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

Anisotropic fluid dynamical simulations of heavy-ion collisions

Mike McNelis Dennis Bazow Ulrich Heinz Department of Physics, The Ohio State University, Columbus, OH 43210-1117, USA Institut für Theoretische Physik, J. W. Goethe Universität, Max-von-Laue-Str. 1, D-60438 Frankfurt am Main, Germany
Abstract

We present VAH, a (3+1)–dimensional simulation that evolves the far-from-equilibrium quark-gluon plasma produced in ultrarelativistic heavy-ion collisions with anisotropic fluid dynamics. We solve the hydrodynamic equations on an Eulerian grid using the Kurganov–Tadmor algorithm in combination with a new adaptive Runge–Kutta method. Our numerical scheme allows us to start the simulation soon after the nuclear collision, largely avoiding the need to integrate it with a separate pre-equilibrium dynamics module. We test the code’s performance by simulating on the Eulerian grid conformal and non-conformal Bjorken flow as well as conformal Gubser flow, whose (0+1)–dimensional solutions are precisely known. Finally, we compare non-conformal anisotropic hydrodynamics to second-order viscous hydrodynamics in central Pb+Pb collisions and find that the former’s longitudinal flow profile responds more consistently to the fluid’s gradients along the spacetime rapidity direction.

keywords:
Ultrarelativistic heavy-ion collisions , quark-gluon plasma , relativistic hydrodynamics , computational fluid dynamics
journal: Computer Physics Communications

PROGRAM SUMMARY
Manuscript Title: Anisotropic fluid dynamical simulations of heavy-ion collisions
Authors: Mike McNelis, Dennis Bazow, Ulrich Heinz
Program Title: VAH 
Licensing provisions: GPLv3
Programming Language: C++
Computer: Laptop, desktop, cluster
Operating System: GNU/Linux distributions, Mac OS X
Global memory usage: 1.2 GB (for a 129×129×63129\times 129\times 63 grid)
Keywords: Ultrarelativistic heavy-ion collisions, quark-gluon plasma, relativistic hydrodynamics, computational fluid dynamics
Classification: 12 Gases and Fluids, 17 Nuclear physics
External routines/libraries: GNU Scientific Library (GSL)
Nature of problem:
Modeling the far-from-equilibrium dynamics of quark-gluon plasma produced in ultrarelativistic heavy-ion collisions.
Solution method:
Kurganov–Tadmor algorithm, adaptive stepsize method
Running time:
A (3+1)–d non-conformal anisotropic fluid dynamical simulation of a central Pb+Pb collision on a 129×129×63129\times 129\times 63 grid takes about 530ss for an Intel Xeon E5-2680 v4 multi-core processor with OpenMP acceleration (see Sec. 5.3).

1 Introduction

The quark-gluon plasma is one of the most extreme phases of matter in our universe. The incredibly high temperatures and densities required to create it (T1012T\sim 10^{12} K, ρ1017\rho\sim 10^{17} kg/m3) are orders of magnitude beyond those of any substance typically encountered in nature. The quark-gluon plasma also has the remarkable property of having the lowest shear viscosity to entropy density ratio of any known fluid; current phenomenological constraints put η/𝒮0.1\eta/\mathcal{S}\sim 0.1 at temperatures around the deconfinement temperature Tc=154T_{c}=154 MeV [1, 2]. This value is very close to the conjectured Kovtun-Son-Starinet bound η/𝒮 1/4π\eta/\mathcal{S}{\,\geq\,}1/4\pi, which follows from the AdS/CFT correspondence in string theory [3, 4]. Therefore, it is of great interest to the physics community to test whether or not the KSS bound holds for such strongly coupled liquids. Today, we can produce quark-gluon plasma by colliding beams of heavy ions (e.g. Pb+Pb) at very high energies. The femtoscopically small droplets of plasma formed in these ultrarelativistic nuclear collisions have incredibly short lifetimes (τf1023\tau_{f}\sim 10^{-23} s), making them difficult to probe directly. To reconstruct the medium’s transport properties, one can simulate the various stages of a heavy-ion collision with quantitative precision and predictive power and fit the theoretical model to large sets of soft-momentum (pT<3p_{T}<3 GeV) hadronic observables [5, 6, 7, 1, 2, 8, 9].111High-energy jets, heavy quarks and electromagnetic radiation can also serve as experimental probes but their full integration into hybrid models [10, 11, 12, 13, 14, 15] is more involved than for soft-momentum hadrons. The hybrid model must be accurate enough that it can make full use of the considerable precision of the available experimental data.

The anisotropically expanding quark-gluon plasma stage is usually modeled with relativistic viscous hydrodynamics [16, 17, 18, 5, 19, 20]. Conventionally, viscous hydrodynamics only works for fluids that are both near local equilibrium (i.e. inverse Reynolds number Re11\text{Re}^{-1}\ll 1) and have small spacetime gradients (i.e. Knudsen number Kn 1\ll 1) [18, 21]. But unlike most fluids found in nature, the quark-gluon plasma is initially far from local-equilibrium (Re11\text{Re}^{-1}\sim 1) and sustains moderately large gradients (Kn 1\sim 1) for a good portion of its lifetime. This is primarily due to quantum fluctuations in the initial-state profile [22, 23, 24] and the rapid longitudinal expansion at early times. Under these extreme conditions, the validity of a fluid dynamical description for the quark-gluon plasma comes into question [25]. Nevertheless, second-order viscous hydrodynamic simulations, in conjunction with other multi-stage modules, have been widely successful at reproducing hadronic observables (e.g. anisotropic flow coefficients vnv_{n}[26, 27]. State-of-the-art hybrid models now have enough predictive power to quantitatively constrain the shear and bulk viscosities of the quark-gluon plasma using heavy-ion experimental data and Bayesian inference [6, 7, 1, 2, 8, 9]. Hydrodynamics has also been found to be applicable in collisions between light and heavy ions (e.g. He3+Au, p+Pb) and in proton–proton collisions [12, 28], although the origin of collectivity in these small systems is still being debated [29, 30, 31, 32].

