ = u + (A.7)
x x
and, for a bilinear form,
T u u
Lu~ T = u + T + T . (A.8)
x x x
Notice that, thanks to the symmetry properties of the Christoel symbols dened by (1.19), the
partial derivatives the above equations can be replaced by the covariant derivatives associated
with the metric g, yielding

k i
X X
1 ...k 1 ...k 1 ... ...k
Lu~ T 1 ...

= u T 1 ... T 1 ... ui + T 1 ...k1 ... ... i u .

i=1 i=1 i
(A.9)

Lu~ = u + u , (A.10)
Lu~ T = u T + T u + T u . (A.11)
78 Lie derivative
Appendix B

Lorene/nrotstar code

The nrotstar code is a free software based on the C++ Lorene library [104]. It comes along
when downloading Lorene, in the directory Lorene/Codes/Nrotstar. It is a descendant of
rotstar, a previous Lorene code described in [45] and used in many astrophysical studies (see
e.g. [51], [100] and references therein).
nrotstar solves the Einstein equations (3.14)-(3.17) via the self-consistent-eld method ex-
posed in 3.5.1. For this purpose it uses a multi-domain spectral method developed in [18] and
[47] (see [48] for a review about spectral methods). At least three domains are employed (cf.
Fig. B.1):

a nucleus, covering the interior of the star;

a intermediate shell, covering the strong eld region outside the star;

an external compactified domain, extending to r = +.

But more intermediate shells can be used, either inside the star or outside it. For instance, the
nucleus can be dedicated to the uid interior and a the rst shell to the solid crust [101], or two
domains inside the star can be used to describe a phase transition [102].

Not finished yet ! (to be continued...)


80 Lorene/nrotstar code

Figure B.1: Domains used in the nrotstar code.


Bibliography

[1] A. Akmal, V.R. Pandharipande & D.G. Ravenhall : Equation of state of nucleon matter
and neutron star structure, Phys. Rev. C 58, 1804 (1998).
[2] M. Alcubierre : Introduction to 3+1 Numerical Relativity, Oxford University Press, Oxford
(2008).
[3] N. Andersson & G. L. Comer : Relativistic Fluid Dynamics: Physics for Many Different
Scales, Living Rev. Relativity 10, 1 (2007);
http://www.livingreviews.org/lrr-2007-1
[4] M. Ansorg, T. Fischer, A. Kleinwchter, R. Meinel, D. Petro & K. Schbel: Equilibrium
configurations of homogeneous fluids in general relativity, Mon. Not. Roy. Astron. Soc. 355,
682 (2004).
[5] M. Ansorg, D. Gondek-Rosiska & L. Villain : On the solution space of differentially rotating
neutron stars in general relativity, Mon. Not. Roy. Astron. Soc. 396, 2359 (2009).
[6] M. Ansorg, A. Kleinwaechter & R. Meinel : Highly accurate calculation of rotating neutron
stars, Astron. Astrophys. 381, L49 (2002).
[7] R. Arnowitt, S. Deser and C.W Misner : The Dynamics of General Relativity, in Gravitation:
an introduction to current research, edited by L. Witten, Wiley, New York (1962), p. 227;
http://arxiv.org/abs/gr-qc/0405109.
[8] A. Ashtekar and A. Magnon-Ashtekar : On conserved quantities in general relativity, J.
Math. Phys. 20, 793 (1979).
[9] J. M. Bardeen : Stability of Circular Orbits in Stationary, Axisymmetric Space-Times,
Astrophys. J. 161, 103 (1970);
http://adsabs.harvard.edu/abs/1970ApJ...161..103B
[10] J. M. Bardeen : A Variational Principle for Rotating Stars in General Relativity, Astrophys.
J. 162, 71 (1970);
http://adsabs.harvard.edu/abs/1970ApJ...162...71B
[11] J. M. Bardeen & T. Piran : General relativistic axisymmetric rotating systems: coordinates
and equations, Phys. Rep. 96, 205 (1983).
[12] J. M. Bardeen & R. V. Wagoner : Relativistic Disks. I. Uniform Rotation, Astrophys. J.
167, 359 (1971);
http://adsabs.harvard.edu/abs/1971ApJ...167..359B
82 BIBLIOGRAPHY

[13] T. W. Baumgarte & S. L. Shapiro : Numerical Relativity. Solving Einsteins Equations on


the Computer, Cambridge Univ. Press, Cambridge, in press.

