3D Relativistic Hydrodynamics
3D Relativistic Hydrodynamics
3D Relativistic Hydrodynamics
1 Introduction
classical and relativistic equations. Sections 3, 4 and 5 form the body of the
review. Section 3 is devoted to discuss different approaches for the integration of
the RHD eqs. paying special attention to the most recent numerical algorithms.
Present numerical 3D RHD codes are reviewed in §4 as well as some compu-
tational issues relevant for three dimensional simulations. Several astrophysics
applications are discussed in §5. We finish this work with a summary (6).
The equations that describe the evolution of a relativistic fluid can be written
as covariant divergences,
∇·J = 0, (1a)
∇·T = 0, (1b)
J = ρuµ , µ, ν = 0, . . . , 3 (2a)
µν µ µ µν
T = ρhu u + pg , (2b)
ρ, p, h, ε and uµ being the rest–mass density, the pressure, the specific enthalpy
(h = 1 + ε + p/ρ), the specific internal energy and the four–velocity of the
fluid, respectively. The system of equations (1a, 1b) is closed making use of the
normalization condition of the four velocity (uµ uµ = −1, where summation is
extended over repeated indices) and an equation of state (EoS), usually of the
form p = p(ρ, ε).
A consistent numerical simulation of the flow evolution within the framework
of Relativity requires, in principle, the solution of the RHD equations (relativis-
tic counterparts of the Euler or Navier–Stockes equations) coupled to the full
set of Einstein field equations (that control the evolution of the space–time).
However, the problem is so complex (especially in multidimensions), and the
variety of astrophysical phenomena is so wide that for practical purposes it is
worthy to consider different physical approximations. The easiest one is to con-
sider that the gravitational field is unimportant and, therefore, the space–time
is flat or, in other words, the metric of the space–time is the Minkowski metric.
This is the approach of Special Relativistic hydrodynamic (SRHD) simulations
and has been successfully applied to, e.g., extragalactic jets (e.g., [101]), after-
glows of gamma–ray bursts (GRBs) [122], and also in other fields of physics,
like e.g., relativistic heavy–ion collisions [150]. Another possibility is to assume
that the simulated flow is a test fluid evolving in an external static field created
3D Relativistic Hydrodynamics 3
where
u · ∂i
vi =
−u · n
ui βi
v i = γ ij vj = + ,
αut α
and Γ ≡ −u · n = (1 − γij v i v j )−1/2 is the Lorentz factor.
λ0 = v i , (7a)
1 ³
i 2
p ´
λ± = v (1 − cs ) ± cs (1 − v 2 )[1 − v 2 c2s − v i v i (1 − c2s )] . (7b)
1 − v 2 c2s
Let us notice that: (i) there exists a strong coupling between the components
of velocity along the different spatial directions through the modulus of the ve-
locity, v; (ii) in the one–dimensional case (1D) the expressions of λ± (associated
to the acoustic waves) reduce to the Lorentz addition of the flow velocity and
the local sound speed
v ± cs
λ± = ; (8)
1 ± vcs
3D Relativistic Hydrodynamics 5
The classical Euler equations are easily recovered from the RHD equations in
the limit c → ∞. The limits of the conserved variables and flux vectors lead to
the corresponding quantities in the classical case:
1
U = (D, S j , τ ) → (ρ, ρv j , ρv2 + ρε)
2
1
F(i) = (Dv i , S j v i + pδ ji , S i − Dv i ) → (ρv i , ρv j v i + pδ ji , v i ( ρv2 + ρε + p))
2
i, j = 1, 2, 3
The equations keep their conservative and hyperbolic characters but there are
several factors that make RHD more complex to solve numerically than classical
hydrodynamics: (i) the RHD equations are tightly coupled through Γ and h and,
therefore, they display a larger non-linearity; (ii) there is no explicit relation be-
tween W and U (except for particular EoS), i.e., obtaining the primitive from
the conserved variables needs an iterative numerical method; (iii) the tangential
flow velocity can change across discontinuities (see § 2.2) and, in addition, the
characteristic speeds may suffer from aberration (see the coupling between dif-
ferent directions in expression (7b)); (iv) in the ultrarelativistic limit (v → 1),
the eigenfields are degenerate (λ0 → λ± → 1) which triggers the appearance
of very thin structures in the flow (like, e.g., in the case of relativistic blast
waves) that may become a source of numerical errors; and (v) relativistic strong
shocks can display unbounded jumps in physical variables; e.g., for an ideal gas
(i.e., with an EoS: p = (γ − 1)ρε), the compression ratio between the post– and
pre–shocked densities of a relativistic strong shock is such that (see, e.g., [155])
ρb /ρa ≤ (γΓa + 1)/(γ − 1) which tends to infinity in the ultrarelativistic limit
(va → 1). This should be compared with the compression ratio in a Newtonian
strong shock: ρb /ρa ≤ (γ + 1)/(γ − 1) which is ∼ 4 − 7 for typical values of γ.