The success of second-order viscous hydrodynamics can be attributed to an underlying resummed hydrodynamic theory that also extends to large gradients [33, 34, 35]. As the associated non-hydrodynamic modes decay on microscopic time scales τrτhydro\tau_{r}{\,\ll\,}\tau_{\text{hydro}},222Hydrodynamic simulations often employ relaxation-type equations from relativistic kinetic theory (e.g. DNMR [18]), but they likely do not precisely capture the transient dynamics in strongly coupled fluids [36, 37]. the system approaches a (generally non-equilibrium) hydrodynamic attractor that is well approximated by viscous hydrodynamics. This happens within the time scale333The hydrodynamization time τhydro\tau_{\text{hydro}} refers to the time when viscous hydrodynamics becomes applicable, but the hydrodynamic simulation starting time τ0\tau_{0} can be different from this. τhydro1\tau_{\text{hydro}}\sim 1 fm/cc, long before the system thermalizes [38, 39, 40]. Whether or not the strongly coupled quark-gluon plasma possesses an attractor at presently experimentally accessible temperatures (T0.150.5T\sim 0.15-0.5 GeV) is currently unknown. A major obstacle to answering this question are the technical difficulties involved in evaluating its transport properties from first principles, including its shear and bulk viscosities [41]. Recently, it was demonstrated that resummed hydrodynamics can be obtained by expanding the microscopic Green’s function of a linearized kinetic system, an alternative to a direct resummation of the gradient expansion [42]. It might be possible to extend this concept to the non-equilibrium quark-gluon plasma by perturbing the system around local-equilibrium and expanding the resulting correlation function(s) [43, 44, 45, 46].

Despite the robustness of second-order viscous hydrodynamics, it is still susceptible to breaking down when gradients are very large (Kn 1\gg 1) [47, 48, 20]. The largest gradients in heavy-ion collisions occur at very early times (τ< 0.2\tau{\,<\,}0.2 fm/cc), especially around the edges of the fireball. If one starts the viscous hydrodynamic simulation too early, one usually encounters negative longitudinal pressures 𝒫L\mathcal{P}_{L} due to huge pressure anisotropies [49]; near the edges of the fireball even the transverse pressure 𝒫\mathcal{P}_{\perp} can turn negative if the bulk viscosity ζ/𝒮\zeta/\mathcal{S} peaks strongly near the quark-hadron phase transition. Not only does this cause excessive viscous heating444Viscous heating refers to internal entropy production by dissipative processes, not due to external thermal sources. but large negative pressures also redirect the matter inward, potentially resulting in cavitation. Regulation schemes can be used to tamp down the viscous pressures and prevent the simulation from crashing, but they cloud the physical predictions of the original hydrodynamic theory [5, 20].555Some regulation schemes are less extreme than others. For example, the hydrodynamic code MUSIC only regulates the dissipative currents outside the fireball region [16] while iEBE-VISHNU regulates the entire grid [5]. Because of the technical issues involved, it is more suitable to use a pre-equilibrium dynamics model before transitioning to viscous hydrodynamics at τ0=τhydro\tau_{0}=\tau_{\text{hydro}} [22, 26, 50, 6, 7, 1, 2, 44, 51, 45, 52]. Still, the conformal approximation usually made in the former results in a mismatch to the latter’s non-conformal equation of state, producing (in spite of the system’s expansion) artificially positive bulk viscous pressures that can be as large as Π𝒫eq\Pi\sim\mathcal{P}_{\text{eq}} around the edges of the fireball [53]. In some situations, one cannot even switch between dynamical models instantaneously. For example, in low-energy collisions (sNN1050\sqrt{s_{\text{NN}}}\sim 10-50 GeV) where the nuclear interpenetration time is comparable to the fireball lifetime, one needs to run viscous hydrodynamics in the background as soon as the participant nucleons start feeding thermal energy and net-baryon number into the newly formed fireball [54, 55, 56, 57, 58].

The shortcomings of second-order viscous hydrodynamics have motivated the development of hydrodynamic models that can better handle far-from-equilibrium situations [59, 60, 61]. The most promising candidate is anisotropic hydrodynamics, which treats the two largest dissipative effects arising in heavy-ion collisions (the pressure anisotropy 𝒫L𝒫\mathcal{P}_{L}-\mathcal{P}_{\perp} and the bulk viscous pressure Π\Pi) non-perturbatively [62, 63, 64, 65, 66]. While current formulations of anisotropic hydrodynamics are partially based on weakly coupled kinetic theory,666Second-order viscous hydrodynamic simulations also rely on microscopic approaches such as relativistic kinetic theory to compute the relaxation times and other second-order transport coefficients [18, 67, 68]. they are able to capture exact kinetic solutions more accurately than standard viscous hydrodynamics [47, 62, 63, 64, 69, 70, 48]. In addition, their ability to maintain positive longitudinal and transverse pressures makes them less prone to cavitation in (3+1)–dimensional simulations [49]. This means they can run at early times with little interference from viscous regulations and remain numerically stable.

The application of anisotropic hydrodynamics to heavy-ion phenomenology is still relatively new; Au+Au collisions at RHIC (sNN=200\sqrt{s_{\text{NN}}}=200 GeV) and Pb+Pb collisions at the LHC (sNN=2.76\sqrt{s_{\text{NN}}}=2.76 and 5.02 TeV) have been modeled reasonably well but so far only using smooth initial conditions and a few model parameters to adjust the fit to experimental data [71, 72, 73]. Anisotropic hydrodynamics has yet to undergo the same extensive tests and trials as viscous hydrodynamics, but it can potentially serve as an additional discrete model for the fluid dynamical stage of heavy-ion collisions. This would further increase the flexibility of the Bayesian framework developed by the JETSCAPE collaboration [1, 2], which has already incorporated a number of discrete models for the particlization stage [74]. The hope is that, by controlling large dissipative flows nonperturbatively, anisotropic hydrodynamics can constrain the transport coefficients of QCD matter more accurately.

