The line element is written as
where
is the 3-metric induced on each spacelike slice.
For a spherically symmetric spacetime, the line element can be written as
m
being a radial (Lagrangian) coordinate, indicating the total
rest-mass enclosed inside the circumference
.
The co-moving character of the coordinates leads, for a perfect fluid, to a stress-energy tensor of the form
In these coordinates the local conservation equation for the
baryonic mass, Equation (2), can be easily integrated to yield the metric potential
b
:
The gravitational field equations, Equation (10), and the equations of motion, Equation (1
), reduce to the following quasi-linear system of partial
differential equations (see also [185
]):
with the definitions
and
, satisfying
. Additionally,
represents the total mass interior to radius
m
at time
t
. The final system, Equations (17), is closed with an EOS of the form given by Equation (9
).
Hydrodynamics codes based on the original formulation of May
and White and on later versions (e.g., [293]) have been used in many nonlinear simulations of supernova and
neutron star collapse (see, e.g., [184
,
280
] and references therein), as well as in perturbative
computations of spherically symmetric gravitational collapse
within the framework of the linearized Einstein equations [251,
252]. In Section
4.1.1
below, some of these simulations are discussed in detail. An
interesting analysis of the above formulation in the context of
gravitational collapse is provided by Miller and Sciama [182]. By comparing the Newtonian and relativistic equations, these
authors showed that the net acceleration of the infalling mass
shells is larger in general relativity than in Newtonian gravity.
The Lagrangian character of May and White's formulation, together
with other theoretical considerations concerning the particular
coordinate gauge, has prevented its extension to
multi-dimensional calculations. However, for one-dimensional
problems, the Lagrangian approach adopted by May and White has
considerable advantages with respect to an Eulerian approach with
spatially fixed coordinates, most notably the lack of numerical
diffusion.
the equations of motion in Wilson's formulation [300,
301
] are
with the ``transport velocity'' given by
. We note that in the original formulation [301
] the momentum density equation, Equation (21
), is only solved for the three spatial components
, and
is obtained through the 4-velocity normalization condition
.
A direct inspection of the system shows that the equations are
written as a coupled set of advection equations. In doing so, the
terms containing derivatives (in space or time) of the pressure
are treated as source terms. This approach, hence, sidesteps an
important guideline for the formulation of nonlinear hyperbolic
systems of equations, namely the preservation of their
conservation form
. This is a necessary condition to guarantee correct evolution in
regions of sharp entropy generation (i.e., shocks). Furthermore,
some amount of numerical dissipation must be used to stabilize
the solution across discontinuities. In this spirit, the first
attempt to solve the equations of general relativistic
hydrodynamics in the original Wilson's scheme [300] used a combination of finite difference upwind techniques with
artificial viscosity terms. Such terms adapted the classic
treatment of shock waves introduced by von Neumann and
Richtmyer [295
] to the relativistic regime (see Section
3.1.1).
Wilson's formulation has been widely used in hydrodynamical
codes developed by a variety of research groups. Many different
astrophysical scenarios were first investigated with these codes,
including axisymmetric stellar core collapse [195,
193
,
199,
22
,
276
,
228
,
79
], accretion onto compact objects [122
,
226
], numerical cosmology [53,
54
,
12] and, more recently, the coalescence of neutron star
binaries [303
,
304
,
169
]. This formalism has also been employed, in the special
relativistic limit, in numerical studies of heavy-ion
collisions [302,
175]. We note that in most of these investigations, the original
formulation of the hydrodynamic equations was slightly modified
by re-defining the dynamical variables, Equation (19
), with the addition of a multiplicative
factor (the lapse function) and the introduction of the Lorentz
factor,
:
As mentioned before, the description of the evolution of
self-gravitating matter fields in general relativity requires a
joint integration of the hydrodynamic equations and the
gravitational field equations (the Einstein equations). Using
Wilson's formulation for the fluid dynamics, such coupled
simulations were first considered in [301], building on a vacuum numerical relativity code specifically
developed to investigate the head-on collision of two black
holes [273]. The resulting code was axially symmetric and aimed to
integrate the coupled set of equations in the context of stellar
core collapse [82].