(a)
ρ e ρ = 1, e=0 (b)
2 2 1 1 10.0
Γ=5/3
Γ=4/3
8.0
Relative error (%)
v=0 |v|<c
2 1
2.0
1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4
W
Fig. 1. (a) Scheme of the relativistic shock reflection test. (b) Relative errors of the
relativistic shock reflection test as a function of the Lorentz factor (W in the plot) of
the inflowing gas for two different values of the adiabatic index (Γ in the plot legends)
using the explicit Eulerian techniques of [24]. Data from Centrella and Wilson [24].
Plot reproduced from Norman and Winkler [118].
dUnj 1 ³
=− F̂j+1/2 (Unj−r , Unj−r+1 , . . . , Unj+q )−
dt ∆x ´
F̂j−1/2 (Unj−r−1 , Unj−r , . . . , Unj+q−1 ) .
where p and q are positive integers. Usually, Unj is an approximation to the zone
average of U defined by
Z xj+1/2
1
Unj = U(tn , x)dx , (17)
∆x xj−1/2
which is consistent with the integral form of the system of conservation laws
and, F̂j±1/2 are the time–averaged fluxes across the interfaces xj±1/2 (xj±1/2 =
(xj + xj±1 )/2) :
Z tn+1
1
F̂j±1/2 ≈ F(U(t, xj±1/2 ))dt . (18)
∆t tn
3D Relativistic Hydrodynamics 9
tends to zero as ∆x → 0 (Ūnj is the average of the true solution in the cell j).
The idea behind definition (19) is that the finer the grid is, the better the numer-
ical solution. Lax equivalence theorem [133] asserts that stability is a necessary
condition to guarantee convergence. A measure of the stability of a solution can
be its total variation at t = tn , TV(Un ), defined as
+∞
X
TV(Un ) = |Unj+1 − Unj | . (20)
j=0
n
u j-1 u jn u nj+1 n
u j+2 n
t
x j-1 x j-1/2 xj xj+1/2 x j+1 x j+3/2 xj+2
continuous solution
discrete solution
Time advance. There are two main procedures to advance the conserved vari-
ables in time. One possibility is a standard discretization of the time derivative
in system (17). In this case, the accuracy of the method depends on the order
of accuracy up to which the numerical fluxes have been computed. The sec-
ond alternative (known as the method of lines) regards (17) as a semidiscrete
(i.e., spatially discretized) system of ordinary differential equations to which any
standard ordinary differential equation solver (e.g., a Runge Kutta solver) can
be applied.
matrices of the flux vectors, B(i) , extending a procedure introduced by Roe [132]
and sometimes referred to as local characteristic approach (LCA). The idea of
the LCA is to use the spectral decomposition of B(i) to rewrite the original sys-
tem as a new one of uncoupled scalar equations, in terms of the characteristic
variables. The locally linear system can then be easily solved to obtain appropri-
ate numerical fluxes for the original system. Sometimes the linearization process
involves an averaged intermediate state at zone interfaces. This is the case of the
original Roe solver [132] and its relativistic extension [39], and Martı́ et al. [97]
or Falle & Komissarov [41] approaches. Donat & Marquina [36] have extended a
numerical flux formula which was first proposed by Shu & Osher [147] for scalar
equations to systems. In the scalar case and for characteristic wave speeds which
do not change sign zone interfaces, Marquina’s flux formula is identical to Roe’s
flux. Otherwise, the scheme switches to the more viscous, entropy satisfying lo-
cal Lax–Friedrichs scheme [147]. In the case of systems, the combination of Roe
and local–Lax–Friedrichs solvers is carried out in each characteristic field after
the local linearization and decoupling of the system of equations [36]. However,
contrary to Roe’s and the previously cited linearized methods, the extension of
Marquina’s method to systems does not require on any averaged intermediate
state. Marquina’s flux formula has been successfully used in the ultrarelativis-
tic regime and in 2D axisymmetric [100,101] and 3D SRHD [3] and even in 3D
GRHD [7,44].