In this paper we introduce our (3+1)–dimensional anisotropic fluid dynamical simulation for heavy-ion collisions called VAH.777The code package can be downloaded from the GitHub repository https://github.com/mjmcnelis/cpu_vah. The C++ module is based on the GPU–accelerated viscous hydrodynamic code GPU VH [20, 49], except that it implements anisotropic hydrodynamics with the (𝒫L\mathcal{P}_{L}, 𝒫\mathcal{P}_{\perp})–matching scheme developed in Ref. [66].888At this point VAH itself has not yet been ported to GPUs. We evolve the dynamical equations on an Eulerian grid using the Kurganov–Tadmor algorithm [75], a popular method also used in other viscous hydrodynamic codes [16, 17, 20, 76]. The main improvement in our code is the ability to automatically adjust the time step Δτn\Delta\tau_{n} after each iteration, as opposed to using a fixed value for the entire simulation. Our adaptive Runge–Kutta scheme initially uses a fine time step to resolve the rapid longitudinal expansion at early times while speeding up the evolution at later times with a coarser time step given by the Courant-Friedrichs-Lewy (CFL) condition. As a result, we can start anisotropic hydrodynamics at a very early time τ0=0.05\tau_{0}=0.05 fm/cc to model both the far-off-equilibrium dynamics stage999We assume the system is longitudinally free-streaming (i.e. 𝒫L/𝒫eq0\mathcal{P}_{L}/\mathcal{P}_{\text{eq}}\approx 0 and 1/τ\mathcal{E}\propto 1/\tau) in the time interval 0<ττ00<\tau\leq\tau_{0} before starting anisotropic hydrodynamics. This closely mirrors the situation found in other pre-hydrodynamic models [22, 50, 45]. It has also been argued [77, 78] that (at least in weakly coupled systems) for τ0\tau\to 0 the far-off-equilibrium hydrodynamic attractor approaches the free-streaming attractor in Bjorken flow (which approximates the early evolution stage in heavy-ion collisions). with a non-conformal QCD equation of state and smoothly transition to viscous hydrodynamics.101010Anisotropic hydrodynamics reduces to second-order viscous hydrodynamics in the limit of small Knudsen and inverse Reynolds numbers (Kn,Re11\text{Kn},\text{Re}^{-1}\ll 1). We do not switch to a separate viscous hydrodynamics model as the Knudsen and inverse Reynolds numbers decrease, but use anisotropic hydrodynamics to evolve both the earliest far-off-equilibrium and subsequent viscous hydrodynamic stages.

The code package contains other useful features, such as the option to run either anisotropic hydrodynamics or second-order viscous hydrodynamics within the same framework. This provides the user with several discrete models to evolve the fluid dynamical stage for comparison. We further implemented user options to accelerate the simulation on a multi-core processor with OpenMP, and to use automated grid settings that optimize the overall size of the spatial grid for a given set of runtime parameters (e.g. the impact parameter and particlization switching temperature).

This paper is organized as follows: In Sec. 2 we review the anisotropic hydrodynamic equations used to evolve the energy-momentum tensor of the quark-gluon plasma. Sec. 3 discusses the numerical implementation of these dynamical equations in the code. In Sec. 4 we test our anisotropic fluid dynamical simulation for a system subject to (non)conformal Bjorken flow and conformal Gubser flow. Furthermore, we compare anisotropic hydrodynamics to second-order viscous hydrodynamics in (3+1)–dimensions. Finally, we benchmark the typical runtimes of (2+1)–d and (3+1)–d non-conformal hydrodynamic simulations in Sec. 5.

In this work we use Milne spacetime coordinates xμ=(τ,x,y,ηs)x^{\mu}=(\tau,x,y,\eta_{s}), where τ=t2z2\tau=\sqrt{t^{2}{-}z^{2}} is the longitudinal proper time and ηs=tanh1(z/t)\eta_{s}=\tanh^{-1}(z/t) is the spacetime rapidity. We adopt the mostly-minus convention for the metric tensor, gμν=diag(1,1,1,1/τ2)g^{\mu\nu}=\mathrm{diag}(1,-1,-1,-1/\tau^{2}). In the code, we solve the hydrodynamic equations in natural units =c=kB=1\hbar=c=k_{B}=1; energy, momentum and temperature have units of fm-1, energy density and pressure have units of fm-4, etc. These units are converted to physical units (e.g. energy densities in GeV/fm3) when outputting and plotting the results. The net-baryon density nBn_{B} and baryon diffusion current VBμV_{B}^{\mu} are neglected in this version of the code.

2 Anisotropic fluid dynamics

In this section we provide an overview of the anisotropic hydrodynamics equations, including the equation of state and transport coefficients, to be implemented in the code that evolves the energy-momentum tensor Tμν(x)T^{\mu\nu}(x) of the quark-gluon plasma. For more details on the derivation of these hydrodynamic equations we refer the reader to Refs. [66, 65].

2.1 Energy-momentum tensor

In anisotropic fluid dynamics, it is convenient to decompose Tμν(x)T^{\mu\nu}(x) in the basis {uμ(x)u^{\mu}(x), zμ(x)z^{\mu}(x), Ξμν(x)\Xi^{\mu\nu}(x)}, where the fluid velocity uμu^{\mu} represents the temporal direction in the local fluid rest frame (LRF) and the spatial vectors are split into the longitudinal direction zμz^{\mu} and the transverse projection tensor Ξμν=gμνuμuν+zμzν\Xi^{\mu\nu}=g^{\mu\nu}-u^{\mu}u^{\nu}+z^{\mu}z^{\nu} [65]. In the Landau frame this results in the decomposition (suppressing the spacetime dependence)