More recently, Wilson's formulation has been applied to the
numerical study of the coalescence of binary neutron stars in
general relativity [303,
304
,
169
] (see Section
4.3.2). These studies adopted an approximation scheme for the
gravitational field, by imposing the simplifying condition that
the 3-geometry (the 3-metric
) is
conformally flat
. The line element, Equation (11
), then reads
The curvature of the 3-metric is then described by a position
dependent conformal factor
times a flat-space Kronecker delta. Therefore, in this
approximation scheme all radiation degrees of freedom are
removed, while the field equations reduce to a set of five
Poisson-like elliptic equations in flat spacetime for the lapse,
the shift vector, and the conformal factor. While in spherical
symmetry this approach is no longer an approximation, being
identical to Einstein's theory, beyond spherical symmetry its
quality degrades. In [139] it was shown by means of numerical simulations of extremely
relativistic disks of dust that it has the same accuracy as the
first post-Newtonian approximation.
Wilson's formulation showed some limitations in handling
situations involving ultrarelativistic flows (), as first pointed out by Centrella and Wilson [54
]. Norman and Winkler [208
] performed a comprehensive numerical assessment of such
formulation by means of special relativistic hydrodynamical
simulations. Figure
1
reproduces a plot from [208
] in which the relative error of the density compression ratio in
the so-called relativistic shock reflection problem - the heating
of a cold gas which impacts at relativistic speeds with a solid
wall and bounces back - is displayed as a function of the Lorentz
factor
W
of the incoming gas. The source of the data is [54
]. This figure shows that for Lorentz factors of about 2 (
), which is the threshold of the ultrarelativistic limit, the
relative errors are between 5% and 7% (depending on the adiabatic
exponent of the gas), showing a linear growth with
W
.
Norman and Winkler [208] concluded that those large errors were mainly due to the way in
which the artificial viscosity terms are included in the
numerical scheme in Wilson's formulation. These terms, commonly
called
Q
in the literature (see Section
3.1.1), are only added to the pressure terms in some cases, namely at
the pressure gradient in the source of the momentum equation,
Equation (21
), and at the divergence of the velocity in the source of the
energy equation, Equation (22
). However, [208
] proposed to add the
Q
terms in a relativistically consistent way, in order to consider
the artificial viscosity as a real viscosity. Hence, the
hydrodynamic equations should be rewritten for a modified
stress-energy tensor of the following form:
In this way, for instance, the momentum equation takes the following form (in flat spacetime):
In Wilson's original formulation,
Q
is omitted in the two terms containing the quantity
. In general,
Q
is a nonlinear function of the velocity and, hence, the quantity
in the momentum density of Equation (26
) is a highly nonlinear function of the velocity and its
derivatives. This fact, together with the explicit presence of
the Lorentz factor in the convective terms of the hydrodynamic
equations, as well as the pressure in the specific enthalpy, make
the relativistic equations much more coupled than their Newtonian
counterparts. As a result, Norman and Winkler proposed the use of
implicit schemes as a way to describe more accurately such
coupling. Their code, which in addition incorporates an adaptive
grid, reproduces very accurate results even for ultrarelativistic
flows with Lorentz factors of about 10 in one-dimensional, flat
spacetime simulations.
Very recently, Anninos and Fragile [13] have compared state-of-the-art artificial viscosity schemes and
high-order non-oscillatory central schemes (see Section
3.1.3) using Wilson's formulation for the former class of schemes and
a conservative formulation (similar to the one considered
in [221
,
218
]; Section
2.2.2) for the latter. Using a three-dimensional Cartesian code, these
authors found that earlier results for artificial viscosity
schemes in shock tube tests or shock reflection tests are not
improved, i.e., the numerical solution becomes increasingly
unstable for shock velocities greater than about
. On the other hand, results for the shock reflection problem
with a second-order finite difference central scheme show the
suitability of such a scheme to handle ultrarelativistic flows,
the underlying reason being, most likely, the use of a
conservative formulation of the hydrodynamic equations rather
than the particular scheme employed (see Section
3.1.3). Tests concerning spherical accretion onto a Schwarzschild
black hole using both schemes yield the maximum relative errors
near the event horizon, as large as
% for the central scheme.
If a numerical scheme written in conservation form converges,
it automatically guarantees the correct Rankine-Hugoniot (jump)
conditions across discontinuities - the shock-capturing property
(see, e.g., [152]). Writing the relativistic hydrodynamic equations as a system
of conservation laws, identifying the suitable vector of
unknowns, and building up an approximate Riemann solver permitted
the extension of state-of-the-art
high-resolution shock-capturing
schemes (HRSC in the following) from classical fluid dynamics
into the realm of relativity [163
].