[14] R. Beig : Arnowitt-Deser-Misner energy and g00 , Phys. Lett. 69A, 153 (1978).

[15] J. Bik & T. Ledvinka : Relativistic disks as sources of the Kerr metric, Phys. Rev. Lett.
71, 1669 (1993).

[16] S. Bonazzola : The virial theorem in general relativity, Astrophys. J. 182, 335 (1973);
http://adsabs.harvard.edu/abs/1973ApJ...182..335B

[17] S. Bonazzola & E. Gourgoulhon : A virial identity applied to relativistic stellar models,
Class. Quantum Grav. 11, 1775 (1994).

[18] S. Bonazzola, E. Gourgoulhon & J.-A. Marck : Numerical approach for high precision 3-D
relativistic star models, Phys. Rev. D 58, 104020 (1998).

[19] S. Bonazzola, E. Gourgoulhon, M. Salgado & J.-A. Marck : Axisymmetric rotating rela-
tivistic bodies: a new numerical approach for exact solutions, Astron. Astrophys. 278, 421
(1993);
http://adsabs.harvard.edu/abs/1993A%26A...278..421B

[20] S. Bonazzola & G. Maschio : Models of Rotating Neutron Stars in General Relativity, in
The Crab Nebula, Proceedings from IAU Symposium no. 46, edited by R. D. Davies and F.
Graham-Smith, Reidel, Dordrecht (1971), p. 346;
http://adsabs.harvard.edu/abs/1971IAUS...46..346B

[21] S. Bonazzola & J. Schneider : An Exact Study of Rigidly and Rapidly Rotating Stars in
General Relativity with Application to the Crab Pulsar, Astrophys. J. 191, 273 (1974);
http://adsabs.harvard.edu/abs/1974ApJ...191..273B

[22] R. H. Boyer : Rotating fluid masses in general relativity, Mathematical Proceedings of the
Cambridge Philosophical Society 61, 527 (1965).

[23] R. H. Boyer : Rotating fluid masses in general relativity. II, Mathematical Proceedings of
the Cambridge Philosophical Society 62, 495 (1966).

[24] R. H. Boyer & R. W. Lindquist : A variational principle for a rotating relativistic fluid,
Phys. Lett. 20, 504 (1966).

[25] E. M. Butterworth : On the structure and stability of rapidly rotating fluid bodies in general
relativity. II - The structure of uniformly rotating pseudopolytropes, Astrophys J. 204, 561
(1976);
http://adsabs.harvard.edu/abs/1976ApJ...204..561B

[26] E. M. Butterworth & J. R. Ipser : Rapidly rotating fluid bodies in general relativity, Astro-
phys J. 200, L103 (1975);
http://adsabs.harvard.edu/abs/1975ApJ...200L.103B
BIBLIOGRAPHY 83

[27] E. M. Butterworth & J. R. Ipser : On the structure and stability of rapidly rotating fluid
bodies in general relativity. I. The numerical method for computing structure and its appli-
cation to uniformly rotating homogeneous bodies, Astrophys J. 204, 200 (1976);
http://adsabs.harvard.edu/abs/1976ApJ...204..200B

[28] B. Carter : Killing Horizons and Orthogonally Transitive Groups in Space-Time, J. Math.
Phys. 10, 70 (1969).

[29] B. Carter : The commutation property of a stationary, axisymmetric system, Commun.


Math. Phys. 17, 233 (1970).

[30] B. Carter : Black hole equilibrium states, in Black holes Les Houches 1972, Eds. C. DeWitt
& B.S. DeWitt, Gordon and Breach, New York (1973), p. 57.

