This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Dissipative Magnetohydrodynamics for Non-Resistive Relativistic Plasmas:
An implicit second-order flux-conservative formulation with stiff relaxation

Elias R. Most emost@princeton.edu Princeton Center for Theoretical Science, Princeton University, Princeton, NJ 08544, USA Princeton Gravity Initiative, Princeton University, Princeton, NJ 08544, USA School of Natural Sciences, Institute for Advanced Study, Princeton, NJ 08540, USA    Jorge Noronha jn0508@illinois.edu Illinois Center for Advanced Studies of the Universe, Department of Physics, University of Illinois at Urbana-Champaign, Urbana, IL 61801, USA
Abstract

Based on a 14-moment closure for non-resistive (general-) relativistic viscous plasmas, we describe a new numerical scheme that is able to handle all first-order dissipative effects (heat conduction, bulk and shear viscosities), as well the anisotropies induced by the presence of magnetic fields. The latter is parameterized in terms of a thermal gyrofrequency or, equivalently, a thermal Larmor radius and allows to correctly capture the thermal Hall effect. By solving an extended Israel-Stewart-like system for the dissipative quantities that enforces algebraic constraints via stiff-relaxation, we are able to cast all first-order dissipative terms in flux-divergence form. This allows us to apply traditional high-resolution shock capturing methods to the equations, making the system suitable for the numerical study of highly turbulent flows. We present several numerical tests to assess the robustness of our numerical scheme in flat spacetime. The 14-moment closure can seamlessly interpolate between the highly collisional limit found in neutron star mergers, and the highly anisotropic limit of relativistic Braginskii magnetohydrodynamics appropriate for weakly collisional plasmas in black-hole accretion problems. We believe that this new formulation and numerical scheme will be useful for a broad class of relativistic magnetized flows.

preprint: APS/123-QED

I Introduction

Relativistic fluid dynamics is widely used as an effective description of long wavelength, long time phenomena in a variety of many-body systems such as the quark-gluon plasma Romatschke and Romatschke (2019), the early universe Weinberg (2008), and the hot and ultradense matter formed in neutron star mergers Baiotti and Rezzolla (2017). In the context of heavy-ion collisions, experiments at the Relativistic Heavy Ion Collider and the Large Hadron Collider have provided overwhelming evidence that the quark-gluon plasma formed in highly energetic nuclear collisions behaves as a strongly-interacting relativistic fluid over distance scales not much larger than the size of a nucleus Heinz and Snellings (2013). Through the last decade (see Romatschke and Romatschke (2019) for a review), detailed comparisons between numerical hydrodynamic simulations and experimental data have shown that the matter formed in such collisions is not a perfect fluid, i.e., viscous effects are needed to describe experimental data. In fact, the quantitative extraction of the transport coefficients (such as shear and bulk viscosities) of the quark-gluon plasma from heavy-ion collision data Bernhard et al. (2019); Nijs et al. (2021); Everett et al. (2021a, b) provides key guidance toward understanding the novel out-of-equilibrium properties that emerge in deconfined quark-gluon matter at extreme temperatures and moderate baryon densities.

Hydrodynamic simulations of the quark-gluon plasma in heavy-ion collisions solve second-order relativistic viscous hydrodynamic equations of motion Israel and Stewart (1979), which may be obtained from entropy considerations, relativistic kinetic theory, or from a resummation of the gradient series Baier et al. (2008). Current formulations in use Baier et al. (2008); Denicol et al. (2012) share the same spirit of Israel-Stewart theory Israel and Stewart (1979) in which dissipative corrections to the energy-momentum tensor and the conserved currents behave as new degrees of freedom that obey additional (relaxation-type) equations of motion, which must be solved together with the conservation laws to determine the evolution of the fluid. Alternative formulations, such as anisotropic hydrodynamics Martinez and Strickland (2010); Florkowski and Ryblewski (2011); Bazow et al. (2014); Alqahtani et al. (2018, 2017), have also been used in heavy-ion simulations.

In the linearized regime around equilibrium, it is known that there are conditions involving the equation of state and the transport coefficients which, once fulfilled, guarantee stability and causality of second-order theories Hiscock and Lindblom (1983); Olson (1990). In the nonlinear regime probed in heavy-ion collisions, however, the backreaction from the dissipative corrections becomes important and more stringent conditions to ensure causality have been derived Bemfica et al. (2019a, 2021). See Chiu and Shen (2021); Plumberg et al. (2021) for discussions on the relevance of such nonlinear causality conditions in the description of the evolution of the quark-gluon plasma.

Numerical implementations of viscous, Israel-Stewart-like equations of motion can be found both in Eulerian Romatschke and Romatschke (2007); Schenke et al. (2012); Bozek (2012); Del Zanna et al. (2013); Karpenko et al. (2014); Shen et al. (2016); Habich et al. (2015); Romatschke (2015); Bazow et al. (2018); Okamoto and Nonaka (2017) and also in Lagrangian algorithms Noronha-Hostler et al. (2013, 2014, 2016). In the former, numerical methods originally written to solve flux conservative problems are adapted Jeon and Heinz (2015) to determine the evolution of the system, even though the second-order viscous formulations Denicol et al. (2012); Baier et al. (2008) employ equations of motion that are, in general, not in flux conservative form. In the latter, a Lagrangian method based on Smoothed Particle Hydrodynamics Monaghan (1992), extended to the relativistic regime Aguiar et al. (2001), is used in heavy-ion simulations Noronha-Hostler et al. (2013, 2014, 2016).

In the context of neutron star mergers, effective dissipation is crucial to model the correct equilibrium of accretion disks Shakura and Sunyaev (1973) and capture potentially out-of-(weak)-equilibrium effects that could affect the gravitational wave emission Alford et al. (2018); Most et al. (2021) or late time evolution of the remnant Radice (2017); Fujibayashi et al. (2018); Shibata et al. (2017); Fujibayashi et al. (2020). Moreover, turbulent dynamo processes in the merger Price and Rosswog (2006); Kiuchi et al. (2015); Aguilera-Miret et al. (2020); Skoutnev et al. (2021) critically depend on the presence of a resistive and a viscous scale. In the case of Rayleigh-Taylor instabilities potentially present at the surface of the merging neutron stars, the amplification and growth scale might depend critically on the shear viscosity Skoutnev et al. (2021).

Several attempts have been made to incorporate viscosity into astrophysical fluid dynamics simulations of relativistic systems. In the context of black hole accretion and neutron star merger remnants, some have considered simple (and acausal Hiscock and Lindblom (1985)) Navier-Stokes approaches, e.g. Ref. Duez et al. (2004); Fragile et al. (2018), whereas more recently Israel-Stewart-like formulations have been considered in Ref. Shibata et al. (2017); Chabanov et al. (2021). Apart from second-order formulations, recent progress in causal and hyperbolic first-order theories Bemfica et al. (2018); Kovtun (2019); Bemfica et al. (2019b); Hoult and Kovtun (2020); Bemfica et al. (2020) has also enabled the first simulations of such theories in the context of conformal relativistic fluids Pandya and Pretorius (2021). Another conceptually very different approach with symmetric hyperbolic structure originating from the study of continuum mechanism was presented by Refs. Peshkov et al. (2019); Romenski et al. (2020). In the context of mimicking effective shear viscous dissipation associated with small-scale magnetic turbulence, large-eddy closures have been proposed Radice (2017, 2020), although additional effort is required to ensure covariance of such a closure Duez et al. (2020). In the same spirit, an effective Israel-Stewart shear viscous formulation was proposed in Shibata et al. (2017) and successfully applied to long-term studies of post-merger remnants Fujibayashi et al. (2018).

In this work we take a different approach in that we consider dissipative evolution equations derived from a moment expansion of the relativistic Boltzmann equation Denicol et al. (2012). Going beyond all works above, we include all first-order dissipative effects, such as heat conduction, bulk and shear viscosity, and also include the anisotropies induced by magnetic fields in the equations of motion of the dissipative quantities in a fully covariant way. The only effect omitted here is resistivity, which we will consider in a forthcoming study. To this end, we incorporate the gyrofrequency as a free parameter, which allows us to control the degree of anisotropy of heat conduction and shear stresses relative to the direction of the (comoving) magnetic field. Critically, Ref. Denicol et al. (2018) has shown that in a systematic expansion in inverse Reynolds and Knudsen numbers, the associated magnetic field coupling considered in this work is the only possible term that appears at first and second order in a gradient expansion, if resistive effects of the plasma are neglected (see Denicol et al. (2019) for the resistive case). For extreme magnetic field strengths relative to the plasma energy, which are not considered here, additional terms might have to be added at second-order, though Panda et al. (2021).

Additionally, one issue affecting all previous numerical (Eulerian) implementations of Israel-Stewart formulations is the appearance of source terms with time derivatives. While the presence of strong shocks and fluid gradients in turbulent astrophysical scenarios calls for high-resolution shock capturing (HRSC) schemes, these derivatives are not commonly treated in this way, see e.g. Del Zanna et al. (2013). This can easily lead to numerical instabilities in the flow in more complex simulations, if the dissipative terms develop gradients themselves, which is associated with regions of numerically under-resolved physical viscosity. In this work, we address this point in full, by carefully designing an HRSC scheme, where all first-order dissipation terms are treated in flux-divergence form Toro (2013). This is facilitated by the use of stiff relaxation terms that are treated with implicit time integration schemes Pareschi and Russo (2005). The use of such stiff relaxation terms is inspired by the implementation of the force-free electrodynamics limit in numerical relativity simulations Alic et al. (2012); Palenzuela (2013); Most and Philippov (2020). Overall, this makes our scheme suitable for applications in heavy-ion collisions and astrophysical fluid dynamics alike.

This paper is structured as follows. In Sec. II we describe the equations of motion we employ and in Sec. III we provide a detailed derivation of how to recast them in flux-divergence form. In Sec. IV, we describe the numerical method used to solve these equations, and in particular how we handle stiff source terms. In Sec. V we demonstrate the ability of this scheme to successfully model relevant test problems, and provide a discussion of the results in Sec. VI. Throughout this paper, we adopt geometrical units where G=c=kB=1G=c=k_{B}=1 and a mostly plus signature for the Lorentzian spacetime metric gμνg_{\mu\nu}.

II Non-Resistive Relativistic Dissipative Magnetohydrodynamics

A relativistic ideal fluid can be described via its rest-mass (“baryon”) density ρ\rho, energy density ee, and four-velocity uμu^{\mu} (normalized such that uμuμ=1u_{\mu}u^{\mu}=-1). The equilibrium pressure PP is defined in terms of the other thermodynamic variables such that, for instance, P=P(ρ,e)P=P(\rho,e). This defines the equation of state, with which one may compute the temperature T=T(ρ,e)T=T(\rho,e) using standard thermodynamic relations. The equilibrium hydrodynamics system can be described by means of an energy-momentum tensor

Thydroμν=euμuν+PΔμν,\displaystyle T^{\mu\nu}_{\rm hydro}=eu^{\mu}u^{\nu}+P\Delta^{\mu\nu}, (1)

where

Δμν=gμν+uμuν,\displaystyle\Delta_{\mu\nu}=g_{\mu\nu}+u_{\mu}u_{\nu}, (2)

is the rank-2 projector orthogonal to the flow velocity, i.e. Δμνuμ=0\Delta^{\mu\nu}u_{\mu}=0. Conservation of baryon number is defined in terms of the rest-mass current Nμ=ρuμN^{\mu}=\rho u^{\mu} and baryon number conservation

μNμ=0.\nabla_{\mu}N^{\mu}=0. (3)

In this work we will also consider the effects of a (co-moving) magnetic field bμb^{\mu} tightly coupled to the fluid, i.e. in the limit of infinite electric conductivity Denicol et al. (2018). We note that bμuμ=0b^{\mu}u_{\mu}=0 so that, in the local rest frame of the fluid where uμ=(1,0,0,0)u^{\mu}=(1,0,0,0), the magnetic field 4-vector only has nonzero spatial components.

The electromagnetic field is described by the electromagnetic field strength tensor and its dual

Fμν:=b2bμν=εμναβuαbβ,\displaystyle F^{\mu\nu}:=\sqrt{b^{2}}b^{\mu\nu}=-\varepsilon^{\mu\nu\alpha\beta}u_{\alpha}b_{\beta}, (4)
Fμν=uμbνbμuν.\displaystyle{{}^{\ast}}F^{\mu\nu}=u^{\mu}b^{\nu}-b^{\mu}u^{\nu}. (5)

We stress again that these expressions only hold in the limit infinite electric conductivity, in which the comoving electric field eμ=0e^{\mu}=0 vanishes. For later convenience, we have also introduced above the shorthand notation bμνb^{\mu\nu}. The evolution of the magnetic field is then governed by the Maxwell equation μFμν=0\nabla_{\mu}{{}^{\ast}}F^{\mu\nu}=0, which may be written as

μ(bμuνbνuμ)=0.\displaystyle\nabla_{\mu}\left(b^{\mu}u^{\nu}-b^{\nu}u^{\mu}\right)=0. (6)

It is useful to introduce the projector

Ξμν=Δμν1b2bμbν,\displaystyle\Xi^{\mu\nu}=\Delta^{\mu\nu}-\frac{1}{b^{2}}b^{\mu}b^{\nu}, (7)

which is orthogonal to uμu_{\mu} and bμb^{\mu}. Using this projector, it can be shown that

FμαΞαν=FμαΔαν=Fμν,\displaystyle F^{\mu\alpha}\Xi^{\nu}_{\alpha}=F^{\mu\alpha}\Delta^{\nu}_{\alpha}=F^{\mu\nu}, (8)

and

FμαFνα=b2Ξνμ.\displaystyle F^{\mu\alpha}F_{\nu\alpha}=b^{2}\Xi^{\mu}_{\nu}. (9)

The electromagnetic fields give rise to an energy-momentum tensor given by Duez et al. (2005)

TEMμν=12b2uμuν+12b2Δμνbμbν.\displaystyle T^{\mu\nu}_{\rm EM}=\frac{1}{2}b^{2}u^{\mu}u^{\nu}+\frac{1}{2}b^{2}\Delta^{\mu\nu}-b^{\mu}b^{\nu}. (10)

In total, the generalized energy-momentum tensor for non-resistive ideal magnetohydrodynamics can be written as

Tμν=Thydroμν+TEMμν.\displaystyle T^{\mu\nu}=T^{\mu\nu}_{\rm hydro}+T^{\mu\nu}_{\rm EM}. (11)

The equations of motion of ideal relativistic magnetohydroynamics stem from the conservation of energy and momentum

μTμν=0,\nabla_{\mu}T^{\mu\nu}=0, (12)

coupled to Eq. (6).

II.1 Second-order hydrodynamic equations

In the following we summarize a formalism for describing out-of-equilibrium dynamics allowing for viscous corrections to the system. These can be encoded by the viscous stress tensor Πμν\Pi_{\mu\nu}, which is added to the fluid’s energy-momentum tensor. In this work, we use the so-called Eckart hydrodynamic frame Eckart (1940) in which the rest-mass current remains Nμ=ρuμN^{\mu}=\rho u^{\mu}, all viscous/dissipative effects appear only in the expression for the energy-momentum tensor, and there are no out-of-equilibrium corrections to the energy density.

Considering the number of degrees of freedom, we can see that TμνT^{\mu\nu} and NμN^{\mu} have a total of 14 (since TμνT^{\mu\nu} is symmetric), but in equilibrium only five of them are actually independent (e.g. {e,ρ,uμ}\{e,\rho,u^{\mu}\}). Accounting for out-of-equilibrium dynamics of the system then amounts to prescribing evolution equations for the remaining 9 degrees of freedom. These can be parameterized in terms of the bulk scalar pressure Π\Pi, the symmetric anisotropic pressure tensor πμν\pi^{\mu\nu}, and the heat flux qμq^{\mu}, which obey a distinct set of constraints on the dissipative momenta and stresses, i.e.

qμuμ\displaystyle q_{\mu}u^{\mu} =0,\displaystyle=0, (13)
πμνuμ\displaystyle\pi_{\mu\nu}u^{\mu} =0,\displaystyle=0, (14)
πμμ\displaystyle\pi^{\mu}_{\mu} =0.\displaystyle=0. (15)

Therefore, once these constraints are fulfilled, the set {πμν,Π,qμ}\{\pi_{\mu\nu},\Pi,q_{\mu}\} indeed only has 9 independent degrees of freedom. As we show in this paper, the way these constraints are imposed heavily influences the numerical methods required to study this system. Overall, we can group these contributions into the viscous stress tensor

Πμν=qμuν+qνuμ+ΠΔμν+πμν,\displaystyle\Pi_{\mu\nu}=q_{\mu}u_{\nu}+q_{\nu}u_{\mu}+\Pi\Delta_{\mu\nu}+\pi_{\mu\nu}, (16)

which is then added to the hydrodynamics equations

TμνTμν+Πμν.\displaystyle T^{\mu\nu}\rightarrow T^{\mu\nu}+\Pi^{\mu\nu}. (17)

The combined stress-energy tensor, including the dissipative contributions, is covariantly conserved

μ(Tμν+Πμν)=0,\displaystyle\nabla_{\mu}\left(T^{\mu\nu}+\Pi^{\mu\nu}\right)=0, (18)

whereas NμN^{\mu} remains unchanged with dynamics given by (3). We point out that we choose the Eckart frame in this work solely for numerical convenience in the context of astrophysical applications. Clearly, other definitions of the hydrodynamic fields, i.e. other hydrodynamic frames Israel and Stewart (1979), could have been used. For instance, one could have removed the heat flux and introduced a diffusive correction to the baryon current. Alternatively, one may employ other frames in which both terms are present. For a deeper discussion about hydrodynamic frames, and how a judicious choice of hydrodynamic variables can be useful to formulate causal, stable and strongly-hyperbolic theories of viscous relativistic fluids at first-order in derivatives, see Bemfica et al. (2018); Kovtun (2019); Bemfica et al. (2019b); Hoult and Kovtun (2020); Bemfica et al. (2020). Second-order Israel-Stewart-like theories in general hydrodynamic frames have also been investigated in Tsumura and Kunihiro (2010), and more recently in Noronha et al. (2021); Rocha and Denicol (2021).