Theoretical advances on the mathematical character of the
relativistic hydrodynamic equations were first achieved studying
the special relativistic limit. In Minkowski spacetime, the
hyperbolic character of relativistic hydrodynamics and
magneto-hydrodynamics (MHD) was exhaustively studied by Anile and
collaborators (see [10] and references therein) by applying Friedrichs' definition of
hyperbolicity [100] to a quasi-linear form of the system of hydrodynamic
equations,
where
are the Jacobian matrices of the system and
is a suitable set of
primitive
(physical) variables (see below). The system (27
) is hyperbolic in the time direction defined by the vector field
with
, if the following two conditions hold: (i)
and (ii) for any
such that
,
, the eigenvalue problem
has only real eigenvalues
and a complete set of right-eigenvectors
. Besides verifying the hyperbolic character of the relativistic
hydrodynamic equations, Anile and collaborators [10] obtained the explicit expressions for the eigenvalues and
eigenvectors in the local rest frame, characterized by
. In Font et al. [93
] those calculations were extended to an arbitrary reference
frame in which the motion of the fluid is described by the
4-velocity
.
The approach followed in [93] for the equations of special relativistic hydrodynamics was
extended to general relativity in [21]. The choice of evolved variables (conserved quantities) in the 3+1 Eulerian formulation developed by Banyuls et
al. [21
] differs slightly from that of Wilson's formulation [300
]. It comprises the rest-mass density (D), the momentum density in the
j
-direction (
), and the total energy density (E), measured by a family of observers which are the natural
extension (for a generic spacetime) of the Eulerian observers in
classical fluid dynamics. Interested readers are directed
to [21
] for more complete definitions and geometrical foundations.
In terms of the so-called
primitive variables
, the conserved quantities are written as
where the contravariant components
of the 3-velocity are defined as
and
W
is the relativistic Lorentz factor
with
.
With this choice of variables the equations can be written in conservation form. Strict conservation is only possible in flat spacetime. For curved spacetimes there exist source terms, arising from the spacetime geometry. However, these terms do not contain derivatives of stress-energy tensor components. More precisely, the first-order flux-conservative hyperbolic system, well suited for numerical applications, reads
with
satisfying
with
. The state vector is given by
with
. The vector of fluxes is
and the corresponding sources
are
The local characteristic structure of the previous system of
equations was presented in [21]. The eigenvalues (characteristic speeds) of the corresponding
Jacobian matrices are all real (but not distinct, one showing a
threefold degeneracy as a result of the assumed directional
splitting approach), and a complete set of right-eigenvectors
exists. System (30
) satisfies, hence, the definition of hyperbolicity. As it will
become apparent in Section
3.1.2
below, the knowledge of the spectral information is essential in
order to construct HRSC schemes based on Riemann solvers. This
information can be found in [21
] (see also [96
]).
The range of applications considered so far in general
relativity employing the above formulation of the hydrodynamic
equations, Equation (30,
31
,
32
,
33
), is still small and mostly devoted to the study of stellar core
collapse and accretion flows onto black holes (see Sections
4.1.1
and
4.2
below). In the special relativistic limit this formulation is
being successfully applied to simulate the evolution of
(ultra-)relativistic extragalactic jets, using numerical models
of increasing complexity (see, e.g., [167,
8
]). The first applications in general relativity were performed,
in one spatial dimension, in [163
], using a slightly different form of the equations. Preliminary
investigations of gravitational stellar collapse were attempted
by coupling the above formulation of the hydrodynamic equations
to a hyperbolic formulation of the Einstein equations developed
by [39
]. These results are discussed in [161
,
38
]. More recently, successful evolutions of fully dynamical
spacetimes in the context of adiabatic stellar core collapse,
both in spherical symmetry and in axisymmetry, have been
achieved [129
,
244
,
67
]. These investigations are considered in Section
4.1.1
below.
An ambitious three-dimensional, Eulerian code which evolves
the coupled system of Einstein and hydrodynamics equations was
developed by Font et al. [96] (see Section
3.3.2). The formulation of the hydrodynamic equations in this code
follows the conservative Eulerian approach discussed in this
section. The code is constructed for a completely general
spacetime metric based on a Cartesian coordinate system, with
arbitrarily specifiable lapse and shift conditions. In [96
] the spectral decomposition (eigenvalues and right-eigenvectors)
of the general relativistic hydrodynamic equations, valid for
general spatial metrics, was derived, extending earlier results
of [21
] for non-diagonal metrics. A complete set of left-eigenvectors
was presented by Ibáñez et al. [127]. Due to the paramount importance of the characteristic
structure of the equations in the design of upwind HRSC schemes
based upon Riemann solvers, we summarize all necessary
information in Section
5.2
of this article.
![]() |
Numerical Hydrodynamics in General Relativity
José A. Font http://www.livingreviews.org/lrr-2003-4 © Max-Planck-Gesellschaft. ISSN 1433-8351 Problems/Comments to livrev@aei-potsdam.mpg.de |