Finally, a very simple approach is an extension of the Harten, Lax, van Leer
(HLL) solver to SRHD [139]. This method avoids the explicit calculation of the
eigenvalues and eigenvectors of the Jacobian matrices and is based on an approx-
imate solution of the original Riemann problems with a single intermediate state,
U∗ , determined by requiring consistency of the approximate Riemann solution
with the integral form of the conservation laws in a grid zone. The algorithm
needs estimates of lower and upper bounds for the smallest and largest signal
velocities, aL and aR , respectively. Good estimates for aL and aR are essential to
guarantee robustness and, at the same time, the minimal amount of numerical
viscosity (the larger the difference ||aL | − |aR || is, the larger the viscosity of the
method; if the difference is too small, undesired numerical oscillations around
discontinuities appear). In the non-relativistic case, Einfeldt [37] proposed to
use the smallest and largest eigenvalues of Roe’s matrix (this corresponds to the
HLLE solver). Duncan & Hughes [33] have generalized the method to 2D SRHD.
Relativistic beam scheme. In the beam scheme [136] and its relativistic
extension [167] the hydrodynamic equations are solved as the limit of the cor-
responding Botzmann equation. The velocity distribution functions are approx-
imated by several Dirac delta functions or discrete beams of particles in each
computational cell, which reproduce the appropriate moments of the distribu-
tion functions. This beam scheme, although being a particle method derived
from a microscopic kinetic description, has all the desirable properties of mod-
ern characteristic–based wave propagation methods based on a macroscopic con-
tinuum description. Yang et al. [167] show that the integration scheme for the
beams can be cast in the form of an upwind conservation scheme in terms of nu-
merical fluxes and build up high–order variants of the relativistic code in terms
of different TVD and ENO interpolations.
There are, according to our knowledge, four different 3D RHD codes, used in
astrophysical applications: Koide–Nishikawa’s code [74,76], GENESIS [3], Cactus
[44] and Shibata–Nakamura’s code [143]. In the following we will describe some
of the main features of each one.
The group of Koide and coworkers has developed a GRMHD code (for a
fixed background metric) based on a sTVD scheme (see Sect. 3.3). The code
has been applied to study the propagation of extragalactic jets through magne-
tized atmospheres [115] and also to simulate the early stages of jet formation by
magnetohydrodynamic mechanisms in background BH spacetimes [75]. A series
of tests of Koide et al.’s method involving mainly continuous solutions can be
found in ref. [76]. Koide et al.’s code has proven to be very stable (although very
difussive) when simulating mildly relativistic flows (maximum Lorentz factors
≈ 4) with discontinuities.
GENESIS [3] is a conservative 3D RHD code (used in astrophysical SRHD
and GRHD applications) based on HRSC techniques. It uses Marquina’s flux for-
mula to compute numerical fluxes and a third order PPM spatial interpolation.
The time advance is performed by means of a high–order Runge–Kutta method.
16 M.A. Aloy and J.M. Martı́
analytic solution
1.0 density/10
pressure/20
total velocity
0.8
0.6
0.4
0.2
0.0
−0.2
−0.5 −0.3 −0.1 0.1 0.3 0.5
main diagonal main
maindiagonal
diagonal
Fig. 3. Comparative performance of Cactus (left panel; figure from [44]) and GEN-
ESIS (right panel; figure from [3]) on a three–dimensional mildly relativistic shock
tube test. The panels show several primitive variables along the main diagonal of
the computational domain (solid line: analytic solution; symbols: numerical solu-
tion). The initial data are a constant left (L) and right (R) state characterized by
(ρL , pL , vL ) = (10, 13.3, 0) and (ρL , pL , vL ) = (1, 0.66 · 10−6 , 0). Numerical grid and
evolution time is similar in both cases.
GENESIS has been extensively tested in problems involving strong shocks even
in the ultrarelativistic regime, being able to handle Lorentz factors as large as
3 · 105 (in the one–dimensional wall reflection test). Simulations of extragalactic
jets (using SRHD, at both parsec [6] and kiloparsec [4] scales) and progenitors
of GRBs (in GRHD with a Schwarzschild background metric, [7]) have been
performed with GENESIS.
Cactus is a numerical tool developed within a collaboration of the Numerical
Relativity divisions at the National Center for Supercomputing Applications,
the Albert Einstein Institut and the Washington University [22]. Cactus is able
to solve the full set of Einstein field equations coupled to a perfect fluid source.