Having parameterized the viscous degrees of freedom, we need to prescribe a set of causal evolution equations for them. In this work, we use equations of motion obtained from a truncation of suitably defined moments of the relativistic Boltzmann-Vlasov equation for a gas of charged particles (without dipole moment or spin) dynamically coupled to a magnetic field. This approach to obtain the second-order hydrodynamic equations employs a systematic expansion of the Boltzmann equation in powers of the Knudsen and inverse Reynolds numbers111The Knudsen numbers measure the ratio between microscopic length scales and the macroscopic scales associated with gradients of the conserved quantities. In the approach of Denicol et al. (2012), one denotes quantities such as Π/(e+P)\Pi/(e+P) as inverse Reynolds numbers, given that they can be seen as relativistic generalizations of the ratio between viscous and inertial forces., and this method was originally applied in Denicol et al. (2012) in the investigation of neutral kinetic systems. The equations of motion, and the transport coefficients obtained within this formulation, are widely used in current heavy-ion collision simulations of the quark-gluon plasma (usually done in the absence of eletromagnetic fields). This formalism was extended to include non-resistive, comoving magnetic field effects using the Boltzmann-Vlasov equation in Denicol et al. (2018). The magnetohydrodynamical equations of motion were obtained in the so-called 14-moment approximation Israel and Stewart (1979) and they essentially provide a generalization of Israel-Stewart fluid dynamics to the case of non-vanishing magnetic fields. The full resistive case (which includes effects from comoving electric and magnetic fields) was worked out later in this approach in Ref. Denicol et al. (2019).

In our work, the dissipative variables obey the following advection-type relaxation equations222We note that the original derivation of the non-resistive equations of motion in Denicol et al. (2018) used the Landau frame Landau and Lifshitz (1987). The Eckart frame version of these equations can be found by taking the non-resistive limit of the equations derived in Denicol et al. (2019).

τΠuμμΠ+δΠΠΠμuμ+Π=ζμuμ\displaystyle\tau_{\Pi}u^{\mu}\nabla_{\mu}\Pi+\delta_{\Pi\Pi}\Pi\,\nabla_{\mu}u^{\mu}+\Pi=-\zeta\nabla_{\mu}u^{\mu} (19)
τqΔναuμμqν+δqqqαμuμ+qα=κTuμμuακΔαμμTδqBFαμqμ\displaystyle\tau_{q}\Delta^{\alpha}_{\nu}u^{\mu}\nabla_{\mu}q^{\nu}+\delta_{qq}q^{\alpha}\nabla_{\mu}u^{\mu}+q^{\alpha}=-\kappa Tu^{\mu}\nabla_{\mu}u^{\alpha}-\kappa\Delta^{\alpha\mu}\nabla_{\mu}T-\delta_{qB}F^{\alpha\mu}q_{\mu} (20)
τπΔλναβuμμπλν+δπππαβμuμ+παβ=ηΔλναβ[νuλ+λuν]δπBFδμπμγΔγδαβ,\displaystyle\tau_{\pi}\Delta^{\alpha\beta}_{\lambda\nu}u^{\mu}\nabla_{\mu}\pi^{\lambda\nu}+\delta_{\pi\pi}\pi^{\alpha\beta}\,\nabla_{\mu}u^{\mu}+\pi^{\alpha\beta}=-\eta\Delta^{\alpha\beta}_{\lambda\nu}\left[\nabla^{\nu}u^{\lambda}+\nabla^{\lambda}u^{\nu}\right]-\delta_{\pi B}F^{\delta\mu}\pi^{\gamma}_{\mu}\,\Delta^{\alpha\beta}_{\gamma\delta}, (21)

where we have introduced the symmetric trace-free projector

Δλναβ=12[ΔλαΔνβ+ΔναΔλβ]13ΔαβΔλν.\displaystyle\Delta^{\alpha\beta}_{\lambda\nu}=\frac{1}{2}\left[\Delta^{\alpha}_{\lambda}\Delta^{\beta}_{\nu}+\Delta^{\alpha}_{\nu}\Delta^{\beta}_{\lambda}\right]-\frac{1}{3}\Delta^{\alpha\beta}\Delta_{\lambda\nu}. (22)

Above, η\eta is the shear viscosity, ζ\zeta is the bulk viscosity, κ\kappa is the thermal conductivity, τΠ\tau_{\Pi} is the bulk relaxation time, τq\tau_{q} is the relaxation time for heat flux, while τπ\tau_{\pi} is the shear viscosity relaxation time. Besides those coefficients, the equations of motion also contain the additional second-order transport coefficients {δΠΠ,δqq,δππ}\{\delta_{\Pi\Pi},\delta_{qq},\delta_{\pi\pi}\} and the new coefficients that determine the coupling between the dissipative quantities and the electromagnetic field tensor, δqB\delta_{qB} and δπB\delta_{\pi B}. Expressions for all of these coefficients, including δqB\delta_{qB} and δπB\delta_{\pi B}, in terms of the temperature and chemical potential for an ultrarelativistic gas can be found in Denicol et al. (2012, 2019, 2018). In this work, we treat the ratios δΠΠ/τΠ\delta_{\Pi\Pi}/\tau_{\Pi},ζ/τΠ\zeta/\tau_{\Pi}, δqq/τq\delta_{qq}/\tau_{q}, κ/τq\kappa/\tau_{q}, δππ/τπ\delta_{\pi\pi}/\tau_{\pi}, η/τπ\eta/\tau_{\pi}, δπB/τπ\delta_{\pi B}/\tau_{\pi}, δqB/τq\delta_{qB}/\tau_{q} as free parameters defined in terms of thermodynamic quantities (ee and ρ\rho). It is important to remark that in the formulation of Ref. Denicol et al. (2018) the form of the equations of motion remains very close to that found in standard Israel-Stewart theory, with additional terms that couple the fluid variables directly to the magnetic field. Also, we note that in this first work we have not included all the possible second-order terms that appear in Denicol et al. (2012, 2018, 2019). A systematic investigation of their effects will be given in a forthcoming study.

We note that any second-order theory with relaxation equations for the dissipative currents given by (19), (20), and (21) has an asymptotic relativistic Navier-Stokes regime. This can be seen by performing a systematic expansion in gradients, together with the conservation laws Baier et al. (2008). In the absence of magnetic fields, to first order in gradients the system reduces to

Π\displaystyle\Pi =ζμuμ+𝒪(2),\displaystyle=-\zeta\nabla_{\mu}u^{\mu}+\mathcal{O}(\partial^{2}), (23)
qα\displaystyle q^{\alpha} =κTuμμuακΔαμμT+𝒪(2),\displaystyle=-\kappa Tu^{\mu}\nabla_{\mu}u^{\alpha}-\kappa\Delta^{\alpha\mu}\nabla_{\mu}T+\mathcal{O}(\partial^{2}), (24)
παβ\displaystyle\pi^{\alpha\beta} =ηΔλναβ[νuλ+λuν]+𝒪(2),\displaystyle=-\eta\Delta^{\alpha\beta}_{\lambda\nu}\left[\nabla^{\nu}u^{\lambda}+\nabla^{\lambda}u^{\nu}\right]+\mathcal{O}(\partial^{2}), (25)

which is the standard relativistic generalization of Navier-Stokes theory Eckart (1940). It is important to understand that this limit can never be reached exactly, since it reduces the character of the hydrodynamic evolution equations from hyperbolic to parabolic, rendering the system acausal Pichon (1965). Also, we note that the standard constraints that stem from a linearized analysis of causality and stability, derived in the absence of a magnetic field in Hiscock and Lindblom (1983), automatically hold for the system of equations we use. For a linearized study of stability and causality in the presence of a nonzero magnetic field, see Biswas et al. (2020).

III Flux-conservative formulation

In the following, we will reformulate the equations for the bulk pressure (19), heat flux (20), and shear-stress tensor (21) in a flux-conservative form Toro (2013); Rezzolla and Zanotti (2013) that is suitable for numerical simulations. In particular, we will suitably extend the system in a way that allows us to recast all first-order gradient terms into flux-divergence form, i.e. μμ\nabla_{\mu}\mathcal{F}^{\mu}, for a suitable flux function μ\mathcal{F}^{\mu}. To this end, to better illustrate our approach, we consider all viscous effects separately.

III.1 Bulk pressure

We start out by considering the Israel-Stewart-like equation (19) for the bulk pressure Π\Pi. After a simple algebraic manipulation it can be shown that Eq. (19) is equivalent to

μ[(Π+ζτΠ)uμ]=\displaystyle\nabla_{\mu}\left[\left(\Pi+\frac{\zeta}{\tau_{\Pi}}\right)u^{\mu}\right]= 1τΠΠ\displaystyle-\frac{1}{\tau_{\Pi}}\Pi
+(1δΠΠτΠ)Πθ+uμμζτΠ.\displaystyle+\left(1-\frac{\delta_{\Pi\Pi}}{{\tau_{\Pi}}}\right)\Pi\theta+u^{\mu}\nabla_{\mu}\frac{\zeta}{\tau_{\Pi}}. (26)

Here we have introduced the shorthand θ=μuμ\theta=\nabla_{\mu}u^{\mu}. We will provide a detailed treatment of arbitrary viscous relaxation terms in Sec. III.5. We note that depending on the application this equation can be further simplified.

In the simplest case of ζ/τΠ= const\zeta/\tau_{\Pi}=\text{ const}, or if ζ/τΠ\zeta/\tau_{\Pi} is advected, i.e. uμμ(ζ/τΠ)=0u^{\mu}\nabla_{\mu}\left(\zeta/\tau_{\Pi}\right)=0 and δΠΠ=τΠ\delta_{\Pi\Pi}=\tau_{\Pi}, we find

μ[(Π+ζτΠ)uμ]=1τΠΠ,\displaystyle\nabla_{\mu}\left[\left(\Pi+\frac{\zeta}{\tau_{\Pi}}\right)u^{\mu}\right]=-\frac{1}{\tau_{\Pi}}\Pi, (27)

which, remarkably, is free of derivatives on the right-hand side (RHS).

III.1.1 Limit of vanishing relaxation time

We now consider a particular limit of these equations, showing that in the limit of vanishing relaxation time τΠ0\tau_{\Pi}\to 0 the system does not always approach to a perfect fluid solution. Indeed, we can see from Eq. (27) that for a choice of transport coefficient ζ=τΠf(ρ,T,)\zeta=\tau_{\Pi}\,f\left(\rho,T,\dots\dots\right), with uμμf=0u^{\mu}\nabla_{\mu}f=0, the limit τΠ0\tau_{\Pi}\rightarrow 0 corresponds to the introduction of a newly conserved quantity, i.e.

μ[(Π+ζτΠ)uμ]0,\displaystyle\nabla_{\mu}\left[\left(\Pi+\frac{\zeta}{\tau_{\Pi}}\right)u^{\mu}\right]\simeq 0, (28)

where conservation is to be understood to hold approximately up to second-order corrections. Physically, this limit is reached when bulk viscous damping happens on viscous scales viscdyn\ell_{\rm visc}\gg\ell_{\rm dyn} much larger than the dynamical scale dyn\ell_{\rm dyn}. We can further see that in this limit, Π+ζ/τΠ\Pi+\zeta/\tau_{\Pi} satisfies a continuity equations, akin to the baryon density ρ\rho. In the absence of effective viscous damping, this implies that Π\Pi provides an effective correction to the equation of state, which now depends on the bulk viscous scalar PP+ΠP\rightarrow P+\Pi, and in turn on the velocity via the continuity equation. This is similar to the use of microphysical equations of state that depend on compositional information, given by an advected scalar, such as the electron fraction YeY_{e} Oertel et al. (2017).

III.2 Heat conduction

Next we turn to the equation for heat conduction. It can trivially be shown that

Δνα[τquμμqν+δqqqνθ]=\displaystyle\Delta^{\alpha}_{\nu}\left[\tau_{q}u^{\mu}\nabla_{\mu}q^{\nu}+\delta_{qq}q^{\nu}\,\theta\right]= Δνα[κTuμμuν\displaystyle\Delta^{\alpha}_{\nu}\left[-\kappa Tu^{\mu}\nabla_{\mu}u^{\nu}\right.
κνTδqBFνμqμ]qα,\displaystyle\left.-\kappa\nabla^{\nu}T-\delta_{qB}F^{\nu\mu}q_{\mu}\right]-q^{\alpha}, (29)

where we have used that uαμuα=0u_{\alpha}\nabla_{\mu}u^{\alpha}=0. The specific form of this equation further allows us to add terms proportional to uνu^{\nu}, since these get projected out by the global Δνα\Delta^{\alpha}_{\nu} projection. Along those lines, it will help us to simplify the equations if we add such a terms in the following way

Δνα[τquμμqν+δqqqνθ]=\displaystyle\Delta^{\alpha}_{\nu}\left[\tau_{q}u^{\mu}\nabla_{\mu}q^{\nu}+\delta_{qq}q^{\nu}\,\theta\right]= Δνα[κuμμ(Tuν)\displaystyle\Delta^{\alpha}_{\nu}\left[-\kappa u^{\mu}\nabla_{\mu}\left(Tu^{\nu}\right)\right.
κνTδqBFνμqμ\displaystyle-\kappa\nabla^{\nu}T-\delta_{qB}F^{\nu\mu}q_{\mu}
κTθuντqTuμuνμκτq]\displaystyle\left.-\kappa T\theta u^{\nu}-\tau_{q}Tu^{\mu}u^{\nu}\nabla_{\mu}\frac{\kappa}{\tau_{q}}\right]
qα.\displaystyle-q^{\alpha}. (30)

We can now understand the meaning of the projector Δνμ\Delta^{\mu}_{\nu} in Eq. (29) as follows. Contracting Eq. (29) with uαu_{\alpha}, we see that the sole purpose of the projector is to enforce Eq. (13), such that qμq_{\mu} is orthogonal to uμu^{\mu}. In practice, the presence of such a projector will lead to a variety of gradient terms, which explicitly impose the constraint (29). Such terms are typically numerically difficult to evaluate and will not only increase the error budget of any numerical solution, but they will also numerically lead to only an approximate enforcement of Eq. (13).
A better approach, which is commonly used in relativistic force-free electrodynamics Alic et al. (2012); Palenzuela (2013), is to instead include the constraint (29) via stiff relaxation Alic et al. (2012). Introducing a new relaxation time ωqτq\omega_{q}\ll\tau_{q}, we can equivalently write

τquμμqν+δqqqνθ=\displaystyle\tau_{q}u^{\mu}\nabla_{\mu}q^{\nu}+\delta_{qq}q^{\nu}\,\theta= κuμμ(Tuν)κνTκTθuν\displaystyle-\kappa u^{\mu}\nabla_{\mu}\left(Tu^{\nu}\right)-\kappa\nabla^{\nu}T-\kappa T\theta u^{\nu}
qνδqBFνμqμτqTuμuνμκτq\displaystyle-q^{\nu}-\delta_{qB}F^{\nu\mu}q_{\mu}-\tau_{q}Tu^{\mu}u^{\nu}\nabla_{\mu}\frac{\kappa}{\tau_{q}}
τqωq(qμuμ)uν.\displaystyle-\frac{\tau_{q}}{\omega_{q}}\left(q_{\mu}u^{\mu}\right)u^{\nu}. (31)

The effect of this term can best be considered in the Navier-Stokes limit (in the absence of magnetic fields), where

qμuμωqκτquμμT.\displaystyle q^{\mu}u_{\mu}\simeq\omega_{q}\frac{\kappa}{\tau_{q}}u^{\mu}\nabla_{\mu}T\,. (32)

Hence, the constraint (13) will indeed be imposed when ωq0\omega_{q}\rightarrow 0. Although Eq. (31) is more desirable from a numerical point of view than the original Israel-Stewart equation (20), we will need implicit numerical schemes in order to properly compute this term. This will be explained in detail in Sec. IV. Applying the same reordering that leads to a flux-conservative form (26) for the bulk scalar we obtain the final form

μ[qνuμ+κτqTΔμν]=\displaystyle\nabla_{\mu}\left[q^{\nu}u^{\mu}+\frac{\kappa}{\tau_{q}}T\Delta^{\mu\nu}\right]= δqBτqFνμqμ1τqqν\displaystyle-\frac{\delta_{qB}}{\tau_{q}}F^{\nu\mu}q_{\mu}-\frac{1}{\tau_{q}}q^{\nu}
+[(1δqqτq)]θqν\displaystyle+\left[\left(1-\frac{\delta_{qq}}{\tau_{q}}\right)\right]\,\theta q^{\nu}
+Tgμνμκτq1ωq(qμuμ)uν.\displaystyle+Tg^{\mu\nu}\nabla_{\mu}\frac{\kappa}{\tau_{q}}-\frac{1}{\omega_{q}}\left(q_{\mu}u^{\mu}\right)u^{\nu}. (33)

For a class of theories, where κ/τq=const\kappa/\tau_{q}=\text{const} or advected (uμμ[κ/τq]=0)\left(u^{\mu}\nabla_{\mu}\left[\kappa/\tau_{q}\right]=0\right), and δqq=τq\delta_{qq}=\tau_{q}, we find the simple flux-conservative stiff relaxation equation

μ[qνuμ+κτqTΔμν]=\displaystyle\nabla_{\mu}\left[q^{\nu}u^{\mu}+\frac{\kappa}{\tau_{q}}T\Delta^{\mu\nu}\right]= 1τqqνδqBτqFνμqμ\displaystyle-\frac{1}{\tau_{q}}q^{\nu}-\frac{\delta_{qB}}{\tau_{q}}F^{\nu\mu}q_{\mu}
1ωq(qμuμ)uν.\displaystyle-\frac{1}{\omega_{q}}\left(q_{\mu}u^{\mu}\right)u^{\nu}. (34)

Nonetheless, our numerical scheme is able to solve systems with arbitrary heat conductivities, see Sec. III.5.