Tμν=uμuν+𝒫Lzμzν𝒫Ξμν+2Wz(μzν)+πμν,T^{\mu\nu}=\mathcal{E}u^{\mu}u^{\nu}+\mathcal{P}_{L}z^{\mu}z^{\nu}-\mathcal{P}_{\perp}\Xi^{\mu\nu}+2W^{(\mu}_{\perp z}z^{\nu)}+\pi_{\perp}^{{\mu\nu}}\,, (1)

where round parentheses denote symmetrization, i.e. Wz(μzν)=12(Wzμzν+Wzνzμ)W^{(\mu}_{\perp z}z^{\nu)}=\frac{1}{2}(W_{\perp z}^{\mu}z^{\nu}+W_{\perp z}^{\nu}z^{\mu}). The major components of TμνT^{\mu\nu} are the LRF energy density =uμuνTμν\mathcal{E}=u_{\mu}u_{\nu}T^{\mu\nu}, the longitudinal pressure 𝒫L=zμzνTμν\mathcal{P}_{L}=z_{\mu}z_{\nu}T^{\mu\nu} and the transverse pressure 𝒫=12ΞμνTμν\mathcal{P}_{\perp}=-\frac{1}{2}\Xi_{\mu\nu}T^{\mu\nu}. Together, the pressures 𝒫L\mathcal{P}_{L} and 𝒫\mathcal{P}_{\perp} capture the largest dissipative flows in heavy-ion collisions: the pressure anisotropy Δ𝒫=𝒫L𝒫\Delta\mathcal{P}=\mathcal{P}_{L}-\mathcal{P}_{\perp} caused by the rapid longitudinal expansion rate at early longitudinal proper times, and the bulk viscous pressure111111Here 𝒫¯=13(𝒫L+2𝒫)\bar{\mathcal{P}}=\frac{1}{3}(\mathcal{P}_{L}+2\mathcal{P}_{\perp}) is the average (isotropic) pressure and 𝒫eq=𝒫eq()\mathcal{P}_{\text{eq}}=\mathcal{P}_{\text{eq}}(\mathcal{E}) is the equilibrium pressure. Π=𝒫¯𝒫eq\Pi=\bar{\mathcal{P}}-\mathcal{P}_{\text{eq}} due to critical fluctuations near the quark-hadron phase transition. The remaining components of TμνT^{\mu\nu} are the longitudinal momentum diffusion current Wzμ=ΞαμzνTανW_{\perp z}^{\mu}=-\Xi^{\mu}_{\alpha}z_{\nu}T^{\alpha\nu} and the transverse shear stress tensor πμν=ΞαβμνTαβ\pi_{\perp}^{{\mu\nu}}=\Xi_{\alpha\beta}^{\mu\nu}T^{\alpha\beta}, with Ξαβμν=12(ΞαμΞβν+ΞβνΞαμΞμνΞαβ)\Xi^{\mu\nu}_{\alpha\beta}=\frac{1}{2}(\Xi^{\mu}_{\alpha}\Xi^{\nu}_{\beta}+\Xi^{\nu}_{\beta}\Xi^{\mu}_{\alpha}-\Xi^{\mu\nu}\Xi_{\alpha\beta}) being the traceless double transverse projector. We refer to these components as residual shear stresses since they are typically smaller than the pressure anisotropy term in the full shear stress tensor

πμν=13(𝒫L𝒫)(2zμzν+Ξμν)+2Wz(μzν)+πμν.\pi^{\mu\nu}=\frac{1}{3}(\mathcal{P}_{L}{-}\mathcal{P}_{\perp})(2z^{\mu}z^{\nu}{+}\Xi^{\mu\nu})+2W^{(\mu}_{\perp z}z^{\nu)}+\pi_{\perp}^{{\mu\nu}}\,. (2)

2.2 Dynamical variables

The dynamical variables that we propagate in the code are

𝒒=(Tτμ,𝒫L,𝒫,Wzμ,πμν).\bm{q}=(T^{\tau\mu},\mathcal{P}_{L},\mathcal{P}_{\perp},W_{\perp z}^{\mu},\pi_{\perp}^{{\mu\nu}})\,. (3)

Their evolution equations will be discussed below. Although WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}} each have only two independent components, we evolve all 14 of their components independently to simplify the workflow of the algorithm.121212When propagating these extraneous components numerically, slight violations of the orthogonality and tracelessness conditions (74) can occur. We correct for these errors at each step of the simulation with the regulation scheme described in Sec. 3.5. In addition, we propagate the energy density \mathcal{E} and the fluid velocity’s spatial components131313The fluid velocity’s temporal component is uτ=1+(ux)2+(uy)2+(τuη)2u^{\tau}=\sqrt{1+(u^{x})^{2}+(u^{y})^{2}+(\tau u^{\eta})^{2}}. 𝒖=\bm{u}= (uxu^{x}, uyu^{y}, uηu^{\eta}) since they appear in the hydrodynamic equations. They are inferred from the components TτμT^{\tau\mu}:

Tτμ=uτuμ+𝒫Lzτzμ𝒫Ξτμ+2Wz(τzμ)+πτμ.T^{\tau\mu}=\mathcal{E}u^{\tau}u^{\mu}+\mathcal{P}_{L}z^{\tau}z^{\mu}-\mathcal{P}_{\perp}\Xi^{\tau\mu}+2W^{(\tau}_{\perp z}z^{\mu)}+\pi_{\perp}^{\tau\mu}\,. (4)

To solve these equations, one also needs to know zμz^{\mu}. From the orthogonality conditions zμzμ=1z_{\mu}z^{\mu}=-1 and zμuμ=0z_{\mu}u^{\mu}=0, there are only two nonzero components that depend on the fluid velocity:

zμ=11+u2(τuη,0,0,uττ),z^{\mu}=\frac{1}{\sqrt{1{+}u_{\perp}^{2}}}\left(\tau u^{\eta},0,0,\frac{u^{\tau}}{\tau}\right)\,, (5)

where u=(ux)2+(uy)2u_{\perp}=\sqrt{(u^{x})^{2}+(u^{y})^{2}} is the transverse velocity. The solution of the inferred variables (\mathcal{E}, 𝒖\bm{u}) from the algebraic equations (4) will be discussed in Sec. 3.3.