The hydrodynamic evolution employs HRSC methods and can be computed by
means of different Riemann solvers (Roe, Marquina) and flux–splitting schemes.
The metric evolution may be followed with several formalisms (ADM, hyperbolic
formulations, conformal–tracefree) combined with a number of integration meth-
ods (leapfrog, Crank–Nicholson, MacCormack along with Strang splitting) and
various gauges (algebraic, maximal slicing, etc.). The main goal of the code is
the simulation of astrophysical processes involving NSs and BHs. Font et al. [44]
have tested the capabilities of the hydro part computing shock tubes. In Fig. 3,
a comparison between Cactus and GENESIS is shown for a mildly relativistic
Riemann problem. Despite the slightly different grids (1003 for GENESIS, 1283
3D Relativistic Hydrodynamics 17
for Cactus) the results are still comparable, the reason being that both codes
use the same HRSC techniques. Font et al. [44] have also tested the GRHD cou-
pling in Cactus simulating Friedman–Robertson–Walker cosmologies with dust
and polytropic NSs (static and boosted). Alcubierre et al. [2] have performed a
number of experiments (Brill waves, single BH, static boson and neutron stars)
comparing different formulations of the Einstein equations. Recently, Landry &
Teukolsky [86] have performed simulations of coalescing binary NSs. The sta-
bility of the Einstein evolution is a main issue of this code, especially when the
ADM or hyperbolic formulations are considered.
Γ = 22360
5 Applications
In this Section we will review briefly some of the most relevant 3D applications
of SRHD and SRMHD codes in astrophysical scenarios (i.e., relativistic jets) as
well as recent GRHD and GRMHD simulations in the context of progenitors of
GRBs and jet formation, respectively. Applications to dynamical space–times,
in which Einstein equations are coupled to the GRHD equations, are beyond the
scope of this article (for a review see, e.g., [45]).
In the standard model [19,138] the elongated radio structures connected to the
center of AGNs in radio galaxies and radio–loud quasars are considered as con-
tinuous ejections of highly collimated, supersonic and very stable plasma. The
emission is produced by synchrotron and inverse Compton processes of electrons
accelerated up to ultrarelativistic energies in the vicinity of a central engine.
The asymmetries in the radio flux of the two oppositely directed jets of a source
and the superluminal motions observed in a few dozens of compact sources are
explained by assuming that both jets propagate with relativistic speeds along
directions at small angles to the line of sight. The relativistic Doppler beaming
of the emission in the direction of motion accounts for the observed emission
asymmetries whereas apparent superluminal speeds are explained by the com-
bination of a finite value of the speed of light and the relativistic motion of the
emitting source.
3D Relativistic Hydrodynamics 19
Large scale jets. At far enough distances from the central object (pc and
kpc scales) the effects of gravity are negligible and we can assume (as a first
approximation to the problem) that magnetic fields are dynamically unimpor-
tant. In this approach, numerical simulations using a pure SRHD treatment have
been performed since the early 90s to study the morphology and dynamics of
relativistic jets.
The development of codes based on HRSC techniques has allowed 2D axisym-
metric time–dependent relativistic hydrodynamic simulations [33,38,99–101,80,134]
to be performed. These simulations led to the conclusion that both the internal
energy and the Lorentz factor of the beam enhance the stability of relativistic
jets compared to their classical counterparts, through the increase of the effective
inertial mass of the beam. Relativistic MHD simulations in 2D using pseudo–
spectral techniques [159,161] or sTVD methods [74,73] have been another step
forward in our understanding of relativistic astrophysical jets.
Only since 1998 the morphology and dynamics of relativistic jets is studied
with 3D SRHD [3,4] or SRMHD [115,116] simulations. Aloy et al. [3] did a 3D
simulation (in Cartesian coordinates) of an axisymmetric relativistic jet propa-
gating through an homogeneous atmosphere. The simulated jet is characterized
by a beam–to–external proper rest–mass density ratio η = 0.01, a beam Mach
number Mb = 6.0, and a beam flow speed vb = 0.99c. An ideal gas equation
of state with an adiabatic exponent γ = 5/3 is assumed to describe both the
jet matter and the ambient gas. The beam is in pressure equilibrium with the
ambient medium which fills a domain (X,Y,Z) with a size of 15Rb × 15Rb × 75Rb
(120 × 120 × 600 cells), where Rb is the beam radius. The jet is injected at z = 0
in the direction of the positive z-axis through a circular nozzle. Simulations were
typically performed with 16 R10000 processors (on a SGI-Origin 2000) and need
about ten thousand time iterations. The execution time was about 100 hours.