III.2.1 Comparison with non-relativistic heat conduction

It is interesting to consider a simple limiting case in order to highlight the nature of heat conduction within this formulation. Restricting ourselves to flat spacetime, zero magnetic field, and assuming a vanishing 3-velocity (i.e. ui0u_{i}\approx 0), it can be shown (neglecting dissipative sources other than heat condution) that the equations reduce to

t(ρε)+iqi=0\displaystyle\partial_{t}\left(\rho\varepsilon\right)+\partial_{i}q^{i}=0 (35)
t(qi)+κτqj(ηijT)=1τqqi.\displaystyle\partial_{t}\left(q^{i}\right)+\frac{\kappa}{\tau_{q}}\partial_{j}\left(\eta^{ij}T\right)=-\frac{1}{\tau_{q}}q^{i}. (36)

For simplicity, adopting the ideal gas law p=ρε(Γ1)=ρkBmbTp=\rho\varepsilon\left(\Gamma-1\right)=\rho\frac{k_{B}}{m_{b}}T, where kBk_{B} is the Boltzmann constant , mbm_{b} the baryon mass and Γ\Gamma the adiabatic coefficient, we find

tT+mb(Γ1)kBρiqi=0.\displaystyle\partial_{t}T+\frac{m_{b}\left(\Gamma-1\right)}{k_{B}\rho}\partial_{i}q^{i}=0. (37)

In this simple example, we have used that tρ=0\partial_{t}\rho=0 for vanishing fluid velocity. In the Navier-Stokes limit where qiκiTq^{i}\rightarrow-\kappa\partial^{i}T, we find that this expression reduces to the standard heat equation. However, because τq>0\tau_{q}>0 this limit will not be reached and instead we find that the temperature evolution obeys

t2Tκmb(Γ1)kBρτqΔT1τqtT=0.\displaystyle\partial_{t}^{2}T-\frac{\kappa m_{b}\left(\Gamma-1\right)}{k_{B}\rho\tau_{q}}\Delta T-\frac{1}{\tau_{q}}\partial_{t}T=0. (38)

This is a damped wave equation for the temperature TT with wave speed

cq2=κmb(Γ1)kBρτq<1,\displaystyle c_{q}^{2}=\frac{\kappa m_{b}\left(\Gamma-1\right)}{k_{B}\rho\tau_{q}}<1, (39)

where causality places a strict bound on κ/τq\kappa/\tau_{q}. This is nothing but the Telegrapher’s equation, which has already been extensively studied in the context of hyperbolic heat conduction (see, e.g. Romatschke (2010a)).

III.3 Shear-viscous stresses

We finally consider the evolution of the shear stress tensor παβ\pi^{\alpha\beta}, with the aim of recasting it into flux-conservative form. Starting from Eq. (21), we write

Δλναβ[τπuμμπλν+2η(αuβ)]=\displaystyle\Delta^{\alpha\beta}_{\lambda\nu}\left[\tau_{\pi}u^{\mu}\nabla_{\mu}\pi^{\lambda\nu}+2\eta\nabla^{\left(\alpha\right.}u^{\left.\beta\right)}\right]= δπππαβθπαβ\displaystyle-\delta_{\pi\pi}\pi^{\alpha\beta}\,\theta-\pi^{\alpha\beta}
δπBFδμπμγΔγδαβ,\displaystyle-\delta_{\pi B}F^{\delta\mu}\pi^{\gamma}_{\mu}\,\Delta^{\alpha\beta}_{\gamma\delta}\,, (40)

where 2(αuβ)=αuβ+βuα2\nabla_{\left(\alpha\right.}u_{\left.\beta\right)}=\nabla_{\alpha}u_{\beta}+\nabla_{\beta}u_{\alpha}. Similar to the observations made for Eq. (29), we find that the introduction of the trace-free projector Δλναβ\Delta^{\alpha\beta}_{\lambda\nu} was done to ensure the validity of the constraints (14) and (15). In the same spirit of deriving Eq. (31), we can replace the projector by the introduction of a stiff relaxation current. More specifically, we write

τπuμμπαβ+2η(αuβ)=\displaystyle\tau_{\pi}u^{\mu}\nabla_{\mu}\pi^{\alpha\beta}+2\eta\nabla^{\left(\alpha\right.}u^{\left.\beta\right)}= δπππαβθπαβ\displaystyle-\delta_{\pi\pi}\pi^{\alpha\beta}\,\theta-\pi^{\alpha\beta}
δπBFδμπμγΔγδαβ\displaystyle-\delta_{\pi B}F^{\delta\mu}\pi^{\gamma}_{\mu}\,\Delta^{\alpha\beta}_{\gamma\delta}
2τπωππλ(αuβ)uλ\displaystyle-2\frac{\tau_{\pi}}{\omega_{\pi}}\pi^{\lambda\left(\alpha\right.}u^{\left.\beta\right)}u_{\lambda}
τπωππ(πμνgμν)gαβ,\displaystyle-\frac{\tau_{\pi}}{\omega_{\pi\pi}}\left(\pi^{\mu\nu}g_{\mu\nu}\right)g^{\alpha\beta}, (41)

where we have introduced the relaxations times ωπ,ωππτπ\omega_{\pi},\omega_{\pi\pi}\ll\tau_{\pi}. Reordering the terms in the same way as for Eqs. (26) and (33), we find

μ[παβuμ+ητπ(gμαuβ+gμβuα)]=\displaystyle\nabla_{\mu}\left[\pi^{\alpha\beta}u^{\mu}+\frac{\eta}{\tau_{\pi}}\left(g^{\mu\alpha}u^{\beta}+g^{\mu\beta}u^{\alpha}\right)\right]= (1δππτπ)παβθ1τππαβ+(gμαuβ+gμβuα)μητπδπBτπFδμπμγΔγδαβ\displaystyle\left(1-\frac{\delta_{\pi\pi}}{\tau_{\pi}}\right)\pi^{\alpha\beta}\,\theta-\frac{1}{\tau_{\pi}}\pi^{\alpha\beta}+\left(g^{\mu\alpha}u^{\beta}+g^{\mu\beta}u^{\alpha}\right)\nabla_{\mu}\frac{\eta}{\tau_{\pi}}-\frac{\delta_{\pi B}}{\tau_{\pi}}F^{\delta\mu}\pi^{\gamma}_{\mu}\,\Delta^{\alpha\beta}_{\gamma\delta}
1ωπ[(παλuλ)uβ+(πβλuλ)uα]1ωππ(πμνgμν)gαβ.\displaystyle-\frac{1}{\omega_{\pi}}\left[\left(\pi^{\alpha\lambda}u_{\lambda}\right)u^{\beta}+\left(\pi^{\beta\lambda}u_{\lambda}\right)u^{\alpha}\right]-\frac{1}{\omega_{\pi\pi}}\left(\pi^{\mu\nu}g_{\mu\nu}\right)g^{\alpha\beta}. (42)

We further note that all terms terms proportional to uμu^{\mu} are already accounted for in the constraint damping term and will get removed in the ω0\omega\rightarrow 0 limit. Remarkably, and differently from the heat conduction and bulk viscosity, this leads to the removal of the μ(η/τπ)\nabla_{\mu}\left(\eta/\tau_{\pi}\right) term from the equations. Hence, no approximation or other special treatment need to be applied to handle non-constant shear viscosity. The final form of the evolution equation then reads

μ[παβuμ+ητπ(gμαuβ+gμβuα)]=\displaystyle\nabla_{\mu}\left[\pi^{\alpha\beta}u^{\mu}+\frac{\eta}{\tau_{\pi}}\left(g^{\mu\alpha}u^{\beta}+g^{\mu\beta}u^{\alpha}\right)\right]= (1δππτπ)παβθδπBτπFδμπμγΔγδαβ\displaystyle\left(1-\frac{\delta_{\pi\pi}}{\tau_{\pi}}\right)\pi^{\alpha\beta}\,\theta-\frac{\delta_{\pi B}}{\tau_{\pi}}F^{\delta\mu}\pi^{\gamma}_{\mu}\,\Delta^{\alpha\beta}_{\gamma\delta}
1τππαβ1ωπ[(παλuλ)uβ+(πβλuλ)uα]1ωππ(πμνgμν)gαβ.\displaystyle-\frac{1}{\tau_{\pi}}\pi^{\alpha\beta}-\frac{1}{\omega_{\pi}}\left[\left(\pi^{\alpha\lambda}u_{\lambda}\right)u^{\beta}+\left(\pi^{\beta\lambda}u_{\lambda}\right)u^{\alpha}\right]-\frac{1}{\omega_{\pi\pi}}\left(\pi^{\mu\nu}g_{\mu\nu}\right)g^{\alpha\beta}. (43)

As a final remark we stress the ambiguity in adding or removing constraint terms related to Eq. (14) and (15). While in the continuum limit Eq. (15) will perfectly hold, the discrete version of these equations might behave differently whether or not additional terms proportional to gμνg^{\mu\nu} would be added. One particular example would be the removal of the trace from the shear tensor in Eq. (43). This would lead to a replacement of the principal part of Eq. (43) with

μ[παβuμ+ητπ(gμαuβ+gμβuα12gαβuμ)].\displaystyle\nabla_{\mu}\left[\pi^{\alpha\beta}u^{\mu}+\frac{\eta}{\tau_{\pi}}\left(g^{\mu\alpha}u^{\beta}+g^{\mu\beta}u^{\alpha}-\frac{1}{2}g^{\alpha\beta}u^{\mu}\right)\right]. (44)

Whether or not such modifications will impact the stability of the system will require a more in-depth analysis, which will be provided in forthcoming work.

III.4 Interpretation of the magnetic field coupling

In the following, we would like to associate a meaning with the magnetic field coupling terms, δqB\delta_{qB} and δπB\delta_{\pi B}. To facilitate this, we will draw on an analogy with the Ohm’s law for a single fluid ideal magnetohydrodynamics in the Newtonian regime. To this end, we consider a simple Ohm’s law from Hall magnetohydrodynamics Sturrock and Andrew (1994),

Ji+ωgτeεijkJjBk=σEi,\displaystyle J^{i}+\omega_{g}\tau_{e}\varepsilon^{ijk}J_{j}B_{k}=\sigma E^{i}, (45)

where ωg=q/(mecB)\omega_{g}=q/\left(m_{e}cB\right) is the electron gyrofrequency, qq the electron charge, mem_{e} the electron mass and JiJ^{i} the electric current. By fixing the gyration velocity vv_{\perp}, we can also express this via the inverse Larmor radius RL1=ωg/vR^{-1}_{L}=\omega_{g}/v_{\perp}.
Taking the Newtonian limit of Eq. (20), i.e. τq0\tau_{q}\rightarrow 0, and going to the rest-frame of the fluid uμ=(1,0,0,0)u^{\mu}=\left(1,0,0,0\right), we obtain

qi+δqBεijkqjbk=κiT.\displaystyle q^{i}+\delta_{qB}\varepsilon^{ijk}q_{j}b_{k}=-\kappa\partial^{i}T. (46)

Since we only consider non-resistive plamas, we can see that the heat conduction current replaces the electric current compared to Eq. (45). In fact, if we were to consider resistive plasmas Denicol et al. (2019), we would obtain exactly the same expression with σei\sigma^{\prime}e^{i} on the RHS, where eμe^{\mu} is the comoving electric field, and σ\sigma^{\prime} the associated conductivity.

Although Eq. (45) describes an electric and Eq. (46) a heat current, it can be shown that the two are related in a dissipative relativistic fluid description. Indeed, the total heat flux of the theory is given by Denicol et al. (2019)

Qμ=qμ+e+PρmbVfμ,\displaystyle Q^{\mu}=q^{\mu}+\frac{e+P}{\rho}m_{b}V_{f}^{\mu}, (47)

where mbm_{b} is the baryon mass and VfμV_{f}^{\mu} is the particle diffusion current, which directly enters the electric current in Maxwell’s equations Denicol et al. (2019). In defining the dissipative system in Eq. (16), we have made a frame choice, to neglect the particle diffusion current, and treat the heat flux vector qμq^{\mu} directly. We could have additionally chosen to operate in the Landau frame, and instead reexpressed the system in terms of the diffusion current VfμV_{f}^{\mu}, being more similar to Eq. (45).

Overall, this comparison allows us to identify the coupling scale with a thermal gyrofrequency

δqBωgτq.\displaystyle\delta_{qB}\sim\omega_{g}\tau_{q}. (48)

We further find it natural to identify the time scale of collisions τe\tau_{e} with the relaxation time τq\tau_{q}, as those should be proportional to each other. Physically this implies that the thermal conductivity (in the Navier-Stokes limit) will split into a part parallel (κ)\left(\kappa_{\parallel}\right) and perpendicular (κ\left(\kappa_{\perp}\right., κH)\left.\kappa_{\rm H}\right) to the magnetic field, where the latter term is the equivalent of the electric Hall conductivity Denicol et al. (2018). In particular, this would lead to an anisotropic modification of the Navier-Stokes limit as follows

qμ\displaystyle q^{\mu}\simeq (κΞμνκbμbνκHbμν)[uααuν+νlogT]T.\displaystyle\left(\kappa_{\perp}\Xi^{\mu\nu}-\kappa_{\parallel}b^{\mu}b^{\nu}-\kappa_{\rm H}b^{\mu\nu}\right)\left[u^{\alpha}\nabla_{\alpha}u_{\nu}+\nabla_{\nu}\log\,T\right]T. (49)

Crucially, in our current prescription these anisotropy effects are controlled by a single parameter δqB\delta_{qB}, which can be freely specified. For a more detailed discussion of this effect, see Ref. Denicol et al. (2018).

III.4.1 Braginskii-limit

Having discussed that the equations presented here are naturally able to capture the effect of (Hall) anisotropies in the heat conduction, we now turn to the strong coupling limit. It will turn out that this limit has a very important interpretation in the limit of weakly collisional plasmas.

Indeed, in the strong coupling limit δπB,δqB\delta_{\pi B},\delta_{qB}\rightarrow\infty, we find that the shear stress and heat flux must satisfy

Fνμqμ=0,\displaystyle F^{\nu\mu}q_{\mu}=0, (50)
FδμπμγΔγδαβ=0.\displaystyle F^{\delta\mu}\pi^{\gamma}_{\mu}\,\Delta^{\alpha\beta}_{\gamma\delta}=0. (51)

Since πμν\pi^{\mu\nu} and qμq^{\mu} are also subject to the constraints (13)-(15), we find that this imposes the following form on the dissipative fluxes and stresses,

qμ=q0bμ,\displaystyle q^{\mu}=q_{0}b^{\mu}, (52)
πμν=π0(bμbν+b23Δμν),\displaystyle\pi^{\mu\nu}=\pi_{0}\left(-b^{\mu}b^{\nu}+\frac{b^{2}}{3}\Delta^{\mu\nu}\right), (53)

where q0q_{0} and π0\pi_{0} are Lorentz scalars, which in the Navier-Stokes limit approach

q0κT1b2[bαuμμuα+bμμlogT],\displaystyle q_{0}\simeq-\kappa T\frac{1}{\sqrt{b^{2}}}\left[b_{\alpha}u^{\mu}\nabla_{\mu}u^{\alpha}+b^{\mu}\nabla_{\mu}\log\,T\right]\,, (54)
π03ηΞμν(μuν).\displaystyle\pi_{0}\simeq-3\eta\,\Xi^{\mu\nu}\nabla_{\left(\mu\right.}u_{\left.\nu\right)}\,. (55)

From a physical point of view, δBRL1ωgλmfp\delta_{B}\sim R_{L}^{-1}\sim\omega_{g}\sim\lambda_{\rm mfp} can be identified with the mean-free-path of the system Denicol et al. (2018) so that, as expected, in the limit of weak collisionality λmfp\lambda_{\rm mfp}\rightarrow\infty a Braginskii-like limit is recovered Braginskii (1965) for our system of equations.