[31] B. Carter : Perfect fluid and magnetic field conservations laws in the theory of black hole ac-
cretion rings, in Active Galactic Nuclei, Eds. C. Hazard & S. Mitton, Cambridge University
Press, Cambridge (1979), p. 273.

[32] S. Chandrasekhar : An introduction to the study of stellar structure, University of Chicago


Press, Chicago (1939).

[33] Y. Choquet-Bruhat and J.W. York : The Cauchy Problem, in General Relativity and Grav-
itation, one hundred Years after the Birth of Albert Einstein, Vol. 1, edited by A. Held,
Plenum Press, New York (1980), p. 99.

[34] G.B. Cook, S.L. Shapiro & S.A. Teukolsky : Spin-up of a rapidly rotating star by angular
momentum loss: Effects of general relativity, Astrophys. J. 398, 203 (1992); errata in [35];
http://adsabs.harvard.edu/abs/1992ApJ...398..203C

[35] G.B. Cook, S.L. Shapiro & S.A. Teukolsky : Rapidly rotating polytropes in general relativity,
Astrophys. J. 422, 227 (1994);
http://adsabs.harvard.edu/abs/1994ApJ...422..227C

[36] G.B. Cook, S.L. Shapiro & S.A. Teukolsky : Recycling pulsars to millisecond periods in
general relativity, Astrophys. J. 423, L117 (1994);
http://adsabs.harvard.edu/abs/1994ApJ...423L.117C

[37] G.B. Cook, S.L. Shapiro & S.A. Teukolsky : Rapidly rotating neutron stars in general
relativity: Realistic equations of state, Astrophys. J. 424, 823 (1994);
http://adsabs.harvard.edu/abs/1994ApJ...424..823C

[38] F. Douchin & P. Haensel : A unified equation of state of dense matter and neutron star
structure, Astron. Astrophys. 380, 151 (2001);
http://adsabs.harvard.edu/abs/2001A%26A...380..151D

[39] J.L. Friedman, J.R. Ipser & L. Parker: Rapidly rotating neutron star models, Astrophys. J.
304, 115 (1986);
http://adsabs.harvard.edu/abs/1986ApJ...304..115F

[40] J.L. Friedman, J.R. Ipser & L. Parker: Implications of a half-millisecond pulsar, Phys. Rev.
Lett. 62, 3015 (1989).
84 BIBLIOGRAPHY

[41] E. Gourgoulhon : An introduction to relativistic hydrodynamics, in Stellar Fluid Dynam-


ics and Numerical Simulations: From the Sun to Neutron Stars, Eds. M. Rieutord & B.
Dubrulle, EAS Publications Series 21, 43 (2006);
http://arxiv.org/abs/gr-qc/0603009

[42] E. Gourgoulhon : 3+1 Formalism and Bases of Numerical Relativity, Lectures given at the
General Relativity Trimester held in the Institut Henri Poincare (Paris, Sept.-Dec. 2006);
http://arxiv.org/abs/gr-qc/0703035

[43] E. Gourgoulhon & S. Bonazzola : Noncircular axisymmetric stationary spacetimes, Phys.


Rev. D 48, 2635 (1993).

[44] E. Gourgoulhon & S. Bonazzola : A formulation of the virial theorem in general relativity,
Class. Quantum Grav. 11, 443 (1994).

[45] E. Gourgoulhon, P. Haensel, R. Livine, E. Paluch, S. Bonazzola, & J.-A. Marck : Fast
rotation of strange stars, Astron. Astrophys. 349, 851 (1999);
http://adsabs.harvard.edu/abs/1999A%26A...349..851G

[46] J. O. Goussard, P. Haensel & J. L. Zdunik : Rapid uniform rotation of protoneutron stars,
Astron. Astrophys. 321, 822 (1997);
http://adsabs.harvard.edu/abs/1997A%26A...321..822G

[47] P. Grandclment, S. Bonazzola, E. Gourgoulhon & J.-A. Marck : A multi-domain spectral


method for scalar and vectorial Poisson equations with non-compact sources, J. Comput.
Phys. 170, 231-260 (2001).