For (3+1)–dimensionally expanding fluids, we have a total of 2020 dynamical variables and four inferred variables to evolve on an Eulerian grid.141414Dynamical variables are evolved directly using the Kurganov–Tadmor algorithm (see Sec. 3.1). Inferred variables are determined from the dynamical variables algebraically.,151515For anisotropic fluid dynamics with a QCD equation of state [66], we further evolve the mean-field BB as a dynamical variable and the anisotropic variables (Λ\Lambda, αL\alpha_{L}, α\alpha_{\perp}) as additional inferred variables (see Secs. 2.4 and 2.6.3). If the system is longitudinally boost-invariant, we do not need to propagate the components TτηT^{\tau\eta}, WzμW_{\perp z}^{\mu}, πμη\pi_{\perp}^{\mu\eta} and uηu^{\eta} since they vanish by symmetry.

2.3 Conservation laws

The evolution of the components TτμT^{\tau\mu} is given by the energy-momentum conservation laws

DμTμν=0,D_{\mu}T^{\mu\nu}=0\,, (6)

where the covariant derivative DμD_{\mu} accounts for the curvilinear nature of the Milne coordinates. The conservation equations can be expanded as

μTμν+ΓμλμTλν+ΓμλνTλμ=0,\partial_{\mu}T^{\mu\nu}+\Gamma^{\mu}_{\mu\lambda}T^{\lambda\nu}+\Gamma^{\nu}_{\mu\lambda}T^{\lambda\mu}=0\,, (7)

where μ\partial_{\mu} is the partial derivative and Γνλμ\Gamma^{\mu}_{\nu\lambda} are the Christoffel symbols. In Milne spacetime, the only nonzero Christoffel symbols are

Γηητ=τ,\displaystyle\Gamma^{\tau}_{\eta\eta}=\tau,\indent Γτηη=Γητη=1τ.\displaystyle\indent\Gamma^{\eta}_{\tau\eta}=\Gamma^{\eta}_{\eta\tau}=\frac{1}{\tau}\,. (8)

Thus, the set of equations (7) can be rewritten as

τTττ+iTτi\displaystyle\partial_{\tau}T^{\tau\tau}+\partial_{i}T^{\tau i} =Tττ+τ2Tηητ,\displaystyle=-\frac{T^{\tau\tau}{+}\tau^{2}T^{\eta\eta}}{\tau}\,, (9a)
τTτx+jTxj\displaystyle\partial_{\tau}T^{\tau x}+\partial_{j}T^{xj} =Tτxτ,\displaystyle=-\frac{T^{\tau x}}{\tau}\,, (9b)
τTτy+jTyj\displaystyle\partial_{\tau}T^{\tau y}+\partial_{j}T^{yj} =Tτyτ,\displaystyle=-\frac{T^{\tau y}}{\tau}\,, (9c)
τTτη+jTηj\displaystyle\partial_{\tau}T^{\tau\eta}+\partial_{j}T^{\eta j} =3Tτητ,\displaystyle=-\frac{3T^{\tau\eta}}{\tau}\,, (9d)

where the Latin indices (i,j)(x,y,η)(i,j)\in(x,y,\eta) are summed over spatial components. We can eliminate the components TτiT^{\tau i} in (9a) by using the identity:

Tτi=(+𝒫)uτui+τi+𝒲τi+πτi=viTττ+vi(𝒫ττ𝒲ττπττ)+τi+𝒲τi+πτi.\begin{split}T^{\tau i}&=(\mathcal{E}{+}\mathcal{P}_{\perp})u^{\tau}u^{i}+\mathcal{L}^{\tau i}+\mathcal{W}^{\tau i}+\pi_{\perp}^{\tau i}\\ &=v^{i}T^{\tau\tau}+v^{i}(\mathcal{P}_{\perp}{-}\mathcal{L}^{\tau\tau}{-}\mathcal{W}^{\tau\tau}{-}\pi_{\perp}^{\tau\tau})+\mathcal{L}^{\tau i}+\mathcal{W}^{\tau i}+\pi_{\perp}^{\tau i}\,.\end{split} (10)

Here we introduced the three-velocity vi=ui/uτv^{i}=u^{i}/u^{\tau} as well as the tensors μν=Δ𝒫zμzν\mathcal{L}^{\mu\nu}=\Delta\mathcal{P}z^{\mu}z^{\nu} and 𝒲μν=2Wz(μzν)\mathcal{W}^{\mu\nu}=2W^{(\mu}_{\perp z}z^{\nu)}. Likewise, we can express the components TijT^{ij} in (9b-d) in terms of TτiT^{\tau i}:

Tij=(+𝒫)uiuj𝒫gij+ij+𝒲ij+πij=vjTτivj(τi+𝒲τi+πτi)𝒫gij+ij+𝒲ij+πij.\begin{split}T^{ij}=&\,(\mathcal{E}{+}\mathcal{P}_{\perp})u^{i}u^{j}-\mathcal{P}_{\perp}g^{ij}+\mathcal{L}^{ij}+\mathcal{W}^{ij}+\pi_{\perp}^{ij}\\ =&\,v^{j}T^{\tau i}-v^{j}(\mathcal{L}^{\tau i}{+}\mathcal{W}^{\tau i}{+}\pi_{\perp}^{\tau i})-\mathcal{P}_{\perp}g^{ij}+\mathcal{L}^{ij}+\mathcal{W}^{ij}+\pi_{\perp}^{ij}\,.\end{split} (11)

After some algebra one obtains [49]