A similar formulation of relativistic viscous magnetohydrodynamics using Israel-Stewart-like equations has been proposed for weakly collisional plasmas in Ref.  (Chandra et al., 2015). This formulation was modelled after non-relativistic Braginskii theory (Braginskii, 1965), where the shear stress aligns with the comoving magnetic fields. We stress that differently from the formulation of Chandra et al. (2015) where the anisotropy of the viscous stresses and heat flux has been imposed from the outset, in the framework considered here the anisotropy emerges naturally as the gyrofrequency diverges in the limit of weak collisionality plasmas. As such, the (truncated first-order) equations of Denicol et al. (2018) in this limit can be considered as a generalization of Braginskii magnetohydrodynamics for non-resistive relativistic plasmas for finite thermal gyrofrequencies.

III.5 Advection terms

Although we have so far managed to eliminate most gradient terms from the RHS of the equations of motion, a set of advection and compression terms still remain. We will now, likewise, reformulate them as a set of implicit rate equations in the comoving frame.

We note that because of baryon number conservation, an arbitrary advected scalar YY satisfies

0=uμμY=μ(ρYuμ).\displaystyle 0=u^{\mu}\nabla_{\mu}Y=\nabla_{\mu}\left(\rho Yu^{\mu}\right). (56)

If we want to enforce that YY follows a certain behavior, Y0Y_{0}, we may implicitly define an autonomous source term such that

Y(𝐱,t)=Y0(𝐱,t).\displaystyle Y\left({\bf x},t\right)=Y_{0}\left({\bf x},t\right). (57)

This gives rise to a relaxation current

Y=u0ωY(YY0),\displaystyle\mathcal{I}_{Y}=-\frac{u^{0}}{\omega_{Y}}\left(Y-Y_{0}\right), (58)

where ωY\omega_{Y} is a stiff relaxation time scale. In order to enforce the condition (57) we may add this current to the advected part,

μ(ρYuμ)=ρY.\displaystyle\nabla_{\mu}\left(\rho Yu^{\mu}\right)=\rho\mathcal{I}_{Y}. (59)

In other words, in the local comoving frame of the fluid, this equation is prescribing a rate equation to enforce the damping of YY towards Y0Y_{0} on a (subgrid) time scale ωY\omega_{Y}. If we can evaluate the stiff current Y\mathcal{I}_{Y} numerically, it can be used to replace advective derivatives of the form

uμμY=Y.\displaystyle u^{\mu}\nabla_{\mu}Y=\mathcal{I}_{Y}. (60)

In the same way, we can also treat gradient terms by noting that

δμνμZ=μ(δνμZ),\displaystyle\delta^{\nu}_{\mu}\nabla_{\mu}Z=\nabla_{\mu}\left({\delta^{\mu}_{\nu}Z}\right), (61)

where δμν\delta^{\nu}_{\mu} is the Kronecker symbol. Introducing the current ZνZ^{\nu}, we can write in non-covariant form:

t(gZν)+i(gZδνi)=gΓμνμ+gνZ,\displaystyle\partial_{t}\left(\sqrt{-g}Z_{\nu}\right)+\partial_{i}\left(\sqrt{-g}Z\delta^{i}_{\nu}\right)=\sqrt{-g}\Gamma^{\mu}_{\mu\nu}+\sqrt{-g}\mathcal{I}^{Z}_{\nu}, (62)

where Γμνλ\Gamma^{\lambda}_{\mu\nu} is the Christoffel symbol associated with gμνg_{\mu\nu} and

νZ=1ωZ(ZνZδν0).\displaystyle\mathcal{I}^{Z}_{\nu}=-\frac{1}{\omega_{Z}}\left(Z_{\nu}-Z\delta^{0}_{\nu}\right). (63)

Here ωZ\omega_{Z} is a stiff relaxation timescale.

Applied to our source terms this results in the following set of stiff advection equations

μ(ρYζuμ)=ρζ,\displaystyle\nabla_{\mu}\left(\rho Y_{\zeta}u^{\mu}\right)=\rho\mathcal{I}_{\zeta}, (64)
μ(ρYκuμ)=ρκ,\displaystyle\nabla_{\mu}\left(\rho Y_{\kappa}u^{\mu}\right)=\rho\mathcal{I}_{\kappa}, (65)
μ(ρYθuμ)=ρθ,\displaystyle\nabla_{\mu}\left(\rho Y_{\theta}u^{\mu}\right)=\rho\mathcal{I}_{\theta}, (66)
μ(ρYηuμ)=ρη,\displaystyle\nabla_{\mu}\left(\rho Y_{\eta}u^{\mu}\right)=\rho\mathcal{I}_{\eta}, (67)
t(gZνκ)+i(gκτqδνi)=gΓμνμ+gνκ.\displaystyle\partial_{t}\left(\sqrt{-g}Z^{\kappa}_{\nu}\right)+\partial_{i}\left(\sqrt{-g}\frac{\kappa}{\tau_{q}}\delta^{i}_{\nu}\right)=\sqrt{-g}\Gamma^{\mu}_{\mu\nu}+\sqrt{-g}\mathcal{I}^{\kappa}_{\nu}. (68)

The source terms of those equations are fixed according to

ζ=1ωζ(YζζτΠ),\displaystyle\mathcal{I}_{\zeta}=-\frac{1}{\omega_{\zeta}}\left(Y_{\zeta}-\frac{\zeta}{\tau_{\Pi}}\right), (69)
κ=1ωκ(Yκκτq),\displaystyle\mathcal{I}_{\kappa}=-\frac{1}{\omega_{\kappa}}\left(Y_{\kappa}-\frac{\kappa}{\tau_{q}}\right), (70)
θ=1ωθ(Yθρ),\displaystyle\mathcal{I}_{\theta}=-\frac{1}{\omega_{\theta}}\left(Y_{\theta}-\rho\right), (71)
η=1ωη(Yηητπ),\displaystyle\mathcal{I}_{\eta}=-\frac{1}{\omega_{\eta}}\left(Y_{\eta}-\frac{\eta}{\tau_{\pi}}\right), (72)
νκ=1ωκ(Zνκκτqδν0),\displaystyle\mathcal{I}^{\kappa}_{\nu}=-\frac{1}{\omega_{\kappa}}\left(Z^{\kappa}_{\nu}-\frac{\kappa}{\tau_{q}}\delta^{0}_{\nu}\right), (73)

where we have introduced new relaxation time scales ωζ,ωκ,ωη,andωθ\omega_{\zeta}\,,\omega_{\kappa}\,,\omega_{\eta}\,,\text{and}\,\omega_{\theta}.

III.6 Summary

In conclusion, our equations of motion for flux-conservative dissipative magnetohydrodynamics read

μ(Tμν+Πμν)=0,\displaystyle\nabla_{\mu}\left(T^{\mu\nu}+\Pi^{\mu\nu}\right)=0\,, (74)
μ(bμuνbνuμ)=0,\displaystyle\nabla_{\mu}\left(b^{\mu}u^{\nu}-b^{\nu}u^{\mu}\right)=0\,, (75)
μ[(Π+Yζ)uμ]=1τΠΠ(1δΠΠτΠ)Π1ρθ+ζ,\displaystyle\nabla_{\mu}\left[\left(\Pi+Y_{\zeta}\right)u^{\mu}\right]=-\frac{1}{\tau_{\Pi}}\Pi-\left(1-\frac{\delta_{\Pi\Pi}}{{\tau_{\Pi}}}\right)\Pi\frac{1}{\rho}\mathcal{I}_{\theta}+\mathcal{I}_{\zeta}, (76)
μ[qνuμ+YκTΔμν]=(1δqqτq)qν1ρθ+Tκν1τqqν1ωq(qμuμ)uνδqBτqFνμqμ,\displaystyle\nabla_{\mu}\left[q^{\nu}u^{\mu}+Y_{\kappa}T\Delta^{\mu\nu}\right]=-\left(1-\frac{\delta_{qq}}{\tau_{q}}\right)q^{\nu}\,\frac{1}{\rho}\mathcal{I}_{\theta}+T\mathcal{I}_{\kappa}^{\nu}-\frac{1}{\tau_{q}}q^{\nu}-\frac{1}{\omega_{q}}\left(q_{\mu}u^{\mu}\right)u^{\nu}-\frac{\delta_{qB}}{\tau_{q}}F^{\nu\mu}q_{\mu}, (77)
μ[παβuμ+ητπ(gμαuβ+gμβuα)]=(1δππτπ)παβ1ρθ1τππαβδπBτπFδμπμγΔ~γδαβ\displaystyle\nabla_{\mu}\left[\pi^{\alpha\beta}u^{\mu}+\frac{\eta}{\tau_{\pi}}\left(g^{\mu\alpha}u^{\beta}+g^{\mu\beta}u^{\alpha}\right)\right]=-\left(1-\frac{\delta_{\pi\pi}}{\tau_{\pi}}\right)\pi^{\alpha\beta}\,\frac{1}{\rho}\mathcal{I}_{\theta}-\frac{1}{\tau_{\pi}}\pi^{\alpha\beta}-\frac{\delta_{\pi B}}{\tau_{\pi}}F^{\delta\mu}\pi^{\gamma}_{\mu}\,\tilde{\Delta}^{\alpha\beta}_{\gamma\delta}
1ωπ[(παλuλ)uβ+(πβλuλ)uα]1ωππ(πμνgμν)gαβ,\displaystyle\phantom{\nabla_{\mu}\left[\pi^{\alpha\beta}u^{\mu}+Y_{\eta}\left(g^{\mu\alpha}u^{\beta}+g^{\mu\beta}u^{\alpha}\right)\right]=}-\frac{1}{\omega_{\pi}}\left[\left(\pi^{\alpha\lambda}u_{\lambda}\right)u^{\beta}+\left(\pi^{\beta\lambda}u_{\lambda}\right)u^{\alpha}\right]-\frac{1}{\omega_{\pi\pi}}\left(\pi^{\mu\nu}g_{\mu\nu}\right)g^{\alpha\beta}, (78)
μ(ρYζuμ)=ρζ,\displaystyle\nabla_{\mu}\left(\rho Y_{\zeta}u^{\mu}\right)=\rho\mathcal{I}_{\zeta}, (79)
μ(ρYκuμ)=ρκ,\displaystyle\nabla_{\mu}\left(\rho Y_{\kappa}u^{\mu}\right)=\rho\mathcal{I}_{\kappa}, (80)
μ(ρYθuμ)=ρθ,\displaystyle\nabla_{\mu}\left(\rho Y_{\theta}u^{\mu}\right)=\rho\mathcal{I}_{\theta}, (81)
μ(ρYηuμ)=ρη,\displaystyle\nabla_{\mu}\left(\rho Y_{\eta}u^{\mu}\right)=\rho\mathcal{I}_{\eta}, (82)
t(gZνκ)+i(gYκδνi)=gΓμνμ+gνκ.\displaystyle\partial_{t}\left(\sqrt{-g}Z^{\kappa}_{\nu}\right)+\partial_{i}\left(\sqrt{-g}Y_{\kappa}\delta^{i}_{\nu}\right)=\sqrt{-g}\Gamma^{\mu}_{\mu\nu}+\sqrt{-g}\mathcal{I}^{\kappa}_{\nu}\,. (83)

Before we proceed, a few comments are in order. All second-order gradient terms (Πθ,qνθ,πμνθ)\left(\Pi\theta\,,q^{\nu}\theta\,,\pi^{\mu\nu}\theta\right) on the RHS of all equations are treated implicitly via the relaxation equations associated with (57). Although such a treatment seems approximate, compared to the flux-divergence operator on the left-hand side (LHS), it is important to point out that we are primarily interested in near-equilibrium behavior where Π\Pi, qνq^{\nu}, and πμν\pi^{\mu\nu} are small. As such, these specific second-order terms constitute only a minor correction and could even be omitted. This is also consistent with neglecting all other second-order transport terms in these equations, which should otherwise be present Denicol et al. (2012). On the other hand, since first-order gradient terms are important for the evolution, they require more sophisticated numerical methods, as outlined in Sec. IV . Second, we point out that in deriving these equations we have made use of our freedom to remove all terms proportional to uμu^{\mu} on the RHS. Due to the projected nature of the Israel-Stewart limit, these terms would get removed when projecting the equations into the fluid frame (in the continuum limit). In the stiff-relaxation approach, all these terms are hence implicitly accounted for in the relaxation operator scaling with the fluid four-velocity uμu^{\mu}.

It would be very interesting to investigate if our formulation of the equations of motion can be proven to be hyperbolic in the full nonlinear limit. This would be especially important when coupling our system to Einstein’s equations, which is needed to investigate viscous phenomena in neutron star mergers Shibata et al. (2017); Radice (2017); Alford et al. (2018); Most et al. (2021) or black hole accretion in non-dynamical spacetimes Foucart et al. (2016, 2017). To the best of our knowledge, there is currently no formulation of general-relativistic dissipative magnetohydrodynamics that is proven to be strongly hyperbolic in the nonlinear regime. For instance, such a statement is not known to hold for the elegant approach proposed in Ref. Chandra et al. (2015). Drawing from our experience with the zero magnetic field limit of Israel-Stewart-like equations Bemfica et al. (2019a, 2021), such a nonlinear analysis of hyperbolicity and causality would give rise to important constraints on the values of the dissipative currents, the transport coefficients, and in this case also the magnetic field.

The equations presented in this section are written in a way that seems to be very natural for for performing such a nonlinear investigation. Indeed, here we have taken the first step towards this result as we have been able to rewrite the system as a set of first-order PDEs in flux-conservative form, which can be beneficial in such studies. We remark, however, that proving that our nonlinear system of equations is strongly hyperbolic is a very challenging mathematical task, which we hope to address in the near future. On the other hand, a linearized analysis of our equations can be performed following Ref. Biswas et al. (2020). The only difference would be that the inclusion of the stiff advection equations introduces additional non-hydrodynamic modes to the system parametrized by the new relaxation times ωζ\omega_{\zeta}, ωκ\omega_{\kappa}, ωη\omega_{\eta} and ωθ\omega_{\theta}. Therefore, in the light of Biswas et al. (2020), we leave the investigation of the linear regime of our equations to a future dedicated study.

IV Numerical methods

Having derived the set of Israel-Stewart-like equations in flux-conservative form with stiff relaxation, we now want to integrate them numerically to investigate their behavior in a series of test problems. One of the key features will be the treatment of the stiff source terms. We first comment on the general numerical approach before providing a detailed discussion of the implicit time stepping in Sec. IV.1.

The equations are discretized following a simple finite-volume scheme common to many fluid dynamical problems Toro (2013); Rezzolla and Zanotti (2013). More specifically, we solve a flux-conservation law of the form

t(g𝒰)+i(gi)=g𝒮,\displaystyle\partial_{t}\left(\sqrt{-g}\,\mathcal{U}\right)+\partial_{i}\left(\sqrt{-g}\,\mathcal{F}^{i}\right)=\sqrt{-g}\,\mathcal{S}, (84)

where 𝒰\mathcal{U} is a conserved state vector, i\mathcal{F}^{i} is the flux vector and 𝒮\mathcal{S} are the source terms. The conserved variables, i.e. the components of 𝒰\mathcal{U}, are given by

ρ=ρu0\displaystyle\rho_{\ast}=\rho u^{0} (85)
e=[[e+P+b2+Π](u0)2+(P+Π+b22)g00\displaystyle e_{\ast}=\left[\left[e+P+b^{2}+\Pi\right]\left(u^{0}\right)^{2}+\left(P+\Pi+\frac{b^{2}}{2}\right)\,g^{00}\right.
+2q0u0+π00b0b0)],\displaystyle\phantom{\tau=}\,\left.+2q^{0}u^{0}+\pi^{00}-b^{0}b^{0}\right)], (86)
Si=[(ρh+Π+b2)u0ui+q0ui+u0qi+πi0b0bj]\displaystyle S_{i}=\left[\left(\rho h+\Pi+b^{2}\right)u^{0}u_{i}+q^{0}u_{i}+u^{0}q_{i}+\pi^{0}_{i}-b^{0}b_{j}\right] (87)
Π~=(Π+Yζ)u0\displaystyle\tilde{\Pi}=\left(\Pi+Y_{\zeta}\right)u^{0} (88)
q~μ=(qμu0+YκTΔ0ν),\displaystyle\tilde{q}^{\mu}=\left(q^{\mu}u^{0}+Y_{\kappa}T\Delta^{0\nu}\right), (89)
π~μν=(πμνu0+Yη[gμ0uν+gν0uμ]).\displaystyle\tilde{\pi}^{\mu\nu}=\left(\pi^{\mu\nu}u^{0}+Y_{\eta}\left[g^{\mu 0}u^{\nu}+g^{\nu 0}u^{\mu}\right]\right). (90)

The precise form of the fluxes and sources depends on the spacetime employed in the simulations and can straightforwardly be derived from Eq. (76)-(83). As the nature of dissipative effects is to provide local heat fluxes and stresses, these will likely act on scales much smaller than the curvature scale of space-time. Hence, for a first exploration we will conduct our numerical experiments exclusively in flat spacetime, and leave general-relativistic test cases for a future study. We provide an explicit representation of the flux-conservative form of the equations in flat Minkowski spacetime in Appendix A.