Genuine multidimensional effects were included by perturbing the axial injection
velocity.
Koide and collaborators [115,116] simulated the evolution of relativistic jets
through a magnetized uniform atmosphere during a very brief period of time. The
numerical setup [116] consisted on a cylindrical jet injected through a circular
nozzle into an oblique (45◦ with respect to the jet axis) magnetic atmosphere.
The computational domain was a cubic box of 20Rb ×20Rb ×20Rb (101×101×101
cells). The jet had the following parameters: vb = 0.98756c, ratio between the
beam and magnetic specific energy densities εb /εm = 1/3, Mb = 4.0, γ = 5/3
and considered two different strengths for the ambient magnetic field, weak and
strong. The simulation lasted for 35 CPU hours on a SGI–Power Challenge and
required about 1 Gb of RAM. It should be remarked that the coarse grid zoning
used in Nishikawa’s [116] simulations (5 cells/Rb ), prevented them from studying
genuine 3D effects in relativistic jets in detail.
Parsec scale jets. The presence of emitting flows at almost light speed en-
hances the importance of relativistic effects in the apparency of jets. This fact
is stressed in the case of parsec scale jets, triggering the combination of syn-
3D Relativistic Hydrodynamics 21
Fig. 5. Figure from [6]. Logarithm of the integrated total ( left) and polarized ( right)
intensity across the jet for different viewing angles. Lines are plotted in intervals of 10◦
from an angle of 10◦ (top line in both plots), to 90◦ (showing a progressive decrease in
emission). Dashed lines (dot dashed) correspond to an observing angle of -100◦ (-140◦ ).
Units are normalized to the maximum in total intensity.
considered and with a maximum velocity of 0.4c if a free falling corona embed-
ding the black hole is assumed. A two layered jet structure (in agreement with
theoretical predictions – e.g., [18] –) is found. The inner part is pressure driven
and moves at relativistic speeds in the hydrostatic corona case. The outermost
part is magnetically driven and subrelativistic. This shell structure might be
the origin of the shear layer mentioned in the previous section. Nevertheless,
for a fast rotating BH [77], the maximum velocities obtained are 0.4c (counter–
rotating disk) and 0.3c (co–rotating disk), and again the two layered outflow
structure is formed (see Fig. 6). As the simulations had to be stopped due to nu-
merical problems quite early [77], the total evolution time was not large enough
to develop highly relativistic jets. Therefore, despite of the promising results of
Koide’s group, the mechanism of jet formation still remains an open and chal-
lenging question (see, e.g., [135]).
GRBs are known observationally since over 30 years [71]. They consist of very
short, non–repeating events (except for a few soft gamma–ray repeaters), with
a typical duration between several milliseconds and several hundreds of seconds,
showing a large variability even at millisecond scale. They show a bimodal time–
distribution, the border between the two groups being at ∼ 2 s. The first group is
composed of short bursts centered around 0.1 s, while the second group consists
of long bursts (more numerous and softer than the first group) centered at
about 15 s. The time–structure is very different from burst to burst.
GRB spectra are non–thermal. The observed energy flux as a function of the
energy can be well described by one or a combination of several power laws.
The maximum of the energy distribution corresponds to an energy (the energy
3D Relativistic Hydrodynamics 23
Fig. 6. Figure from [76]. Initial and final time snapshots of the logarithm of the density
around a non rotating Schwarzschild BH. Top panels: free falling corona case. Bottom
panels: hydrostatic corona case. The solid lines are the magnetic field lines. The vector
plots show the flow velocity. On the initial state a uniform axial magnetic field is set
up. The disk (in white color) rotates around the BH with Keplerian velocity. In the
right panels the jet is formed almost along the magnetic field lines.
peak), which is characteristic of each GRB and usually is about several hundreds
of keV. The observed fluence on earth is 10−5 − 10−7 erg/cm2 .
For years it was unclear whether GRBs take place at local or cosmological
distances (see e.g., [105]). However, a galactic origin can ve excluded, because
the BATSE catalog shows an isotropic distribution of GRBs over the sky [104].
BeppoSAX spacecraft [29] has provided accurate coordinates (∼ arc minutes) of
the fading X–ray counterparts of GRBs, which has allowed for subsequent ground
based observations of faint GRB afterglows at optical and radio wavelengths.