τTττ+i(viTττ)=\displaystyle\partial_{\tau}T^{\tau\tau}+\partial_{i}(v^{i}T^{\tau\tau})= Tττ+τ2Tηητ+(ττ+𝒲ττ+πττ𝒫)ivi\displaystyle-\frac{T^{\tau\tau}{+}\tau^{2}T^{\eta\eta}}{\tau}+(\mathcal{L}^{\tau\tau}{+}\mathcal{W}^{\tau\tau}{+}\pi_{\perp}^{\tau\tau}{-}\mathcal{P}_{\perp})\partial_{i}v^{i} (12a)
+vii(ττ+𝒲ττ+πττ𝒫)ητηi𝒲τiiπτi,\displaystyle+v^{i}\partial_{i}(\mathcal{L}^{\tau\tau}{+}\mathcal{W}^{\tau\tau}{+}\pi_{\perp}^{\tau\tau}{-}\mathcal{P}_{\perp})-\partial_{\eta}\mathcal{L}^{\tau\eta}-\partial_{i}\mathcal{W}^{\tau i}-\partial_{i}\pi_{\perp}^{\tau i}\,,\quad
τTτx+i(viTτx)=\displaystyle\partial_{\tau}T^{\tau x}+\partial_{i}(v^{i}T^{\tau x})= Tτxτx𝒫+(𝒲τx+πτx)ivi\displaystyle-\frac{T^{\tau x}}{\tau}-\partial_{x}\mathcal{P}_{\perp}+(\mathcal{W}^{\tau x}{+}\pi_{\perp}^{\tau x})\partial_{i}v^{i} (12b)
+vii(𝒲τx+πτx)η𝒲xηiπxi,\displaystyle+v^{i}\partial_{i}(\mathcal{W}^{\tau x}{+}\pi_{\perp}^{\tau x})-\partial_{\eta}\mathcal{W}^{x\eta}-\partial_{i}\pi_{\perp}^{xi}\,,
τTτy+i(viTτy)=\displaystyle\partial_{\tau}T^{\tau y}+\partial_{i}(v^{i}T^{\tau y})= Tτyτy𝒫+(𝒲τy+πτy)ivi\displaystyle-\frac{T^{\tau y}}{\tau}-\partial_{y}\mathcal{P}_{\perp}+(\mathcal{W}^{\tau y}{+}\pi_{\perp}^{\tau y})\partial_{i}v^{i} (12c)
+vii(𝒲τy+πτy)η𝒲yηiπyi,\displaystyle+v^{i}\partial_{i}(\mathcal{W}^{\tau y}{+}\pi_{\perp}^{\tau y})-\partial_{\eta}\mathcal{W}^{y\eta}-\partial_{i}\pi_{\perp}^{yi}\,,
τTτη+i(viTτη)=\displaystyle\partial_{\tau}T^{\tau\eta}+\partial_{i}(v^{i}T^{\tau\eta})= 3Tτητη𝒫τ2+(τη+𝒲τη+πτη)ivi\displaystyle-\frac{3T^{\tau\eta}}{\tau}-\frac{\partial_{\eta}\mathcal{P}_{\perp}}{\tau^{2}}+(\mathcal{L}^{\tau\eta}{+}\mathcal{W}^{\tau\eta}{+}\pi_{\perp}^{\tau\eta})\partial_{i}v^{i} (12d)
+vii(τη+𝒲τη+πτη)ηηηi𝒲ηiiπηi.\displaystyle+v^{i}\partial_{i}(\mathcal{L}^{\tau\eta}{+}\mathcal{W}^{\tau\eta}{+}\pi_{\perp}^{\tau\eta})-\partial_{\eta}\mathcal{L}^{\eta\eta}-\partial_{i}\mathcal{W}^{\eta i}-\partial_{i}\pi_{\perp}^{\eta i}\,.

For the numerical algorithm the hydrodynamic equations must be written in conservative flux form [75]:161616For a conformal fluid, the energy density’s spatial derivatives are needed to evaluate the source term i𝒫=12(ii𝒫L)\partial_{i}\mathcal{P}_{\perp}=\frac{1}{2}(\partial_{i}\mathcal{E}-\partial_{i}\mathcal{P}_{L}).

τ𝒒(x)+i𝑭i(x)=𝑺(τ,𝒒(x),𝒖(x),(x),m𝒒(x),μ𝒖(x)).\partial_{\tau}\bm{q}(x)+\partial_{i}\bm{F}^{i}(x)=\bm{S}(\tau,\bm{q}(x),\bm{u}(x),\mathcal{E}(x),\partial_{m}\bm{q}(x),\partial_{\mu}\bm{u}(x))\,. (13)

Here 𝑭i=vi𝒒\bm{F}^{i}=v^{i}\bm{q} are the currents, 𝑺\bm{S} are the source terms and m(x,y,η)m\in(x,y,\eta) is a spatial index. Naturally, the evolution equations for TτμT^{\tau\mu} already assume this form. In the next subsection we will see that the relaxation equations for the dissipative flows require additional manipulations.

2.4 Relaxation equations

The relaxation equations for the dissipative flows 𝒫L\mathcal{P}_{L}, 𝒫\mathcal{P}_{\perp}, WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}} are [66, 65]