We discretize this set of equations by adopting a second-order accurate finite-volume algorithm. In particular we evolve cell averaged volume quantities U=(ΔV)1𝒰dVU=\left(\Delta V\right)^{-1}\int\mathcal{U}{\rm d}V, where ΔV\Delta V is the cell volume over which the integral is carried out. We compute an upwinded discretization of the flux by solving the local Riemann problem at each cell interface Toro (2013). In particular, we first perform a limited interpolation step using the WENO-Z algorithm (Borges et al., 2008) for the right and left states, URU_{R} and ULU_{L}, at the interface. From those, we can compute an upwinded flux adopting the Rusanov Riemann solver Toro (2013),

F=12(FL+FR)cc2(URUL),\displaystyle F=\frac{1}{2}\left(F_{L}+F_{R}\right)-\frac{c_{c}}{2}\left(U_{R}-U_{L}\right), (91)

where ccc_{c} is the fastest characteristic speed at the interface. For simplicity, we adopt this to be the speed of light cc=cc_{c}=c. While such a choice leads to more diffusive numerical solutions, it is common in relativistic simulations of non-ideal magnetohydrodynamics, when the characteristics of the system are not known Dionysopoulou et al. (2013); Ripperda et al. (2019). The flux-update of the ii\partial_{i}\mathcal{F}^{i} term is performed using a second-order accurate discretization of the divergence operator. Especially for (barely) resolved dissipative length scales (e.g. in the presence of strong gradients) with smooth profiles it might be beneficial to used improved high-order flux-update schemes Del Zanna et al. (2007); McCorquodale and Colella (2011) as used in (Most et al., 2019; Felker and Stone, 2018). Nonetheless, we find a second-order scheme to be sufficient for the simple test problems considered here. The numerical implementation is done on top of a newly developed version of the Athena++ framework Stone et al. (2020), which utilizes the Kokkos library (Edwards et al., 2014) to achieve parallelization across modern CPU and GPU architectures.

IV.1 Implicit time-stepping

An integral part of the new formulation presented here is the enforcement of the constraints (13), (14) and (15) by means of stiff relaxation source terms. Due to the fact that there must be a hierarchy among the relaxation times involved, i.e., τω0\tau\gg\omega\rightarrow 0, this severe stiffness limit can only be handled using implicit time integration schemes. One such scheme, consistent with the explicit strong-stability preserving Runge-Kutta schemes, is the RK3-ImEx SSP3 (4,3,3) scheme Pareschi and Russo (2005). Within this scheme, we split the source terms into two distinct contributions. In particular we write

t𝒰i=i,\displaystyle\partial_{t}\mathcal{U}_{i}=\mathcal{H}_{i}, (92)
t𝒱i=i+i.\displaystyle\partial_{t}\mathcal{V}_{i}=\mathcal{E}_{i}+\mathcal{I}_{i}. (93)

Here, we have split the hydrodynamic variables 𝒰i=(e,ρ,Si)\mathcal{U}_{i}=\left(e_{\ast},\rho_{\ast},S_{i}\right) from the dissipative variables 𝒱i=(Π~,q~ν,π~μν)\mathcal{V}_{i}=\left(\tilde{\Pi},\tilde{q}^{\nu},\tilde{\pi}^{\mu\nu}\right). We have further split the RHS of the dissipative variables 𝒱i\mathcal{V}_{i} into explicit, i\mathcal{E}_{i}, and implicit, i\mathcal{I}_{i}, terms. The hydrodynamic variables remain explicit with source terms i\mathcal{H}_{i}. Explicit terms are evaluated at the current time tnt^{n}, while implicit terms are evaluated at the next time step. We will give a detailed description below.

The multistep ImEx scheme then proceeds with n=4n=4 internal stages Pareschi and Russo (2005),

𝒰i(0)=𝒰i(t)\displaystyle\mathcal{U}^{\left(0\right)}_{i}=\mathcal{U}_{i}\left(t\right) (94)
𝒰i(k)=𝒰i+Δtl=1k1akli(𝒰i(l),𝒱i(l))\displaystyle\mathcal{U}^{\left(k\right)}_{i}=\mathcal{U}_{i}+{\Delta t}\sum_{l=1}^{k-1}a_{kl}\mathcal{H}_{i}\left(\mathcal{U}_{i}^{\left(l\right)},\mathcal{V}_{i}^{\left(l\right)}\right) (95)
𝒱i(k)=𝒱i+Δtl=1k1akli(𝒰i(l),𝒱i(l))\displaystyle\mathcal{V}^{\left(k\right)}_{i}=\mathcal{V}_{i}+{\Delta t}\sum_{l=1}^{k-1}a_{kl}\mathcal{E}_{i}\left(\mathcal{U}_{i}^{\left(l\right)},\mathcal{V}_{i}^{\left(l\right)}\right)
+Δtl=1ka~kli(𝒰i(l),𝒱i(l)),\displaystyle\phantom{\mathcal{V}^{\left(k\right)}_{i}=\mathcal{V}_{i}}+{\Delta t}\sum_{l=1}^{k}\tilde{a}_{kl}\mathcal{I}_{i}\left(\mathcal{U}_{i}^{\left(l\right)},\mathcal{V}_{i}^{\left(l\right)}\right), (96)
𝒰i(t+Δt)=𝒰i+Δtl=1nbli(𝒰i(l),𝒱i(l)),\displaystyle\mathcal{U}_{i}\left(t+{\Delta t}\right)=\mathcal{U}_{i}+{\Delta t}\sum_{l=1}^{n}b_{l}\mathcal{H}_{i}\left(\mathcal{U}_{i}^{\left(l\right)},\mathcal{V}_{i}^{\left(l\right)}\right), (97)
𝒱i(t+Δt)=𝒱i+Δtl=1nbli(𝒰i(l),𝒱i(l))\displaystyle\mathcal{V}_{i}\left(t+{\Delta t}\right)=\mathcal{V}_{i}+{\Delta t}\sum_{l=1}^{n}b_{l}\mathcal{E}_{i}\left(\mathcal{U}_{i}^{\left(l\right)},\mathcal{V}_{i}^{\left(l\right)}\right)
+Δtl=1nb~li(𝒰i(l),𝒱i(l)).\displaystyle\phantom{\mathcal{V}_{i}\left(t+{\Delta t}\right)=\mathcal{V}_{i}}+{\Delta t}\sum_{l=1}^{n}\tilde{b}_{l}\mathcal{I}_{i}\left(\mathcal{U}_{i}^{\left(l\right)},\mathcal{V}_{i}^{\left(l\right)}\right). (98)

The coefficients aij,a~ij,bi,b~ia_{ij}\,,\tilde{a}_{ij}\,,b_{i}\,,\tilde{b}_{i} can be found in Table 1.

0 0 0 0 0
0 0 0 0 0
1 0 1 0 0
1/2 0 1/4 1/4 0
0 1/6 1/6 2/3
α\alpha α\alpha 0 0 0
0 -α\alpha α\alpha 0 0
1 0 1α1-\alpha α\alpha 0
1/2 β\beta η\eta γ\gamma α\alpha
0 1/6 1/6 2/3
Table 1: Butcher tableau representation of the third-order IMEX-SSP3(4,3,3) scheme (Pareschi and Russo, 2005). The explicit part aija_{ij} (left) and the implicit part a~ij\tilde{a}_{ij}(right) are both given in matrix form. The bottom row displays the coefficients bib_{i} (left) and b~i\tilde{b}_{i} (right). The numerical coefficients are given as follows α=0.24169426078821,β=0.06042356519705,η=0.12915286960590,γ=1/2αβη\alpha=0.24169426078821~{},~{}\beta=0.06042356519705~{},~{}\eta=0.12915286960590~{},~{}\gamma=1/2-\alpha-\beta-\eta.

At every substep we need to solve an implicit equation for the substep 𝒱ik\mathcal{V}_{i}^{k}. Since the implicit matrix of Table 1 is lower triangular, each substep kk, the following expression can be directly computed using the information of previous substeps,

𝒱i(k)=𝒱i+Δtl=1k1akli(𝒰i(l),𝒱i(l))\displaystyle\mathcal{V}^{\ast\,\left(k\right)}_{i}=\mathcal{V}_{i}+{\Delta t}\sum_{l=1}^{k-1}a_{kl}\mathcal{E}_{i}\left(\mathcal{U}_{i}^{\left(l\right)},\mathcal{V}_{i}^{\left(l\right)}\right)
+Δtl=1k1a~kli(𝒰i(l),𝒱i(l)).\displaystyle\phantom{\mathcal{V}^{\ast\,\left(k\right)}_{i}=\mathcal{V}_{i}}+{\Delta t}\sum_{l=1}^{k-1}\tilde{a}_{kl}\mathcal{I}_{i}\left(\mathcal{U}_{i}^{\left(l\right)},\mathcal{V}_{i}^{\left(l\right)}\right). (99)

With that, the implicit equation can be written as

𝒱i(k)=𝒱i(k)+αΔti(𝒰i(k),𝒱i(k)),\displaystyle\mathcal{V}^{\left(k\right)}_{i}=\mathcal{V}^{\ast\,\left(k\right)}_{i}+\alpha\Delta t\mathcal{I}_{i}\left(\mathcal{U}_{i}^{\left(k\right)},\mathcal{V}_{i}^{\left(k\right)}\right)\,, (100)

where aii=αa_{ii}=\alpha. Since this equation is non-linear, numerical root-finding will be needed in order to obtain the intermediate stage 𝒱i(k)\mathcal{V}^{\left(k\right)}_{i}.

IV.1.1 Solving the implicit equation

We are now going to apply the implicit time stepping to the dissipative magnetohydrodynamics system. In doing so, we need to treat all source terms in Eqs. (76)-(83) using the ImEx method outlined above, see Eq. (99). In particular, we treat all stiff contributions proportional to τ1\tau^{-1} or ω1\omega^{-1} implicitly. We further introduce the definitions

qν=1ωq(qμuμ)uν\displaystyle\mathcal{I}_{q}^{\nu}=-\frac{1}{\omega_{q}}\left(q_{\mu}u^{\mu}\right)u^{\nu} (101)
παβ=1ωπ[(παλuλ)uβ+(πβλuλ)uα]\displaystyle\mathcal{I}_{\pi}^{\alpha\beta}=-\frac{1}{\omega_{\pi}}\left[\left(\pi^{\alpha\lambda}u_{\lambda}\right)u^{\beta}+\left(\pi^{\beta\lambda}u_{\lambda}\right)u^{\alpha}\right]
1ωππ(πμνgμν)gαβ.\displaystyle\phantom{\mathcal{I}_{\pi}^{\alpha\beta}=}-\frac{1}{\omega_{\pi\pi}}\left(\pi^{\mu\nu}g_{\mu\nu}\right)g^{\alpha\beta}. (102)

Since the baryon-number, energy- and momentum equations do not have stiff sources, the hydrodynamic variables ρ,T,uμ\rho,T,u^{\mu} are, hence, the same in the implicit and explicit stages. We can then write the full set of implicit equations as follows,

[Π+Yζ]u0=Π~αΔt[1τΠ+(1δΠΠτΠ)1ρθ]Π\displaystyle\left[\Pi+Y_{\zeta}\right]u^{0}=\tilde{\Pi}^{\ast}-\alpha\Delta t\left[\frac{1}{\tau_{\Pi}}+\left(1-\frac{\delta_{\Pi\Pi}}{\tau_{\Pi}}\right)\frac{1}{\rho}\mathcal{I}_{\theta}\right]\Pi
+αΔtζ\displaystyle\phantom{\left[\Pi+Y_{\zeta}\right]u^{0}}+\alpha\Delta t\mathcal{I}_{\zeta} (103)
qνu0+YκTΔ0ν=q~ναΔtωq(qμuμ)uναΔtδqBτqFνμqμ\displaystyle q^{\nu}\,u^{0}+Y_{\kappa}T\Delta^{0\nu}=\tilde{q}^{\ast\,\nu}-\frac{\alpha\Delta t}{\omega_{q}}\left(q^{\mu}u_{\mu}\right)u^{\nu}-\alpha\Delta t\frac{\delta_{qB}}{\tau_{q}}F^{\nu\mu}q_{\mu}
+αΔt[κν+qν],\displaystyle\phantom{q^{\nu}\,u^{0}+\frac{\kappa}{\tau_{q}}T\Delta^{0\nu}}+\alpha\Delta t\left[\mathcal{I}_{\kappa}^{\nu}+\mathcal{I}_{q}^{\nu}\right], (104)
παβu0=π~αβ2Yη(g0(αuβ))\displaystyle\pi^{\alpha\beta}u^{0}=\tilde{\pi}^{\ast\,\alpha\beta}-2Y_{\eta}\left(g^{0\left(\alpha\right.}u^{\left.\beta\right)}\right)
αΔt[1τπ+(1δππτπ)1ρθ]\displaystyle\phantom{\pi^{\alpha\beta}u^{0}=}-\alpha\Delta t\left[\frac{1}{\tau_{\pi}}+\left(1-\frac{\delta_{\pi\pi}}{\tau_{\pi}}\right)\frac{1}{\rho}\mathcal{I}_{\theta}\right]
+αΔtπαβαΔtδπBτπFδμπμγΔγδαβ,\displaystyle\phantom{\pi^{\alpha\beta}u^{0}=}+\alpha\Delta t\mathcal{I}_{\pi}^{\alpha\beta}-\alpha\Delta t\frac{\delta_{\pi B}}{\tau_{\pi}}F^{\delta\mu}\pi^{\gamma}_{\mu}\,\Delta^{\alpha\beta}_{\gamma\delta}, (105)
ρu0Yζ=ρu0Y~ζαΔtρζ,\displaystyle\rho u^{0}Y_{\zeta}=\rho u^{0}\tilde{Y}_{\zeta}^{\ast}-\alpha\Delta t\rho\mathcal{I}_{\zeta}, (106)
ρu0Yκ=ρu0Y~ζαΔtρκ,\displaystyle\rho u^{0}Y_{\kappa}=\rho u^{0}\tilde{Y}_{\zeta}^{\ast}-\alpha\Delta t\rho\mathcal{I}_{\kappa}, (107)
ρu0Yη=ρu0Y~ηαΔtρη,\displaystyle\rho u^{0}Y_{\eta}=\rho u^{0}\tilde{Y}_{\eta}^{\ast}-\alpha\Delta t\rho\mathcal{I}_{\eta}, (108)
ρu0Yθ=ρu0Y~θαΔtρθ,\displaystyle\rho u^{0}Y_{\theta}=\rho u^{0}\tilde{Y}_{\theta}^{\ast}-\alpha\Delta t\rho\mathcal{I}_{\theta}, (109)
Zν=Z~ναΔtρκν.\displaystyle Z_{\nu}=\tilde{Z}_{\nu}^{\ast}-\alpha\Delta t\rho\mathcal{I}_{\kappa}^{\nu}. (110)

In the stiff limit of ωq/π/ππ0\omega_{q/\pi/\pi\pi}\rightarrow 0, these equations will have the following solution for a given velocity vector uiu_{i} and temperature TT,