[48] P. Grandclment & J. Novak : Spectral Methods for Numerical Relativity, Living Rev. Relat.
12, 1 (2009);
http://www.livingreviews.org/lrr-2009-1

[49] P. Haensel & B. Pichon : Experimental nuclear masses and the ground state of cold dense
matter, Astron. Astrophys. 283, 313 (1994);
http://adsabs.harvard.edu/abs/1994A%26A...283..313H

[50] P. Haensel, A. Y. Potekhin & D. G. Yakovlev : Neutron stars 1: Equation of state and
structure, Springer, New York (2007).

[51] P. Haensel, J.L. Zdunik, M. Bejger, J.M. Lattimer : Keplerian frequency of uniformly ro-
tating neutron stars and quark stars, Astron. Astrophys. 502, 605 (2009).

[52] J. B. Hartle : Slowly Rotating Relativistic Stars. I. Equations of Structure, Astrophys. J.


150, 1005 (1967);
http://adsabs.harvard.edu/abs/1967ApJ...150.1005H

[53] J. B. Hartle : Gravity: An Introduction to Einsteins General Relativity, Addison Wesley


(Pearson Education), San Fransisco (2003).

[54] J. B. Hartle & D. H. Sharp : Variational Principle for the Hydrostatic Equilibrium of a
Relativistic, Rotating Fluid, Phys. Rev. Lett. 15, 909 (1965).
BIBLIOGRAPHY 85

[55] J. B. Hartle & D. H. Sharp : Variational Principle for the Equilibrium of a Relativistic,
Rotating Star, Astrophys. J. 147, 317 (1967);
http://adsabs.harvard.edu/abs/1967ApJ...147..317H

[56] J. B. Hartle & K.S. Thorne: Slowly Rotating Relativistic Stars. II. Models for Neutron Stars
and Supermassive Stars, Astrophys. J. 153, 807 (1968);
http://adsabs.harvard.edu/abs/1968ApJ...153..807H

[57] U. Heilig : On the existence of rotating stars in general relativity, Commun. Math. Phys.
166, 457 (1995); http://projecteuclid.org/euclid.cmp/1104271700

[58] J. W. T. Hessels, S. M. Ransom, I. H. Stairs, P. C. C. Freire, V. M. Kaspi & F. S. Camilo :


A Radio Pulsar Spinning at 716 Hz, Science 311, 1901 (2006).

[59] J. L. Jaramillo & E. Gourgoulhon : Mass and Angular Momentum in General Relativity, in
Mass and Motion in General Relativity, eds. L. Blanchet, A. Spallicci and B. Whiting, in
press;
http://arxiv.org/abs/1001.5429

[60] A. Komar : Covariant Conservation Laws in General Relativity, Phys. Rev. 113, 934 (1959).

[61] H. Komatsu, Y. Eriguchi & I. Hachisu : Rapidly rotating general relativistic stars I.
Numerical method and its application to uniformly rotating polytropes, Mont. Not. Roy.
Astron. Soc. 237, 355 (1989);
http://adsabs.harvard.edu/abs/1989MNRAS.237..355K

[62] H. Komatsu, Y. Eriguchi & I. Hachisu : Rapidly rotating general relativistic stars II.
Differentially rotating polytropes, Mont. Not. Roy. Astron. Soc. 239, 153 (1989);
http://adsabs.harvard.edu/abs/1989MNRAS.239..153K

[63] W. Kundt & M. Trmper : Orthogonal Decomposition of Axi-symmetric Stationary Space-


times, Zeitschrift fr Physik 192, 419 (1966).

[64] K. Lanczos : ber eine stationre Kosmologie im Sinne der Einsteinschen Gravitationsthe-
orie, Zeitschrift fr Physik 21, 73 (1924).

[65] J. M. Lattimer, M. Prakash, D. Masak & A. Yahil : Rapidly rotating pulsars and the equation
of state, Astrophys. J. 355, 241 (1990);
http://adsabs.harvard.edu/abs/1990ApJ...355..241L

[66] J. Lense & H. Thirring : ber den Einfluss der Eigenrotation der Zentralkrper auf die Be-
wegung der Planeten und Monde nach der Einsteinschen Gravitationstheorie, Physikalische
Zeitschrift 19, 156 (1918).