𝒫˙L=\displaystyle\dot{\mathcal{P}}_{L}= 𝒫eq𝒫¯τΠ𝒫L𝒫3τπ/2+ζ¯zLθL+ζ¯Lθ2Wzμz˙μ\displaystyle\,\frac{\mathcal{P}_{\text{eq}}{-}\bar{\mathcal{P}}}{\tau_{\Pi}}-\frac{\mathcal{P}_{L}{-}\mathcal{P}_{\perp}}{3\tau_{\pi}/2}+\bar{\zeta}^{L}_{z}\theta_{L}+\bar{\zeta}^{L}_{\perp}\theta_{\perp}-2W_{\perp z}^{\mu}\dot{z}_{\mu} (14a)
+λ¯WuLWzμDzuμ+λ¯WLWzμzν,μuνλ¯πLπμνσ,μν,\displaystyle+\bar{\lambda}^{L}_{Wu}W_{\perp z}^{\mu}D_{z}u_{\mu}+\bar{\lambda}^{L}_{W\perp}W_{\perp z}^{\mu}z_{\nu}\nabla_{\perp,\mu}u^{\nu}-\bar{\lambda}^{L}_{\pi}\pi_{\perp}^{{\mu\nu}}\sigma_{\perp,{\mu\nu}}\,,
𝒫˙=\displaystyle\dot{\mathcal{P}}_{\perp}= 𝒫eq𝒫¯τΠ+𝒫L𝒫3τπ+ζ¯zθL+ζ¯θ+Wzμz˙μ\displaystyle\,\frac{\mathcal{P}_{\text{eq}}{-}\bar{\mathcal{P}}}{\tau_{\Pi}}+\frac{\mathcal{P}_{L}{-}\mathcal{P}_{\perp}}{3\tau_{\pi}}+\bar{\zeta}^{\perp}_{z}\theta_{L}+\bar{\zeta}^{\perp}_{\perp}\theta_{\perp}+W_{\perp z}^{\mu}\dot{z}_{\mu} (14b)
+λ¯WuWzμDzuμλ¯WWzμzν,μuν+λ¯ππμνσ,μν,\displaystyle+\bar{\lambda}^{\perp}_{Wu}W_{\perp z}^{\mu}D_{z}u_{\mu}-\bar{\lambda}^{\perp}_{W\perp}W_{\perp z}^{\mu}z_{\nu}\nabla_{\perp,\mu}u^{\nu}+\bar{\lambda}^{\perp}_{\pi}\pi_{\perp}^{{\mu\nu}}\sigma_{\perp,{\mu\nu}}\,,
W˙z{μ}=\displaystyle\dot{W}^{\{\mu\}}_{\perp z}= Wzμτπ+2η¯uWΞμνDzuν2η¯Wzνμuν(τ¯zWΞμν+πμν)z˙ν\displaystyle\,-\frac{W_{\perp z}^{\mu}}{\tau_{\pi}}+2\bar{\eta}^{W}_{u}\Xi^{\mu\nu}D_{z}u_{\nu}-2\bar{\eta}^{W}_{\perp}z_{\nu}\nabla_{\perp}^{\mu}u^{\nu}-\big{(}\bar{\tau}^{W}_{z}\Xi^{\mu\nu}{+}\pi_{\perp}^{{\mu\nu}}\big{)}\dot{z}_{\nu} (14c)
λ¯WuWWzμθL+δ¯WWWzμθ+λ¯WWσμνWz,ν+ωμνWz,ν\displaystyle-\bar{\lambda}^{W}_{Wu}W_{\perp z}^{\mu}\theta_{L}+\bar{\delta}^{W}_{W}W_{\perp z}^{\mu}\theta_{\perp}+\bar{\lambda}^{W}_{W\perp}\sigma_{\perp}^{\mu\nu}W_{\perp z,\nu}+\omega_{\perp}^{\mu\nu}W_{\perp z,\nu}
+λ¯πuWπμνDzuνλ¯πWπμνzλ,νuλ,\displaystyle+\bar{\lambda}^{W}_{\pi u}\pi_{\perp}^{{\mu\nu}}D_{z}u_{\nu}-\bar{\lambda}^{W}_{\pi\perp}\pi_{\perp}^{{\mu\nu}}z_{\lambda}\nabla_{\perp,\nu}u^{\lambda}\,,
π˙{μν}=\displaystyle\dot{\pi}^{\{{\mu\nu}\}}_{\perp}= πμντπ+2η¯σμν2Wz{μz˙ν}+λ¯πππμνθLδ¯πππμνθ\displaystyle\,-\frac{\pi_{\perp}^{{\mu\nu}}}{\tau_{\pi}}+2\bar{\eta}_{\perp}\sigma_{\perp}^{\mu\nu}-2W_{\perp z}^{\{\mu}\dot{z}^{\nu\}}+\bar{\lambda}^{\pi}_{\pi}\pi_{\perp}^{{\mu\nu}}\theta_{L}-\bar{\delta}^{\pi}_{\pi}\pi_{\perp}^{{\mu\nu}}\theta_{\perp} (14d)
τ¯πππλ{μσ,λν}+2πλ{μω,λν}λ¯WuπWz{μDzuν}.\displaystyle-\bar{\tau}^{\pi}_{\pi}\pi_{\perp}^{\lambda\{\mu}\sigma^{\nu\}}_{\perp,\lambda}+2\pi_{\perp}^{\lambda\{\mu}\omega^{\nu\}}_{\perp,\lambda}-\bar{\lambda}^{\pi}_{Wu}W_{\perp z}^{\{\mu}D_{z}u^{\nu\}}\,.

Here a dot above any quantity denotes the co-moving time derivative uγDγu^{\gamma}D_{\gamma}, and curly brackets denote either the transverse projection of a vector, t{μ}=Ξαμtαt^{\{\mu\}}=\Xi^{\mu}_{\alpha}t^{\alpha}, or the traceless double transverse projection of a rank-2 tensor, t{μν}=Ξαβμνtαβt^{\{\mu\nu\}}=\Xi^{\mu\nu}_{\alpha\beta}t^{\alpha\beta}. We also define the longitudinal and transverse expansion rates θL=zμDzuμ\theta_{L}=z_{\mu}D_{z}u^{\mu} and θ=,μuμ\theta_{\perp}=\nabla_{\perp,\mu}u^{\mu}, where Dz=zνDνD_{z}=-z^{\nu}D_{\nu} is the LRF longitudinal derivative and μ=ΞμνDν\nabla_{\perp}^{\mu}=\Xi^{\mu\nu}D_{\nu} is the transverse gradient. The transverse velocity-shear tensor is σμν=ΞαβμνD(αuβ)\sigma_{\perp}^{\mu\nu}=\Xi^{\mu\nu}_{\alpha\beta}D^{(\alpha}u^{\beta)}, while the transverse vorticity tensor is given by ωμν=ΞαμΞβνD[αuβ]\omega_{\perp}^{\mu\nu}=\Xi^{\mu}_{\alpha}\Xi^{\nu}_{\beta}D^{[\alpha}u^{\beta]}, where the square brackets D[αuβ]=12(DαuβDβuα)D^{[\alpha}u^{\beta]}=\frac{1}{2}(D^{\alpha}u^{\beta}-D^{\beta}u^{\alpha}) denote anti-symmetrization. The shear and bulk viscosity to entropy density ratios η/𝒮\eta/\mathcal{S} and ζ/𝒮\zeta/\mathcal{S} and their associated relaxation times τπ\tau_{\pi} and τΠ\tau_{\Pi}, along with the anisotropic transport coefficients coupled to the gradient forces, will be discussed in Sec. 2.6.