Π=(u0+αΔtΔΠ)1[Π~/u0Yζ],\displaystyle\Pi=\left(u^{0}+\alpha\Delta t\Delta_{\Pi}\right)^{-1}\left[\tilde{\Pi}^{\ast}/u^{0}-Y_{\zeta}^{\ast}\right], (111)
qμ=(u0+αΔtΔq)1𝒬νμ[(q)νT(Z)ν],\displaystyle q^{\mu}=\left(u^{0}+\alpha\Delta t\Delta_{q}\right)^{-1}\mathcal{Q}^{\mu}_{\nu}\left[\left(q^{\ast}\right)^{\nu}-T\left(Z^{\ast}\right)^{\nu}\right], (112)
παβ=((u0)2+αΔtΔπu0)1𝒫γδαβΔμνγδ(π)μν,\displaystyle\pi^{\alpha\beta}=\left(\left(u^{0}\right)^{2}+\alpha\Delta t\Delta_{\pi}u^{0}\right)^{-1}\mathcal{P}_{\gamma\delta}^{\alpha\beta}\Delta^{\gamma\delta}_{\mu\nu}\left(\pi^{\ast}\right)^{\mu\nu}, (113)
Yζ=ωζωζ+αΔtYζ+αΔtωζ+αΔtζτΠ,\displaystyle Y_{\zeta}=\frac{\omega_{\zeta}}{\omega_{\zeta}+\alpha\Delta t}Y^{\ast}_{\zeta}+\frac{\alpha\Delta t}{\omega_{\zeta}+\alpha\Delta t}\frac{\zeta}{\tau_{\Pi}}, (114)
Yκ=ωκωκ+αΔtYκ+αΔtωκ+αΔtκτq,\displaystyle Y_{\kappa}=\frac{\omega_{\kappa}}{\omega_{\kappa}+\alpha\Delta t}Y^{\ast}_{\kappa}+\frac{\alpha\Delta t}{\omega_{\kappa}+\alpha\Delta t}\frac{\kappa}{\tau_{q}}, (115)
Yη=ωηωη+αΔtYη+αΔtωη+αΔtητπ,\displaystyle Y_{\eta}=\frac{\omega_{\eta}}{\omega_{\eta}+\alpha\Delta t}Y^{\ast}_{\eta}+\frac{\alpha\Delta t}{\omega_{\eta}+\alpha\Delta t}\frac{\eta}{\tau_{\pi}}, (116)
Yθ=ωθωθ+αΔtYθ+αΔtωθ+αΔtρ,\displaystyle Y_{\theta}=\frac{\omega_{\theta}}{\omega_{\theta}+\alpha\Delta t}Y^{\ast}_{\theta}+\frac{\alpha\Delta t}{\omega_{\theta}+\alpha\Delta t}\rho, (117)
Zν=ωκωκ+αΔtZν+αΔtωκ+αΔtδν0κτq,\displaystyle Z_{\nu}=\frac{\omega_{\kappa}}{\omega_{\kappa}+\alpha\Delta t}Z_{\nu}^{\ast}+\frac{\alpha\Delta t}{\omega_{\kappa}+\alpha\Delta t}\delta^{0}_{\nu}\frac{\kappa}{\tau_{q}}, (118)

where we have used that

ΔΠ=[1τΠ+(1δΠΠτΠ)1ρθ],\displaystyle\Delta_{\Pi}=\left[\frac{1}{\tau_{\Pi}}+\left(1-\frac{\delta_{\Pi\Pi}}{\tau_{\Pi}}\right)\frac{1}{\rho}\mathcal{I}_{\theta}\right]\,, (119)
Δq=[1τq+(1δqqτq)1ρθ],\displaystyle\Delta_{q}=\left[\frac{1}{\tau_{q}}+\left(1-\frac{\delta_{qq}}{\tau_{q}}\right)\frac{1}{\rho}\mathcal{I}_{\theta}\right]\,, (120)
Δπ=[1τπ+(1δππτπ)1ρθ],\displaystyle\Delta_{\pi}=\left[\frac{1}{\tau_{\pi}}+\left(1-\frac{\delta_{\pi\pi}}{\tau_{\pi}}\right)\frac{1}{\rho}\mathcal{I}_{\theta}\right]\,, (121)
Δtπ=αΔtδπB/τπ,\displaystyle\Delta t_{\pi}^{\prime}=\alpha\Delta t\delta_{\pi B}/\tau_{\pi}\,, (122)
Δtq=αΔtδqB/τq,\displaystyle\Delta t_{q}^{\prime}=\alpha\Delta t\delta_{qB}/\tau_{q}\,, (123)
𝒬μν=Δμν(Δtq)2b21+(Δtq)2b2ΞμνΔtq1+(Δtq)2b2Fμν,\displaystyle\mathcal{Q}^{\mu\nu}=\Delta^{\mu\nu}-\frac{\left(\Delta t_{q}^{\prime}\right)^{2}b^{2}}{1+\left(\Delta t_{q}^{\prime}\right)^{2}b^{2}}\Xi^{\mu\nu}-\frac{\Delta t_{q}^{\prime}}{1+\left(\Delta t_{q}^{\prime}\right)^{2}b^{2}}F^{\mu\nu}, (124)
𝒫γδαβ=b2bαbβbγbδ+bα𝔭γδβ+bβ𝔭γδα+𝔓γδαβ,\displaystyle\mathcal{P}^{\alpha\beta}_{\gamma\delta}=b^{-2}\,b^{\alpha}b^{\beta}b_{\gamma}b_{\delta}+b^{\alpha}\mathfrak{p}^{\beta}_{\gamma\delta}+b^{\beta}\mathfrak{p}^{\alpha}_{\gamma\delta}+\mathfrak{P}^{\alpha\beta}_{\gamma\delta}, (125)
b2𝔭μνα=(1+2(Δtπ)2b2)1[bμΞνα(Δtπ)bμFνα]\displaystyle\sqrt{b^{2}}\,\mathfrak{p}^{\alpha}_{\mu\nu}=\left(1+2\left(\Delta t_{\pi}^{\prime}\right)^{2}b^{2}\right)^{-1}\left[b_{\mu}\Xi^{\alpha}_{\nu}-\left(\Delta t_{\pi}^{\prime}\right)b_{\mu}F^{\alpha}_{\nu}\right] (126)
𝔓μναβ=(1+2(Δtπ)2b2)(1+4(Δtπ)2b2)ΞμαΞνβ\displaystyle\mathfrak{P}^{\alpha\beta}_{\mu\nu}=\frac{\left(1+2\left(\Delta t_{\pi}^{\prime}\right)^{2}b^{2}\right)}{\left(1+4\left(\Delta t_{\pi}^{\prime}\right)^{2}b^{2}\right)}\Xi^{\alpha}_{\mu}\Xi^{\beta}_{\nu}
Δtπ1+4(Δtπ)2b2[FμαΞνβ+FμβΞνα]\displaystyle\phantom{\mathfrak{P}^{\alpha\beta}_{\mu\nu}}-\frac{\Delta t_{\pi}^{\prime}}{1+4\left(\Delta t_{\pi}^{\prime}\right)^{2}b^{2}}\left[F^{\alpha}_{\mu}\Xi^{\beta}_{\nu}+F^{\beta}_{\mu}\Xi^{\alpha}_{\nu}\right]
+2(Δtπ)21+4(Δtπ)2b2FμαFνβ.\displaystyle\phantom{\mathfrak{P}^{\alpha\beta}_{\mu\nu}}+\frac{2\left(\Delta t_{\pi}^{\prime}\right)^{2}}{1+4\left(\Delta t_{\pi}^{\prime}\right)^{2}b^{2}}F^{\alpha}_{\mu}F^{\beta}_{\nu}. (127)
Refer to caption
Figure 1: One-dimensional blast wave problem for different choices of the bulk viscous relaxation time τΠ\tau_{\Pi} for fixed κ/τΠ=1\kappa/\tau_{\Pi}=1. The left panel shows the rest-mass density ρ\rho at time t=0.4t=0.4, the right panel the four-velocity component uxu^{x}.

While those relations can straightforwardly be used when the four-velocity uμu^{\mu} and the temperature TT are known, this is not the case at the end of the explicit stage, when only the conserved variables ρ\rho_{\ast}, ee_{\ast} and SiS_{i} are given. Hence, the solution of the implicit equation needs to be embedded in an outer root-finding loop trying to recover those variables simultaneously. This problem has been studied extensively in the context of relativistic (ideal-) (magneto-)hydrodynamics Noble et al. (2006); Galeazzi et al. (2013); Newman and Hamlin (2014); Siegel et al. (2018); Kastaun et al. (2021), and we refer to these studies for further details.

In short, our iterative procedure to solve the implicit equations proceeds as follows:

  1. 1.

    Given the current guess for the spatial part of the four velocity u¯i\bar{u}^{i} and the temperature T¯\bar{T}, compute the full four-velocity u¯0\bar{u}^{0} from u¯μu¯νgμν=1\bar{u}^{\mu}\bar{u}^{\nu}g_{\mu\nu}=-1 and the baryon density ρ¯=ρ/u¯0\bar{\rho}=\rho_{\ast}/\bar{u}^{0}. Together with the equation of state a full guess for the hydrodynamical state (ρ¯,T¯,u¯i)\left(\bar{\rho},\bar{T},\bar{u}^{i}\right) is then available.

  2. 2.

    Compute all transport coefficients using the above guess for the hydrodynamical state, compute all first-order transport coefficients (η,κ,ζ)\left(\eta,\kappa,\zeta\right) and relaxation times (τΠ,τq,τπ)\left(\tau_{\Pi},\tau_{q},\tau_{\pi}\right), as well as the second-order transport terms (δΠΠ,δqq,δππ)\left(\delta_{\Pi\Pi},\delta_{qq},\delta_{\pi\pi}\right) and the magnetic field couplings (δqB,δπB)\left(\delta_{qB},\delta_{\pi B}\right).

  3. 3.

    Using the approximate state (ρ¯,T¯,u¯i)\left(\bar{\rho},\bar{T},\bar{u}^{i}\right), solve the implicit equations (111)-(118) in order to recover the dissipative variables (Π¯,q¯μ,π¯μν)\left(\bar{\Pi},\bar{q}^{\mu},\bar{\pi}^{\mu\nu}\right).

  4. 4.

    Next, use the approximation Π¯,q¯μ,π¯μν\bar{\Pi},\bar{q}^{\mu},\bar{\pi}^{\mu\nu} to the dissipative variables obtained in the previous step to compute the purely hydrodynamic part of the ee_{\ast} and SiS_{i},

    ehydro=\displaystyle e^{\rm hydro}_{\ast}= e+b0b0π002q0u0\displaystyle\,e_{\ast}+b^{0}b^{0}-\pi^{00}-2q^{0}u^{0}
    (12b2+Π)Δ0012b2(u0)2,\displaystyle-\left(\frac{1}{2}b^{2}+\Pi\right)\Delta^{00}-\frac{1}{2}b^{2}\left(u^{0}\right)^{2}\,, (128)
    Sihydro=\displaystyle S_{i}^{\rm hydro}= Siπi0q0uiqiu0+b0bi\displaystyle\,S_{i}-\pi_{i}^{0}-q^{0}u_{i}-q_{i}u^{0}+b^{0}b_{i}
    (12b2+Π)Δi012b2u0ui,\displaystyle-\left(\frac{1}{2}b^{2}+\Pi\right)\Delta^{0}_{i}-\frac{1}{2}b^{2}u^{0}u_{i}\,, (129)

    where all variables are to be interpreted as the approximate solutions obtained before, although we suppress the notation here for readability. These two equations can then be inverted using relations for a standard hydrodynamical inversion. In particular, we use Galeazzi et al. (2013),

    ε\displaystyle\varepsilon =u0eg00ρSiSiρ1,\displaystyle=u^{0}\frac{e_{\ast}}{g^{00}\rho_{\ast}}-\frac{\sqrt{S_{i}S^{i}}}{\rho_{\ast}}-1\,, (131)

    where ε=e/ρ1\varepsilon=e/\rho-1 is the specific internal energy density.

  5. 5.

    Using the equation of state, we can compute the residuals via

    resT=T(ε¯,ρ¯,)T¯,\displaystyle{\rm res}_{T}=T\left(\bar{\varepsilon},\bar{\rho},\dots\right)-\bar{T}\,, (132)
    resSi=Si(e¯+P¯)u¯0u¯i,\displaystyle{\rm res}_{S}^{i}=\frac{S_{i}}{\left(\bar{e}+\bar{P}\right)\bar{u}^{0}}-\bar{u}_{i}\,, (133)

    where the second terms are always given by the initial guesses from Step 1. Using this residual the root-finding algorithm repeats this procedure restarting at the first step, until these residuals are small. We typically demand that max(|resT/T|,|resSi|)<1011\max\left(\left|{\rm res}_{T}/T\right|,\left|{\rm res}_{S}^{i}\right|\right)<10^{-11} for successful convergence.

To solve the implicit equation numerically, it turned out to be crucial to use a stable root-finding algorithm as the Jacobian of the system could easily become singular. While we were able to obtain good results for heat conduction only with a multi-dimensional Newton-method Broyden (1965), solving the shear stresses required the use of a Newton-Krylov solver Kelley (1995). Moreover, using a good initial guess is crucial to the successful solution of this system. We empirically found that for initial guesses to the root-finder that differed substantially from the true solution, the inversion converged to unphysical states. Differently from ideal magnetohydrodynamics inversions Newman and Hamlin (2014); Siegel et al. (2018); Kastaun et al. (2021), unphysical in this context means that the viscous degrees of freedom were unphysically large, whereas the hydrodynamic variables were still well defined, in terms of positive pressures and finite velocities.
We therefore found it beneficial to first compute an initial guess for the Newton-Krylov solver using a standard ideal magnetohydrodynamics inversion scheme Newman and Hamlin (2014), which neglects the viscous contributions. Further development and investigation will be needed to improve the robustness and computational cost of this step, but the current approach is sufficient to provide accurate evolutions of all test problems presented in this work.

V Numerical tests

Having introduced a new numerical formulation of the non-resistive viscous relativistic hydrodynamics system, we want to assess its ability to handle

each of the dissipative contributions. In the following, we present an initial set of problems designed to test each of those transport contributions individually. Starting from one-dimensional problems for bulk viscosity and heat conduction, we continue with a two-dimensional test of shear viscosity and the impact of varying the thermal gyrofrequency parameter in the presence of a magnetic field. For simplicity, all tests adopt a simplified Γ\Gamma-law equation of state, i.e. P=ρε(Γ1)P=\rho\varepsilon\left(\Gamma-1\right), where Γ\Gamma will be given in the description of each test. As is common in numerical code testing (see e.g. Del Zanna et al. (2007); Stone et al. (2020)) we will adopt code units throughout. That is, the units of all quantities are implicitly specified relative to each other.

V.1 Bulk viscosity

Refer to caption
Figure 2: Thermal dissipation of an initial Gaussian temperature, TT, profile. The transport coefficients κ\kappa and τq\tau_{q} are varied in each case. The different colors denote different times tt during the diffusion process.
Refer to caption
Figure 3: Comparison of the dissipation of the top, x=0x=0 of a Gaussian temperature, TT, profile. Shown are two evolution profiles for two sets of transport coefficients ζ\zeta and τq\tau_{q}. These are compared to analytic solutions of the heat equation, Eq. (137), and the Telegrapher’s equation, Eq. (138).

In this test we investigate the impact of bulk viscosity on a one-dimensional relativistic shocktube problem, see also Ref. Chabanov et al. (2021) for a similar test. Following Radice and Rezzolla (2012), we adopt the following initial conditions for a one-dimensional blast wave launched into an ambient medium. We adopt a domain [0.5;0.5]\left[-0.5;0.5\right] along the xx-axis . All primitive variables, including the dissipative sector, are initialized to zero. The only non-zero quantities are given separately for x<0x<0, and x>0x>0 as follows :

ρ=(103,103),\displaystyle\rho=\left(10^{-3},10^{-3}\right), (134)
P=(1,105).\displaystyle P=\left(1,10^{-5}\right). (135)

The system is closed by adopting an equation of state with Γ=5/3\Gamma=5/3. In order to study the behavior of bulk viscous dissipation, we adopt a fixed ζ/τΠ=1\zeta/\tau_{\Pi}=1 and vary τΠ\tau_{\Pi}. This will allow us to probe the τΠ0\tau_{\Pi}\to 0 limit discussed in Sec. III.1.1 . All other dissipative coefficients are set to zero, i.e. η=κ=0\eta=\kappa=0 with their corresponding relaxation timescales fixed to values smaller than the dynamical time of the problem, i.e. τq=τπ=105\tau_{q}=\tau_{\pi}=10^{-5} . With this choice, the system will only be subject to bulk viscous dissipation. The evolution of the baryon density ρ\rho and velocity uxu^{x} is shown in Fig. 1 for different values of the bulk viscous relaxation time τΠ\tau_{\Pi}. We can anticipate that for this particular choice of the transport coefficients the code will converge to two limiting solutions, see Sec. III.1.1. In the case of τΠ0\tau_{\Pi}\rightarrow 0 (magenta curve), the effective timescale associated lengthscale viscζ1=τΠ1\ell_{\rm visc}\sim\zeta^{-1}=\tau_{\Pi}^{-1} of the bulk viscosity will be too small to affect the dynamics of the shock problem, which happens on scales dyn0.01\ell_{\rm dyn}\gg 0.01 . Hence, the solution will approach the perfect fluid solution. On the other hand, following the discussion in Sec. III.1.1, we approach a non-perfect fluid limit already in the case of τΠ=10\tau_{\Pi}=10, which changes the shock structure (blue curve in Fig. 1).

Intermediate values of the bulk viscosity, dynvisc\ell_{\rm dyn}\simeq\ell_{\rm visc}, on the other hand, have the ability to interpolate between the two limits. This transition can be best understood when looking at the velocity uxu^{x}, where for increasing relaxation time τΠ\tau_{\Pi} the velocity profile transitions from the higher speed perfect fluid solution (magenta curve) to the advected solution (blue curve) at slightly lower velocities.

V.2 One-dimensional heat conduction

Refer to caption
Figure 4: Anisotropic heat conductivity test. Starting from an initially hot inner cylinder at temperature ThT_{h} surrounded by an ambient medium at temperature TcT_{c} (left column), the evolution of relativistic causal heat conduction is shown for two different times in the middle and right columns. The rows correspond to various degrees of anisotropy with respect to the magnetic field (shown as white streamlines). We can see that in the most anisotropic case (bottom row), heat can only flow along the magnetic field lines. A more detailed description is given in Sec. V.3.