The recent redshift determinations, obtained from the optical spectra, prove
24 M.A. Aloy and J.M. Martı́
the cosmological origin at least of the majority of GRBs (see [48] for more
information). Observed redshift values are in the range 0.7 ≤ z ≤ 3.4 implying
emitted gamma-ray energies of 2 × 1051 ≤ E ≤ 2.3 × 1054 erg for an isotropically
radiating source. The cosmological origin of the GRBs is consistent with the
distribution of bursts in the log N − log P plane.
Nonetheless, the accuracy of the positioning is neither sufficient to determine
the host galaxies of GRBs nor their progenitors. This picture was challenged by
the detection of the Type Ib/c supernova SN 1998bw [46,47] within the error box
of GRB 980425 [149,126] whose explosion time and location is consistent with
that of the GRB. This suggests a relationship between GRBs and SNe Ib/c,
i.e., core collapse supernovae of massive stellar progenitors which have lost their
hydrogen and helium envelopes [47,66]. However, the observation of a second fad-
ing X–ray source within the error box of GRB 980425 (different from SN 1998bw)
still causes some doubts on the GRB–supernova connection, although the prob-
ability of chance coincidence of GRB 980425 and SN 1998bw is almost negligible
[126].
Another clue on the nature of the progenitors of GRBs comes from the du-
ration of the shorter bursts and the temporal substructure of the longer bursts
(∼ 1 msec). If the time variation is intrinsic, the length scales involved in the
production of a GRB are of about 1 light–millisecond, which in turn points to-
wards compact objects, like NSs or BHs. Furthermore, the non recurrence of the
events points towards cataclysmic astrophysical events.
The compact nature of GRB sources, the observed flux and the cosmological
distance taken together imply a large photon density and, therefore, a large op-
tical depth for pair production. This is, however, inconsistent with the optically
thin source indicated by the non–thermal gamma–ray spectrum, which extends
well beyond the pair production threshold at 0.5 MeV. This problem ( compact-
ness problem) can be resolved by assuming an ultra–relativistic expansion of the
emitting region. The bulk Lorentz factor required are Γ > 100 (see, e.g., [128])
Various catastrophic collapse events have been proposed in order to ex-
plain the energies released in a GRB. Among those proposals we find neutron–
star/neutron–star mergers [120,52], neutron–star/black–hole mergers [108], col-
lapsars [166,90] and hypernovae [121]. These models rely on the existence of
a stellar mass BH hole which accretes several solar masses of matter from
a disk (formed during a merger or by a non–spherical collapse) at a rate of
∼ 1 M⊙ s−1 [131]. A fraction of the gravitational binding energy released by
accretion is converted into neutrino and anti–neutrino pairs, which in turn an-
nihilate into electron–positron pairs. This creates a pair fireball, which will also
include baryons present in the environment surrounding the black hole. If the
baryon load (the ratio of the fireball mass to its energy) of the fireball is small
enough, the baryons are accelerated together with the e+ e− pairs to ultra–
relativistic speeds with Lorentz factors > 102 [23,129]. The existence of such
relativistic flows is supported by radio observations of GRB 980425 [82]. The
rapid temporal decay of several GRB afterglows is inconsistent with spherical
(isotropic) blast wave models propagating through the interstellar medium, and
3D Relativistic Hydrodynamics 25
instead is more consistent with the evolution of a relativistic jet after it slows
down and spreads laterally [137]. Finally, the bulk kinetic energy of the fireball is
thought to be converted into gamma–rays via cyclotron radiation and/or inverse
Compton processes (see, e.g., [105,128]).
One–dimensional numerical simulations of spherically symmetric relativis-
tic fireballs have been performed by several authors to model GRB sources
(e.g., [129,123,124,59]). Multi–dimensional modeling of ultra–relativistic jets in
the context of GRBs has for the first time been attempted by Aloy et al. [7]. Us-
ing a collapsar progenitor model (from [90]) they have simulated the propagation
of an axisymmetric jet through the mantle and envelope of a collapsing massive
star using a version of GENESIS [3] that includes a background Schwarzschild
metric. The jet forms as a consequence of an assumed energy deposition rate
of 1050 − 1051 erg/sec within a 30◦ cone around the rotation axis. When the
jet reaches the surface of the stellar progenitor, the maximum Lorentz factor
attained by the flow is about 20. The latter fact implies that Newtonian sim-
ulations of this phenomenon [90] are clearly inadequate. The simulations also
try to address the ulterior acceleration of the fireball when the jet propagates
through an atmosphere of declining density. At the end of the simulations (when
the jet has gone over ∼ 1011 cm) the maximum Lorentz factor is about 50 (for an
energy deposition rate of 1051 erg/sec). The baryonic contamination is very het-
erogenous having an average value of 1. However, there are regions (coincident
with the parts of the flow having the largest Lorentz factor) where this value is
as small as 10−5 , which is in agreement with the theoretical expectations (see
above). Although the final Lorentz factor is small compared with the predictions
of the standard model [23,129], the distance up to which the jet propagation has
been tracked (8 × 1010 cm) is much smaller than the one assumed to produce
efficient internal shocks (1012 −1014 cm) and the fireball to become optically thin
(∼ 1013 cm, [128]). Therefore, there is still room for further acceleration of the
jet until it becomes transparent. Finally, there are recent claims pointing to the
possibility of GRB generation without extremely high Lorentz factors and with
much smaller masses than in the standard model [153].