[67] T. Lewis : Some Special Solutions of the Equations of Axially Symmetric Gravitational
Fields, Proc. Roy. Soc. London Ser. A 136, 176 (1932).

[68] A. Lichnerowicz : Sur un thorme dhydrodynamique relativiste, Comptes Rendus des


Sances de lAcadmie des Sciences 211, 117 (1940);
http://gallica.bnf.fr/ark:/12148/bpt6k3163d/f117.page
86 BIBLIOGRAPHY

[69] A. Lichnerowicz : Sur linvariant intgral de lhydrodynamique relativiste, Annales scien-


tiques de lcole Normale Suprieure 58, 285 (1941);
http://www.numdam.org/item?id=ASENS_1941_3_58__285_0

[70] A. Lichnerowicz : Thories relativistes de la gravitation et de llectromagntisme, Masson,


Paris (1955); reprinted by ditions Jacques Gabay, Paris (2008).

[71] A. Lichnerowicz : Relativistic hydrodynamics and magnetohydrodynamics, Benjamin, New


York (1967).

[72] L. Lindblom : Stationary stars are axisymmetric, Astrophys. J. 208, 873 (1976);
http://adsabs.harvard.edu/abs/1976ApJ...208..873L

[73] M. Mars & J. M. M. Senovilla : On the construction of global models describing rotating
bodies; uniqueness of the exterior gravitational field, Mod. Phys. Lett. A 13, 1509 (1998).

[74] J. Martn, A. Molina & E. Ruiz : Can rigidly rotating polytropes be sources of the Kerr
metric?, Class. Quantum Grav. 25, 105019 (2008).

[75] C.W. Misner, K.S. Thorne, & J.A. Wheeler : Gravitation, Freeman, New York (1973).

[76] T. Nozawa, N. Stergioulas, E. Gourgoulhon & Y. Eriguchi : Construction of highly accurate


models of rotating neutron stars comparison of three different numerical schemes, Astron.
Astrophys. Suppl. Ser. 132, 431 (1998).

[77] J. P. Ostriker & J. W.-K. Mark : Rapidly rotating stars. I. The self-consistent-field method,
Astrophys. J. 151, 1075 (1968);
http://adsabs.harvard.edu/abs/1968ApJ...151.1075O

[78] A. Papapetrou : Champs gravitationnels stationnaires symtrie axiale, Ann. Inst. Henri
Poincar (A) Phys. thor. 4, 83 (1966);
http://www.numdam.org/item?id=AIHPA_1966__4_2_83_0

[79] H. Pster & U. M. Schaudt : Steps towards an existence proof for rotating stars solutions
in general relativity, Ann. Phys. (Leipzig) 11, 507 (2000).

[80] M. Rieutord : Modeling rapidly rotating stars, in Compte-rendus de la Semaine de


lAstrophysique Franaise 2006, edited by D. Barret, F. Casoli, T. Contini, G. Lagache,
A. Lecavelier & L. Pagani,
http://proc.sf2a.asso.fr/2006/2006sf2a.conf..0501R.pdf

[81] M. Rieutord : The dynamics of rotating fluids and binary stars, in Tidal Effects in Stars,
Planets and Disks, edited by M.-J. Goupil & J.-P. Zahn, EAS Publications Series 29, 127
(2008).

[82] D. M. Sedrakian & E. V. Chubaryan : Stationary axially symmetric gravitational fields (in
Russian), Astrozika 4, 239 (1968); English translation in Astrophysics 4, 87 (1968).

[83] U. M. Schaudt : On the Dirichlet Problem for Stationary and Axisymmetric Einstein Equa-
tions, Commun. Math. Phys. 190, 509 (1998).
BIBLIOGRAPHY 87

[84] U. M. Schaudt & H. Pster : The boundary value problem for the stationary and axisym-
metric Einstein equations is generically solvable, Phys. Rev. Lett. 77, 3284 (1996).