We proceed by analyzing the ability of the system to conduct heat in a one-dimensional test setup. In order to pick a suitable set of initial conditions, we recall that in the limit of ui0u^{i}\rightarrow 0, the system reduces to the Telegrapher’s equation (38). In the Navier-Stokes limit of τq0\tau_{q}\rightarrow 0 we further recover,

κmb(Γ1)kBρΔTtT=0,\displaystyle\frac{\kappa m_{b}\left(\Gamma-1\right)}{k_{B}\rho}\Delta T-\partial_{t}T=0, (136)

which is the standard heat equation. The fundamental solution to this equation, i.e. starting from initial conditions T(x,t=0)=δ(x)T\left(x,t=0\right)=\delta\left(x\right), is a Gaussian,

T(x,t)=14πξtexp(x24ξt).\displaystyle T\left(x,t\right)=\frac{1}{\sqrt{4\pi\xi t}}\exp\left(-\frac{x^{2}}{4\xi t}\right). (137)

Here, we have introduced the effective diffusion constant ξ=κmb(Γ1)kBρ\xi=\frac{\kappa m_{b}\left(\Gamma-1\right)}{k_{B}\rho}. While this solution only holds approximately in the limit of τq0\tau_{q}\rightarrow 0, the full solution of the Telegrapher’s equation for the above initial condition instead reads Romatschke (2010b),

T(x,t)=\displaystyle T\left(x,t\right)= Θ(t)Θ(t2ξτqx2)14ξτq\displaystyle\Theta\left(t\right)\Theta\left(\frac{t^{2}\xi}{\tau_{q}}-x^{2}\right)\frac{1}{\sqrt{4\xi\tau_{q}}}
×et2/(2τq)I0(t24τqx24ξτq),\displaystyle\times e^{-t^{2}/\left(2\tau_{q}\right)}I_{0}\left(\sqrt{\frac{t^{2}}{4\tau_{q}}-\frac{x^{2}}{4\xi\tau_{q}}}\right), (138)

where the Heaviside functions Θ\Theta ensure causality of the solution. Here, I0I_{0} is a modified Bessel function of the first kind. Motivated by these observations we initialize an initial temperature distribution to resemble a Gaussian in hydrostatic equilibrium, i.e.,

T=exp(x2/0.01),\displaystyle T=\exp\left(-x^{2}/0.01\right), (139)
p=0.1,\displaystyle p=0.1, (140)
ρ=p/T(assuming kB=mb=1),\displaystyle\rho=p/T\ \left(\text{assuming\ }k_{B}=m_{b}=1\right), (141)

where we adopt Γ=4/3\Gamma=4/3. Our simulations use a grid resolution corresponding to Nx=1600N_{x}=1600 grid points in the interval [0.5:0.5][-0.5:0.5]. We then show the resulting profiles for different times in Fig. 2. We can see that the evolution differs drastically for fixed κ/τq=0.02\kappa/\tau_{q}=0.02 . For low τq=0.05\tau_{q}=0.05 and, hence, fast relaxation (left panel), we can see that the initial Gaussian slowly diffuses. On the other hand, for larger τq=0.5\tau_{q}=0.5 (center panel) the diffusive process happens more rapidly but the Gaussian is eventually flattening. For even larger relaxation time, τq=5.0\tau_{q}=5.0 , which is much larger than the dynamical time scale of the simulation, we see that the wave part in the Telegrapher’s equation (38) takes over and the Gaussian begins to split apart. Hence, in this limit of large mean-free path the dynamics is more similar to a damped wave equation than to a diffusion equation.

Refer to caption
Figure 5: Two-dimensional Kelvin-Helmholtz test with heat conduction at t=3.4t=3.4. Shown is the rest-mass density ρ\rho. In the absence of shear viscosity secondary vortices form at higher resolutions, denoted by the number of grid points Ny×NxN_{y}\times N_{x}.

In order to assess whether we are indeed recovering the correct analytic solution of the Telegrapher’s (38) and Heat equations (136), we compare the evolution of the center point, T(x=0,t)T\left(x=0,t\right) with the analytic solutions (138) and (137). This comparison is shown in Fig. 3. We can see that for different choices of the transport coefficient κ/τq\kappa/\tau_{q} the two limits can be reliably recovered. In particular, it is worth highlighting that causality, which limits the instantaneous spreading of heat, can significantly slow down dissipation compared to standard parabolic heat conduction, for which T(x=0,t)t1/2T\left(x=0,t\right)\propto t^{-1/2}.

V.3 Two-dimensional heat conduction

We now continue to explore the effects of thermal conductivity in a two-dimensional setting. In particular, we follow a setup first proposed by Ref. Chandra et al. (2017),

ρ={0.8x2+y2<0.08,1.0x2+y2>0.08,\displaystyle\rho=\begin{cases}0.8&\quad\sqrt{x^{2}+y^{2}}<0.08,\\ 1.0&\quad\sqrt{x^{2}+y^{2}}>0.08,\end{cases} (142)
P=Γ1,\displaystyle P=\Gamma-1, (143)
Bx=104,\displaystyle B^{x}=10^{-4}, (144)
By=104sin(16πx).\displaystyle B^{y}=10^{-4}\sin\left(16\pi x\right). (145)

As before, all quantities not explicity listed have been initialized to zero. We emphasize that the initial condition is in pressure equilibrium and, in the absence of initial velocities, would remain static if dissipative effects were not present. Adopting an equation of state Γ=4/3\Gamma=4/3, this corresponds to an inner hot region with temperature Th=5/12T_{h}=5/12 and an outer cold region with temperature Tc=4/12T_{c}=4/12. In this problem, we explicitly include a global magnetic field to study the dependence with the gyrofrequency parameter δqB\delta_{qB}. Inspired by the functional form of this parameter found for ultrarelativistic gases Denicol et al. (2019), we chose

τq=0.60,\displaystyle\tau_{q}=0.60, (147)
κ=0.02,\displaystyle\kappa=0.02, (148)
δqBτqT2=const.\displaystyle\frac{\delta_{qB}}{\tau_{q}}T^{2}=\rm const. (149)

The evolution for different values of δqB\delta_{qB} is shown in Fig. 4. We can see that in the almost isotropic case, i.e. δqB0\delta_{qB}\rightarrow 0, heat conduction proceeds independently of the magnetic field geometry. More specifically, we can see in the middle and right panels of the first row in Fig. 4 that the temperature evolution retains its cylindrical symmetry. With increasing degree of anisotropy (middle row), we can see that the temperature evolution begins to be affected by the presence of the magnetic field. Physically, the increase in gyrofrequency beings to suppress cross-conductivity. In particular, the initially cylindrical temperature profile begins to split up (center panel) and heat conduction along the magnetic field starts to be enhanced. Comparing the final times (right column), we still find that the overall profile of the heat conduction is almost isotropic with only a small degree of anisotropy being present. Finally, by further increasing the degree of anisotropy δqB\delta_{qB}, the heat flux begins to fully align with the magnetic field, i.e. qμq0bμq^{\mu}\rightarrow q_{0}b^{\mu}, where the latter approximation is valid in the limit of small velocities. This is the limit of extended magnetohydrodynamics discussed in Sec. III.4.1.

We can see from the bottom row of Fig. 4 that in this limit heat conduction across the magnetic field lines is essentially absent and that the evolution is similar to the test case presented in Ref. Chandra et al. (2017). We note, however, that because the closure adopted in the present formulation allows to dynamically adjust the degree of anisotropy, we can naturally interpolate between the highly collisional (top row) and weakly collisional limits (bottom row).

V.4 Two-dimensional Kelvin-Helmholtz instability

Refer to caption
Refer to caption
Refer to caption
Refer to caption
Figure 6: Two-dimensional Kelvin-Helmholtz test with heat conduction and shear viscosity at t=3.4t=3.4. In the presence of shear viscosity the same converged vortex structure is recovered also at higher resolutions, denoted by the number of grid points Ny×NxN_{y}\times N_{x}. Shown are the rest-mass density ρ\rho (top row), the four-velocity uyu^{y} along the shear layer (second row), and the shear stress tensor components πxy\pi^{xy} (third row) and πyy\pi^{yy} (bottom row).

Having discussed the behavior of bulk viscosity and heat conductivity in a series of numerical tests, we now turn to the effect of shear viscosity. In particular, we focus on a standard test problem routinely studied in the context of the Newtonian Navier-Stokes equations Lecoanet et al. (2016). Adopting the initial conditions presented in Beckwith and Stone (2011), we study the formation of vortices in a two-dimensional Kelvin-Helmholtz unstable shear layer. More precisely, we use

p=1,\displaystyle p=1\,, (150)
ρ={0.505+0.495tanh[(x0.5)/0.01],x>0,0.5050.495tanh[(x+0.5)/0.01],x0,\displaystyle\rho=\begin{cases}0.505+0.495\,{\rm tanh}\left[\left(x-0.5\right)/0.01\right]\,,&\quad x>0\,,\\ 0.505-0.495\,{\rm tanh}\left[\left(x+0.5\right)/0.01\right]\,,&\quad x\leq 0\,,\end{cases} (151)
uxu0={0.05sin(2πy)exp[(x0.5)2/0.1],x>0,0.05sin(2πy)exp[(x+0.5)2/0.1],x0.\displaystyle\frac{u^{x}}{u^{0}}=\begin{cases}0.05\,\sin\left(2\pi y\right){\rm exp}\left[-\left(x-0.5\right)^{2}/0.1\right]\,,&\quad x>0\,,\\ -0.05\,\sin\left(2\pi y\right){\rm exp}\left[-\left(x+0.5\right)^{2}/0.1\right]\,,&\quad x\leq 0\,.\end{cases} (152)
uyu0={0.5tanh[(x0.5)/0.01],x>0,0.5tanh[(x+0.5)/0.01],x0,\displaystyle\frac{u^{y}}{u^{0}}=\begin{cases}0.5\,{\rm tanh}\left[\left(x-0.5\right)/0.01\right]\,,&\quad x>0\,,\\ 0.5\,{\rm tanh}\left[\left(x+0.5\right)/0.01\right]\,,&\quad x\leq 0\,,\end{cases} (153)

where we have further adopted Γ=4/3\Gamma=4/3. As before, all quantities not listed above have been initialized to zero. For this test problem we include both shear viscosity and heat conduction by fixing

τq=0.05,\displaystyle\tau_{q}=0.05, (154)
τπ=0.1,\displaystyle\tau_{\pi}=0.1, (155)
κ=6×104τq,\displaystyle\kappa=6\times 10^{-4}\tau_{q}, (156)
η=5×103τπ,\displaystyle\eta=5\times 10^{-3}\tau_{\pi}, (157)

which gives rise to an effective thermal Prandtl number Pr:=η/κ8{\rm Pr}:=\eta/\kappa\approx 8.

Shear stresses act in providing a fixed cut-off for small scale turbulence, which is set by the length scale of the shear viscosity shearη/(ρv¯)\ell_{\rm shear}\simeq\eta/\left(\rho\bar{v}\right), where v¯\bar{v} is a characteristic velocity scale. If instead of an explicit viscosity this cut-off was solely provided by the grid scale, i.e. shearΔx\ell_{\rm shear}\sim\Delta x, the outcome of the simulation would strongly depend on the given resolution. Hence, demonstrating resolved and converged vortex formation in a Kelvin-Helmholtz unstable shear layer has become a standard test problem for non-relativistic hydrodynamics Lecoanet et al. (2016). Following the same logic, we will perform simulations at three different resolutions, labeled by the number of grid points Nx×(2Nx)N_{x}\times\left(2N_{x}\right), where Nx[256,512,1024]N_{x}\in\left[256,512,1024\right].

In order to establish a baseline for the effect of underresolved shear viscosity, we perform a first set of simulations for η=0\eta=0. The resulting baryon density evolution during vortex formation is shown in Fig. 5. Comparing the lowest (left panel) and highest resolutions (right panel), we clearly find differences in the small scale evolution, particularly inside the vortex. The presence of grid dependent effective viscosity can best be appreciated by comparing the medium (middle panel) and high (right panel) resolution cases. While on first glance the density distributions look very similar, one can clearly spot the presence of secondary vortices being formed in the shear layers of the primary vortex. If the shear viscosity was instead constant and resolved by the numerical resolution, we would expect convergent behavior in the vortex evolution. That is, getting the same vortex above a certain sufficiently resolved resolution.

To demonstrate that this is indeed the case when using our viscous scheme, we perform the same simulation, but with the shear viscosity η\eta given by Eq. (157). The resulting evolutions for the baryon density ρ\rho, the velocity component uyu^{y} along the shear layer, and the shear stresses πxy\pi^{xy} and πyy\pi^{yy} are shown in Fig. 6. Comparing the density evolutions with those shown in Fig. 5, the effect of a resolved shear viscosity is strikingly obvious. The formation of secondary vortices is suppressed at all resolutions and the vortices have the same internal structure in all cases, establishing that the relativistic form of the shear viscosity in flux-conservative form is working as expected. To better illustrate the dynamics, we next focus on the velocity uyu^{y} along the shear layer (second row in Fig. 6). We can indeed see that relativistic velocities uy1u^{y}\simeq 1 are present in the vortices. Furthermore we can see that strong shock fronts propagate through the vortices along the shear layer, coinciding with the jump (red to green color) in the rest-mass density ρ\rho. Looking at the shear stresses πxy\pi^{xy} (third row in Fig. 6), we can see that the shear stresses are perfectly resolved and converged at all resolutions, with the strongest stresses present inside the vortices, as expected.

The most remarkable feature of our numerical simulations now arise in the bottom panel of Fig. 6. Since these simulations do not include explicit bulk viscosity, i.e. ζ=0\zeta=0, the strong shock fronts propagating along the shear layer are not resolved and should manifest almost as discontinuities, similar to the shock tube solutions for perfect fluids shown in Fig. 1. Taking πyy\pi^{yy} as a proxy for these stresses, we indeed find very sharp almost discontinuous jumps across the shocks, getting sharper when going from the lowest (left panel) to the highest (right panel) simulation. While handling such strong discontinuities might be difficult with more traditional finite-difference approaches to the shear stresses Del Zanna et al. (2013); Bazow et al. (2018), our way of incorporating the first-order dissipative forms in flux-conservative form clearly allows us to perfectly handle those steep jumps using high-resolution shock capturing methods.

Refer to caption
Refer to caption
Refer to caption
Figure 7: Two-dimensional Kelvin-Helmholtz test with anisotropic heat conduction, shear viscosity and finite thermal gyrofrequency at t=3.4t=3.4. the left column shows results in the extended magnetohydrodynamics (Braginskii-like) limit of the closure. The center column in addition adds a subgrid model to include kinetic effects (limiting the pressure anisotropies according to mirror and firehose instabilities) that increase collisionality. The right column corresponds to a non-resistive viscous simulation. The rows show the pressure anisotropy |PP|\left|P_{\parallel}-P_{\perp}\right| relative to the magnetic field, the in-plane shear-stresses πxy\pi^{xy} and the norm of the in-plane heat flux |𝐪|\left|\bf q\right|.

V.5 Two-dimensional Kelvin-Helmholtz instability in the presence of magnetic fields

As a final test, we investigate the weakly collisional regime when shear stresses and heat fluxes are included. As discussed in Sec. III.4.1, in the limit of large mean-free-path the couplings δqB\delta_{qB} and δπB\delta_{\pi B} diverge, leading to the heat flux and shear stresses to align with the co-moving magnetic field.
In the case of globally sheared accretion flows, hybrid-kinetic simulations of shearing flows in accretion disks have shown that in this limit mirror and firehose instabilities lead to a local enhancement of collisionality Kunz et al. (2016). These instabilities are expected to kick in, once Foucart et al. (2016)

π0>32b2(firehose instability),\displaystyle\pi_{0}>\frac{3}{2}b^{2}~{}(\text{firehose instability}), (158)
π0<32b2π0+Pπ0P(mirror instability).\displaystyle\pi_{0}<\frac{3}{2}b^{2}\frac{\pi_{0}+P}{\pi_{0}-P}~{}(\text{mirror instability}). (159)

Additionally, the heat flux is expected to be bound Cowie and McKee (1977); Foucart et al. (2016),

|q0|ρcs3,\displaystyle\left|q_{0}\right|\lesssim\rho c_{s}^{3}, (160)

where csc_{s} is the sound speed in the fluid. Once these instabilities are triggered they lead to a local enhancement of collisionality and effectively pin the stresses at the threshold values for the instability. Within the framework of dissipative hydrodynamics, this can most easily be incorporated by locally adjusting the relaxation time τq/πλmfp\tau_{q/\pi}\propto\lambda_{\rm mfp}, while keeping the ratios of the transport coefficients fixed, i.e. κ/τq,η/τπconst\kappa/\tau_{q}\,,\eta/\tau_{\pi}\,\simeq\,\rm const.

To do this, we adjust the relaxation times during the implicit solve of Eq. (112) and (113). We then consider the same setup presented in Sec. V.4, but add a uniform magnetic field

By=103,\displaystyle B^{y}=10^{-3}, (161)

while keeping the other magnetic field components at zero initially. This results in an initial plasma parameter β=P/B2=106\beta=P/B^{2}=10^{6}, for which the magnetic field in itself is not dynamically important. That is, in the absence of magnetic field induced anisotropies of heat fluxes and shear stresses, the simulation outcome will not differ from a purely hydrodynamical simulation. However, because of the anisotropic coupling with the magnetic field, the differences observed between the simulations will be purely because of the anisotropic nature of the dissipative variables. Furthermore, for such a choice of β\beta the mirror and firehose instability are expected to quench any anisotropy of the shear stresses.