6 Summary
Hydrodynamic relativistic processes are on the basis of a variety of challenging
astrophysical phenomena. On the other hand, relativistic astrophysics has ben-
efited of the recent development of accurate numerical techniques which have
allowed, for the first time, the simulation of ultrarelativistic multidimensional
flows. Two main applications are currently addressed. One is in the field of
relativistic jets, where very important advances have been made in problems
like the jet formation mechanisms or the nature of superluminal sources. The
other main application is in the field of GRBs. In order to stress the importance
of relativistic hydrodynamic simulations in these fields, let us remind that the
generation of both relativistic jets and GRBs is hidden to present observations
making numerical simulations the only means to confront the theoretical models.
26 M.A. Aloy and J.M. Martı́
In the present review we have rewritten the equations governing the dynam-
ics of relativistic perfect fluids paying special attention to their conservative and
hyperbolic character (two properties which are extensively exploited by mod-
ern numerical techniques). We have also reviewed the evolution of numerical
relativistic hydrodynamics, that started as a branch of relativistic astrophysics
more than thirty years ago with the pioneering simulations of May and White
of stellar core collapse, and which has culminated in the last decade with the
introduction of high–resolution shock–capturing (HRSC) methods. The basics
of HRSC methods have been summarized, too.
Most of the problems that may be treated by means of numerical simula-
tions require a multidimensional modeling of the flows. Fully three–dimensional
simulations in relativistic hydrodynamics are particularly challenging because of
their intrinsic, technical difficulties. These difficulties have been addressed only
by a few scientific groups, so far. We have listed those codes that have been used
in 3D special or general RHD simulations. Two of them (Cactus and Shibata–
Nakamura codes) have implemented a full consistent evolution of the metric of
the space–time coupled to the GRHD equations. This fact has allowed them to
study, for the first time, problems as complex as the coalescence of simplified
models of NSs. The other two codes (GENESIS and Koide codes) allow for a
static background metric of the space–time and even for the inclusion of the
magnetic field in the equations (Koide’s code).
The last Section has been devoted to outline the main results in the simula-
tion of relativistic extragalactic jets and GRBs.
Acknowledgements
This work has been supported in part by the Spanish DGES (grant PB 97-
1432). M.A. Aloy thanks to MPA for financial support under its guest program
and to the Spanish MEC for the a grant (EX 00 22566499). The authors thank
J.M. Ibáñez and E. Müller for the critical reading of the manuscript.
References
1. C.E. Akujor: Astron. Astrophys. 259, L61 (1992)
2. M. Alcubierre, B. Bruegmann, et al. :Phis. Review D 62, 044034 (2000)
3. M.A. Aloy, J.Ma . Ibáñez, J.Ma . Martı́, E. Müller: Astrophys. J. Suppl. 122, 151
(1999)
4. M.A. Aloy, J.Ma . Ibáñez, J.Ma . Martı́, J.L. Gómez, E. Müller: Astrophys. J. Lett.
523, L125 (1999)
5. M.A. Aloy, J. Pons, J.Ma . Ibáñez, Comput. Phys. Commun. 120, 115 (1999)
6. M.A. Aloy, J.Ma . Ibáñez, J.Ma . Martı́, J.L. Gómez, E. Müller: Astrophys. J. Lett.
528, L85 (2000)