[85] J. M. M. Senovilla : Stationary and Axisymmetric Perfect-Fluid Solutions to Einsteins


Equations, in Rotating Objects and Relativistic Physics, Proceedings of the El Escorial Sum-
mer School on Gravitation and General Relativity 1992, edited by F. J. Chinea & L. M.
Gonzales-Romero, Springer-Verlag, Berlin (1993), p. 73.

[86] H. Stephani, D. Kramer, M. MacCallum, C. Hoenselaers & E. Herlt : Exact Solutions of


Einsteins Field Equations (2nd edition), Cambridge University Press, Cambridge (2003).

[87] N. Stergioulas : Rotating Stars in Relativity, Living Rev. Relativity 6, 3 (2003)


http://www.livingreviews.org/lrr-2003-3

[88] N. Stergioulas & J.L. Friedman : Comparing models of rapidly rotating relativistic stars
constructed by two numerical methods, Astrophys. J. 444, 306 (1995);
http://adsabs.harvard.edu/abs/1995ApJ...444..306S

[89] W. J. van Stockum : The gravitational field of a distribution of particles rotating around an
axis of symmetry, Proc. Roy. Soc. Edinburgh 57, 135 (1937).

[90] N. Straumann : General Relavity, with Applications to Astrophysics, Springer-Verlag, Berlin


(2004).

[91] J. L. Synge : Relativistic Hydrodynamics, Proceedings of the London Mathematical Society


43, 376 (1937); reprinted in Gen. Relat. Grav. 34, 2177 (2002).

[92] L. B. Szabados : Quasi-Local Energy-Momentum and Angular Momentum in General Rel-


ativity, Living Rev. Relativity 12, 4 (2009);
http://www.livingreviews.org/lrr-2009-4

[93] J.-L. Tassoul : Stellar rotation, Cambridge University Press, Cambridge (2000).

[94] K. S. Thorne & D. Macdonald : Electrodynamics in curved spacetime: 3+1 formulation,


Mon. Not. R. Astron. Soc. 198, 339 (1982);
http://adsabs.harvard.edu/abs/1982MNRAS.198..339T

[95] L. Villain : Instabilities of rotating compact stars: A brief overview, in Stellar Fluid Dy-
namics and Numerical Simulations: From the Sun to Neutron Stars, Eds. M. Rieutord &
B. Dubrulle, EAS Publications Series 21, 335 (2006);
http://arxiv.org/abs/astro-ph/0602234

[96] R. M. Wald : General Relativity, Univ. Chicago Press, Chicago (1984).

[97] H. Weyl : Zur Gravitationtheorie, Annalen der Physik 54, 117 (1917).

[98] J. R. Wilson : Models of Differentially Rotating Stars, Astrophys. J. 176, 195 (1972);
http://adsabs.harvard.edu/abs/1972ApJ...176..195W

[99] J. R. Wilson : Rapidly Rotating Neutron Stars, Phys. Rev. Lett. 30, 1082 (1973).
88 BIBLIOGRAPHY

[100] J.L. Zdunik, M. Bejger, P. Haensel, & E. Gourgoulhon : Strong first-order phase transition
in a rotating neutron star core and the associated energy release, Astron. Astrophys. 479,
515 (2008).

[101] J.L. Zdunik, P. Haensel & E. Gourgoulhon : The crust of rotating strange quark stars,
Astron. Astrophys. 372, 535 (2001);
http://adsabs.harvard.edu/abs/2001A%26A...372..535Z

[102] J.L. Zdunik, P. Haensel, E. Gourgoulhon & M. Bejger : Hyperon softening of the EOS of
dense matter and the spin evolution of isolated neutron stars, Astron. Astrophys. 416, 1013
(2004);
http://adsabs.harvard.edu/abs/2004A%26A...416.1013Z

[103] J.L. Zdunik & E. Gourgoulhon : Small strange stars and marginally stable orbit in Newto-
nian theory, Phys. Rev. D 63, 087501 (2001).

[104] Lorene library and codes : http://www.lorene.obspm.fr/

[105] RNS code : http://www.gravity.phys.uwm.edu/rns/

Vous aimerez peut-être aussi