We present the result of these simulations in Fig. 7 at time t=3.4t=3.4. Specifically, we consider the case of δqB=δπB=105τπ\delta_{qB}=-\delta_{\pi B}=10^{5}\tau_{\pi}, in which the equations approach the Braginskii-like limit of extended magnetohydrodynamics Chandra et al. (2015). We also consider the same parameters, but supplemented with the subgrid model discussed above. Starting from the top, we show the pressure anisotropy |π0|=|PP|\left|\pi_{0}\right|=\left|P_{\parallel}-P_{\perp}\right|, where PP_{\parallel} and PP_{\perp} are the pressures along and perpendicular to the magnetic field. The scalar π0\pi_{0} was defined in Eq. (53). We can see that the global magnetic field evolution is very similar in all cases, except in the absence of the magnetic field coupling (δqB=δπB=0)(\delta_{qB}=\delta_{\pi B}=0), in which the vortices are much more collapsed (right panel). In the limit of extended magnetohydrodynamics the vortices are much more circular, although there are differences within the vortex, when comparing the purely extended magnetohydrodynamics case (left panel) with the subgrid modeling (middle panel). Most importantly, we can see that the shear stresses are highly anisotropic, especially in the case of extended magnetohydrodynamics (left). However, as expected from the firehose and mirror instabilities, the effective subgrid model eliminates this anisotropy almost entirely with the pressure anisotropy being about a 100-times smaller (middle panel).

In the middle row, we show the xyxy-component of the shear stress tensor πxy\pi^{xy}. Starting from the limit of extended magnetohydrodynamics with subgrid model (middle panel), we can see that the shear-stresses overall are 100100 times weaker than in the other cases, as expected from the discussion of |PP|\left|P_{\parallel}-P_{\perp}\right| above. Looking at the case of pure Extended magnetohydrodynamics (left panel), we can also see that small-scale substructures are present inside the vortices. Comparing the the structures in the magnetic field (top row), we can clearly see that shear stresses nicely align with the magnetic field geometry. When comparing with the case without the coupling (right panel), we can see that the stresses are uniformly present inside the vortices, where strongest shear flows are present. As expected for an enhancement of collisionality, the simulation including the subgrid model locally suppresses the shear stresses inside the vortices. The disappearance of substructure in the main vortices is also broadly consistent with the outcome of a similar Newtonian simulation assessing the impact of the subgrid model Berlok et al. (2020). Since the shear stresses are very small in this case, they might not be able to perfectly suppress the formation of secondary vortices, as in the isotropic viscous case, see Fig. 6. Indeed we see the (re-)appearance of secondary vortices when adding the sugrid model (middel panel).

Finally, we turn to the heat flux present in these simulations, which is shown in the bottom row of Fig. 7. We can see that in the case without magnetic field coupling (right panel), a strong heat flux is present along the shear layer. This heat flux is able to diffuse the part of the pressure support of the vortices helping the collapse of the vortex compared to the extended magnetohydrodynamics cases (left and center panels). Since for those the heat flux has to align with the magnetic field, no heat conduction across the shear layer can happen, as the magnetic field is aligned with the layer so that the vortices will retain an additional pressure support. Moreover, when considering the shock that propagates along the interface, we can see that in the unlimited case (left panel) a strong heat flux is present that travels with the shock. In the case of the subgrid model, where the heat flux is clamped, see Eq. (160), the transport of heat is suppressed as these low density regions do not permit strong heat fluxes.

VI Conclusions

In this work, we have presented a new formulation to numerically model relativitic dissipative effects in the presence of magnetic fields. More specifically, starting from a collisional 14-moment closure Denicol et al. (2012) for non-resistive relativistic magnetohydrodynamics Denicol et al. (2018), we have derived a set of equations suitable for the study of heat conduction as well as bulk and shear viscosities in an astrophysical context. By treating the fluid frame projector implicitly we were able to recast the equations in fully flux conservative form. Different from earlier approaches (e.g. Del Zanna et al. (2013); Bazow et al. (2018)), this allows us to treat all dissipative variables using high-resolution shock capturing methods, which removes any ambiguities of how to compute either explicit time or spatial derivatives in the dissipative sources.

In addition, the numerical scheme includes two coupling terms that project the shear stresses and heat flux onto the comoving magnetic fields. This coupling has been identified with the gyrofrequency, allowing us for the first time to fully include this effect in a relativistic magnetohydrodynamics simulations. We have further demonstrated that for large coupling the system approaches the relativistic Braginskii-like limit of extended magnetohydrodynamics investigated in Ref. Chandra et al. (2015). This allows us to study effects of weakly collisional plasmas that could be relevant for certain types of black hole accretion Foucart et al. (2016, 2017).

To investigate the properties of the solutions of the equations of motion, we have also presented a set of numerical simulations aimed at testing each of the dissipative effects individually. In particular, we have considered one- and two-dimensional tests of heat conduction and viscous shear flows in flat-spacetime. Our results showed that all effects can be correctly captured with or without (coupled) magnetic fields being present. As such, we expect that this scheme will be highly suitable to numerically study highly relativistic magnetohydrodynamical turbulence Zrake and MacFadyen (2012, 2013), black hole accretion Foucart et al. (2016), and dissipative effects in neutron star mergers Shibata et al. (2017); Alford et al. (2018); Most et al. (2021); Hammond et al. (2021).

An important aspect not yet addressed in this work are causality bounds on the system presented here. While conditions that ensure causality in the nonlinear regime of Israel-Stewart-like equations formulated in the Landau frame (in the absence of magnetic fields or baryon density) have been derived in Bemfica et al. (2021) (and strong hyperbolicity has been proven in Bemfica et al. (2019a) in the case of only bulk viscosity), no such conditions exist either when magnetic fields are included, or for the equations presented here, which have additional relaxation equations. Although our equations do approach standard Israel-Stewart-like equations in the stiff limit, we currently have no proof under which conditions the equations presented here maintain causality in the nonlinear regime (the linear regime of the theory derived in Denicol et al. (2018) was investigated in Biswas et al. (2020)). Such an analysis will have to be performed for our system as well, although our ability to solve this system in a stable manner under all conditions investigated here provides an optimistic indication that causality and hyperbolicity can be established in this system 333We stress at this point that a flux-conservative treatment requires the specification of an upper-limit of the system of characteristics, which we approximate with the speed of light. If the actual characteristics were much larger, the system would loose its upwind properties leading to numerical instabilities..

Beyond our current approach, several extensions are possible, in particular when considering applications to the hydrodynamic evolution of the quark-gluon plasma Romatschke and Romatschke (2019). In that regard, the next step would be to check how our framework handles Gubser flow Gubser (2010), which has become the standard test for numerical codes in the field of heavy-ion collisions Marrochio et al. (2015). In fact, when considering heavy-ion applications, we would have to also incorporate second-order coupling terms Denicol et al. (2012), though the effects of some of those terms on causality are not yet known in the nonlinear regime even at zero baryon chemical potential (for instance, Ref. Bemfica et al. (2021) neglected terms involving the coupling between the vorticity tensor and the shear stress tensor).

Furthermore, we have limited ourselves to non-resistive plasmas. A natural extension we are planning is the incorporation of resistivity effects, which have been investigated in Denicol et al. (2019). Extrapolating our results, this will allow us to correctly treat the relativistic Hall effect Zanotti and Dumbser (2011) and relativistic reconnection Bransgrove et al. (2021), which could be highly relevant for flaring process around accreting supermassive black holes Nathanail et al. (2020); Ripperda et al. (2020).

Acknowledgements

The authors thank Lev Arzamasskiy, Fabio Bemfica, Amitava Bhattarcharjee, Mani Chandra, Gabriel Denicol, Marcelo Disconzi, Charles Gammie, James Juno, Alex Pandya, Frans Pretorius, Alexander Philippov, Bart Ripperda and James Stone for insightful discussions and comments related to this work. ERM gratefully acknowledges support from postdoctoral fellowships at the Princeton Center for Theoretical Science, the Princeton Gravity Initiative, and the Institute for Advanced Study. JN is partially supported by the U.S. Department of Energy, Office of Science, Office for Nuclear Physics under Award No. DE-SC0021301. All simulations were performed on the Helios Cluster at the Institute for Advanced Study.

References

Appendix A Non-resistive dissipative magnetohydrodynamics in flat spacetime

The numerical tests presented in this paper are performed in flat Minkowski spacetime, where gμν=ημν=diag(1,1,1,1)g_{\mu\nu}=\eta_{\mu\nu}={\rm diag}\left(-1,1,1,1\right). We can further introduce a magnetic field

Bi=F0i,\displaystyle B^{i}=\,{{}^{\ast}\!F^{0i}}, (162)

as seen by a Eulerian observer. In terms of this magnetic field, the comoving field reads

bμ=(Bjuj,1u0[Bi+(Bjuj)ui]).\displaystyle b^{\mu}=\left(B_{j}u^{j},\frac{1}{u^{0}}\left[B^{i}+\left(B_{j}u^{j}\right)u^{i}\right]\right). (163)

The overall evolution equations of non-resistive dissipative magnetohydrodynamics in flat Minkowski spacetime are then given by

t[ρu0]+i[ρui]=0\displaystyle\partial_{t}\left[\rho u^{0}\right]+\partial_{i}\left[\rho u^{i}\right]=0 (164)
t[(ρh+b2+Π)(u0)2+(2q0ρ)u0(P+Π+12b2)b0b0+π00]\displaystyle\partial_{t}\left[\left(\rho h+b^{2}+\Pi\right)\left(u^{0}\right)^{2}+\left(2q^{0}-\rho\right)u^{0}-\left(P+\Pi+\frac{1}{2}b^{2}\right)-b^{0}b^{0}+\pi^{00}\right]
+i[(ρh+Π+b2)(u0)2uiu0+(q0ρ)u0uiu0+qiu0+π0ib0bi]=0\displaystyle\phantom{\partial_{t}\left[\left(\rho h+b^{2}+\Pi\right)\left(u^{0}\right)^{2}\right.}+\partial_{i}\left[\left(\rho h+\Pi+b^{2}\right)\left(u^{0}\right)^{2}\frac{u^{i}}{u^{0}}+\left(q^{0}-\rho\right)u^{0}\frac{u^{i}}{u^{0}}+q^{i}u^{0}+\pi^{0i}-b^{0}b^{i}\right]=0 (165)
t[(ρh+Π+b2)u0uj+qju0+q0uj+πj0b0bj]\displaystyle\partial_{t}\left[\left(\rho h+\Pi+b^{2}\right)u^{0}u_{j}+q_{j}u^{0}+q^{0}u_{j}+\pi^{0}_{j}-b^{0}b_{j}\right]
+i[(ρh+Π+b2)u0ujuiu0+qiuj+qjui+πji+δji(P+Π+12b2)bibj]=0\displaystyle\phantom{\partial_{t}\left[\left(\rho h+b^{2}+\Pi\right)\left(u^{0}\right)^{2}\right.}+\partial_{i}\left[\left(\rho h+\Pi+b^{2}\right)u^{0}u_{j}\frac{u^{i}}{u^{0}}+q^{i}u_{j}+q_{j}u^{i}+\pi^{i}_{j}+\delta^{i}_{j}\left(P+\Pi+\frac{1}{2}b^{2}\right)-b^{i}b_{j}\right]=0 (166)
tBj+i[uiu0Bjuju0Bi]=0\displaystyle\partial_{t}B^{j}+\partial_{i}\left[\frac{u^{i}}{u^{0}}B^{j}-\frac{u^{j}}{u^{0}}B^{i}\right]=0 (167)
t[(Π+Yζ)u0]+i[(Π+Yζ)u0uiu0]=1τΠΠ(1δΠΠτΠ)Π1ρθ+ζ,\displaystyle\partial_{t}\left[\left(\Pi+Y_{\zeta}\right)u^{0}\right]+\partial_{i}\left[\left(\Pi+Y_{\zeta}\right)u^{0}\frac{u^{i}}{u^{0}}\right]=-\frac{1}{\tau_{\Pi}}\Pi-\left(1-\frac{\delta_{\Pi\Pi}}{{\tau_{\Pi}}}\right)\Pi\frac{1}{\rho}\mathcal{I}_{\theta}+\mathcal{I}_{\zeta}, (168)
t[(qν+YκTuν)u0+YκTη0ν]+i[(qν+YκTuν)u0uiu0+YκTηiν]=\displaystyle\partial_{t}\left[\left(q^{\nu}+Y_{\kappa}Tu^{\nu}\right)u^{0}+Y_{\kappa}T\eta^{0\nu}\right]+\partial_{i}\left[\left(q^{\nu}+Y_{\kappa}Tu^{\nu}\right)u^{0}\frac{u^{i}}{u^{0}}+Y_{\kappa}T\eta^{i\nu}\right]=
(1δqqτq)qν1ρθ+Tκν1τqqν1ωq(qμuμ)uνδqBτqFνμqμ\displaystyle\phantom{\partial_{t}\left[\left(q^{\nu}+Y_{\kappa}Tu^{\nu}\right)u^{0}+Y_{\kappa}T\eta^{0\nu}\right]}-\left(1-\frac{\delta_{qq}}{\tau_{q}}\right)q^{\nu}\,\frac{1}{\rho}\mathcal{I}_{\theta}+T\mathcal{I}_{\kappa}^{\nu}-\frac{1}{\tau_{q}}q^{\nu}-\frac{1}{\omega_{q}}\left(q_{\mu}u^{\mu}\right)u^{\nu}-\frac{\delta_{qB}}{\tau_{q}}F^{\nu\mu}q_{\mu} (169)
t[παβu0+Yη(η0αuβ+η0βuα)]+i[παβu0uiu0+Yη(ηiαuβ+ηiβuα)]=δπBτπFδμπμγΔ~γδαβ1τππαβ\displaystyle\partial_{t}\left[\pi^{\alpha\beta}u^{0}+Y_{\eta}\left(\eta^{0\alpha}u^{\beta}+\eta^{0\beta}u^{\alpha}\right)\right]+\partial_{i}\left[\pi^{\alpha\beta}u^{0}\frac{u^{i}}{u^{0}}+Y_{\eta}\left(\eta^{i\alpha}u^{\beta}+\eta^{i\beta}u^{\alpha}\right)\right]=-\frac{\delta_{\pi B}}{\tau_{\pi}}F^{\delta\mu}\pi^{\gamma}_{\mu}\,\tilde{\Delta}^{\alpha\beta}_{\gamma\delta}-\frac{1}{\tau_{\pi}}\pi^{\alpha\beta}
(1δππτπ)παβ1ρθ1ωπ[(παλuλ)uβ+(πβλuλ)uα]1ωππ(πμνημν)ηαβ.\displaystyle\phantom{\partial_{t}\left[\pi^{\alpha\beta}u^{0}+\frac{\eta}{\tau_{\pi}}\left(\eta^{0\alpha}u^{\beta}+\eta^{0\beta}u^{\alpha}\right)\right]}-\left(1-\frac{\delta_{\pi\pi}}{\tau_{\pi}}\right)\pi^{\alpha\beta}\,\frac{1}{\rho}\mathcal{I}_{\theta}-\frac{1}{\omega_{\pi}}\left[\left(\pi^{\alpha\lambda}u_{\lambda}\right)u^{\beta}+\left(\pi^{\beta\lambda}u_{\lambda}\right)u^{\alpha}\right]-\frac{1}{\omega_{\pi\pi}}\left(\pi^{\mu\nu}\eta_{\mu\nu}\right)\eta^{\alpha\beta}. (170)
t(ρYζu0)+i(ρYζu0uiu0)=ρζ,\displaystyle\partial_{t}\left(\rho Y_{\zeta}u^{0}\right)+\partial_{i}\left(\rho Y_{\zeta}u^{0}\frac{u^{i}}{u^{0}}\right)=\rho\mathcal{I}_{\zeta}, (171)
t(ρYκu0)+i(ρYκu0uiu0)=ρκ,\displaystyle\partial_{t}\left(\rho Y_{\kappa}u^{0}\right)+\partial_{i}\left(\rho Y_{\kappa}u^{0}\frac{u^{i}}{u^{0}}\right)=\rho\mathcal{I}_{\kappa}, (172)
t(ρYηu0)+i(ρYηu0uiu0)=ρη,\displaystyle\partial_{t}\left(\rho Y_{\eta}u^{0}\right)+\partial_{i}\left(\rho Y_{\eta}u^{0}\frac{u^{i}}{u^{0}}\right)=\rho\mathcal{I}_{\eta}, (173)
t(ρYθu0)+i(ρYθu0uiu0)=ρθ,\displaystyle\partial_{t}\left(\rho Y_{\theta}u^{0}\right)+\partial_{i}\left(\rho Y_{\theta}u^{0}\frac{u^{i}}{u^{0}}\right)=\rho\mathcal{I}_{\theta}, (174)
t(Zνκ)+i(Yκδνi)=νκ.\displaystyle\partial_{t}\left(Z^{\kappa}_{\nu}\right)+\partial_{i}\left(Y_{\kappa}\delta^{i}_{\nu}\right)=\mathcal{I}^{\kappa}_{\nu}. (175)