7. M.A. Aloy, E. Müller, J.Ma . Ibáñez, J.Ma . Martı́, A. MacFadyen: Astrophys. J.
Lett. 531, L119 (2000)
8. A.M. Anile. Relativistic Fluids and Magnetofluids, (Cambridge University Press,
Cambridge 1989)
9. R. Arnowitt, S. Deser, C.W. Misner. Gravitation: An Introduction to Current Re-
search, ed., Witten, L., (John Wiley, New York, 1962) p. 227
10. J.M. Attridge, D.H. Roberts, J.F.C. Wardle: Astrophys. J. Lett. 518, L87 (1999)
3D Relativistic Hydrodynamics 27
129. T. Piran, A. Shemi, R. Narayan: Mon. Not. Roy. Astronom. Soc. 263, 861 (1993)
130. J.A. Pons, J.Ma Martı́, E. Müller: J. Comp. Phys. 422, 125 (2000)
131. R. Popham, S.E. Woosley, C. Fryer: Astrophys. J. 518, 356 (1999)
132. P.L. Roe: J. Comp. Phys. 43, 357 (1981)
133. R.D. Richtmyer, K.W. Morton: Difference Methods for Initial–value Problems,
(Wiley–Interscience, New York 1967)
134. Rosen, A., Hughes, et al. : Astrophys. J. 516, 729 (1999)
135. C. Sauty (2000), this volume.
136. R.H. Sanders, K.H. Prendergast: Astrophys. J. 188, 489 (1974)
137. R. Sari, T. Piran, J.P. Halpern: Astrophys. J. Lett. 519, L17 (1999)
138. P.A.G. Scheuer: Mon. Not. Roy. Astronom. Soc. 166, 513 (1974)
139. V. Schneider, U. Katscher, et al. : J. Comp. Phys. 105, 92 (1993)
140. M. Shibata, T. Nakamura: Phis. Review D 52, 5428 (1995)
141. M. Shibata: Prog. Theor. Phys. 101, 251 (1999)
142. M. Shibata: Prog. Theor. Phys. 101, 1199 (1999)
143. M. Shibata: Phis. Review D 60, 104052 (1999)
144. M. Shibata, T.W. Baumgarte, S.L. Shapiro: Phis. Review D 61, 044012 (2000)
145. S. Siegler, H. Riffert: Astrophys. J. Suppl. 531, 1053 (2000)
146. C.W. Shu: Math. Comp. 49, 105 (1987)
147. C.W. Shu, S.J. Osher: J. Comp. Phys. 83, 32 (1989)
148. L. Smarr, J.R. Wilson, R.T. Barton, R.L. Bowers: Astrophys. J. 246, 515 (1981)
149. P. Soffitta, M. Feroci, et al. : IAU Circ. 6884 (1998)
150. J. Sollfrank, P. Huovinen, M. Kataja, P.V. Ruuskanen, M. Prakash, R. Venu-
gopalan: Phys. Rev. C 55, 392 (1997)
151. R.F. Stark, T. Piran: Comput. Phys. Rep. 5, 221 (1987)
152. M.R. Swain, A.H. Bridle, S.A. Baum: Astrophys. J. 507, L29 (1998)
153. J.C. Tan, C.D. Matzner, C.F. McKee: Astrophys. J. , submitted, (2000)
154. M. Taub: Phys. Rev. 74, 328 (1948)
155. K.W. Thomson: J. Fluid Mech. 171, 365 (1986)
156. S.J. Tingay, D.L. Jauncey, R.A. Preston, J.E. Reynolds, D.L. Meier et al. : Nature
374, 141 (1995)
157. E. Toro: Riemann solvers and numerical methods for fluid dynamics: a practical
introduction, 1st edn. (Springer, Berlin 1997)
158. B. van Leer: J. Comp. Phys. 32, 101 (1979)
159. M.H.P.M. van Putten: J. Comp. Phys. 105, 339 (1993a)
160. M.H.P.M. van Putten: Astrophys. J. Lett. 408, L21 (1993b)
161. M.H.P.M. van Putten: Astrophys. J. Lett. 467, L57 (1996)
162. J. von Neumann, R.D. Richtmyer: J. Appl. Phys. 21, 232 (1950)
163. L. Wen, A. Panaitescu, P. Laguna: Astrophys. J. 486, 919 (1997)
164. J.R. Wilson: Astrophys. J. 173, 431 (1972)
165. J.R. Wilson. In: Sources of Gravitational Radiation, ed. by L.L. Smarr, (Cam-
bridge University Press, Cambridge 1979) p. 423
166. S.E. Woosley: Astrophys. J. 405, 273 (1993)
167. Yang, J.Y., Chen, M.H., Tsai, I-N. and J.W. Chang: J. Comp. Phys. 136, 19
(1997)