We recast the relaxation equations in conservative flux form by using the product rule identities

W˙z{μ}\displaystyle\dot{W}_{\perp z}^{\{\mu\}} =ΞαμuγDγWzα=uγDγWzμWzαuγDγΞαμ,\displaystyle=\Xi^{\mu}_{\alpha}u^{\gamma}D_{\gamma}W_{\perp z}^{\alpha}=u^{\gamma}D_{\gamma}W_{\perp z}^{\mu}-W_{\perp z}^{\alpha}u^{\gamma}D_{\gamma}\Xi^{\mu}_{\alpha}\,, (15a)
π˙{μν}\displaystyle\dot{\pi}_{\perp}^{\{{\mu\nu}\}} =ΞαβμνuγDγπαβ=uγDγπμνπαβuγDγΞαβμν\displaystyle=\Xi^{\mu\nu}_{\alpha\beta}u^{\gamma}D_{\gamma}\pi_{\perp}^{\alpha\beta}=u^{\gamma}D_{\gamma}\pi_{\perp}^{{\mu\nu}}-\pi_{\perp}^{\alpha\beta}u^{\gamma}D_{\gamma}\Xi^{\mu\nu}_{\alpha\beta} (15b)

to rewrite the l.h.s. of Eqs. (14a-d) as

𝒫˙L=uγγ𝒫L,\displaystyle\dot{\mathcal{P}}_{L}=u^{\gamma}\partial_{\gamma}\mathcal{P}_{L}\,, (16a)
𝒫˙=uγγ𝒫,\displaystyle\dot{\mathcal{P}}_{\perp}=u^{\gamma}\partial_{\gamma}\mathcal{P}_{\perp}\,, (16b)
W˙z{μ}=uγγWzμ+uγΓγλμWzλ+Wzα(uμaαzμz˙α),\displaystyle\dot{W}_{\perp z}^{\{\mu\}}=u^{\gamma}\partial_{\gamma}W_{\perp z}^{\mu}+u^{\gamma}\Gamma^{\mu}_{\gamma\lambda}W_{\perp z}^{\lambda}+W_{\perp z}^{\alpha}(u^{\mu}a_{\alpha}{-}z^{\mu}\dot{z}_{\alpha})\,, (16c)
π˙{μν}=uγγπμν+uγΓγλμπνλ+uγΓγλνπμλ\displaystyle\dot{\pi}_{\perp}^{\{{\mu\nu}\}}=u^{\gamma}\partial_{\gamma}\pi_{\perp}^{{\mu\nu}}+u^{\gamma}\Gamma^{\mu}_{\gamma\lambda}\pi_{\perp}^{\nu\lambda}+u^{\gamma}\Gamma^{\nu}_{\gamma\lambda}\pi_{\perp}^{\mu\lambda} (16d)
+πμα(uνaαzνz˙α)+πνα(uμaαzμz˙α),\displaystyle\qquad\quad+\pi_{\perp}^{\mu\alpha}(u^{\nu}a_{\alpha}{-}z^{\nu}\dot{z}_{\alpha})+\pi_{\perp}^{\nu\alpha}(u^{\mu}a_{\alpha}{-}z^{\mu}\dot{z}_{\alpha})\,,

where aμ=u˙μa^{\mu}=\dot{u}^{\mu} is the fluid acceleration. Finally, we use the product rule identity

uγγ𝒫L,=uτ[τ𝒫L,+i(vi𝒫L,)𝒫L,ivi]u^{\gamma}\partial_{\gamma}\mathcal{P}_{L,\perp}=u^{\tau}\left[\partial_{\tau}\mathcal{P}_{L,\perp}+\partial_{i}(v^{i}\mathcal{P}_{L,\perp})-\mathcal{P}_{L,\perp}\partial_{i}v^{i}\right] (17)

to rewrite Eqs. (16a-b) as

τ𝒫L+i(vi𝒫L)\displaystyle\partial_{\tau}\mathcal{P}_{L}+\partial_{i}(v^{i}\mathcal{P}_{L}) =𝒫Livi+1uτ[𝒫eq𝒫¯τΠ𝒫L𝒫3τπ/2+L],\displaystyle=\mathcal{P}_{L}\partial_{i}v^{i}+\frac{1}{u^{\tau}}\left[\frac{\mathcal{P}_{\text{eq}}{-}\bar{\mathcal{P}}}{\tau_{\Pi}}-\frac{\mathcal{P}_{L}{-}\mathcal{P}_{\perp}}{3\tau_{\pi}/2}+\mathcal{I}_{L}\right]\,, (18a)
τ𝒫+i(vi𝒫)\displaystyle\partial_{\tau}\mathcal{P}_{\perp}+\partial_{i}(v^{i}\mathcal{P}_{\perp}) =𝒫ivi+1uτ[𝒫eq𝒫¯τΠ+𝒫L𝒫3τπ+];\displaystyle=\mathcal{P}_{\perp}\partial_{i}v^{i}+\frac{1}{u^{\tau}}\left[\frac{\mathcal{P}_{\text{eq}}{-}\bar{\mathcal{P}}}{\tau_{\Pi}}+\frac{\mathcal{P}_{L}{-}\mathcal{P}_{\perp}}{3\tau_{\pi}}+\mathcal{I}_{\perp}\right]\,; (18b)

here

L=\displaystyle\mathcal{I}_{L}= ζ¯zLθL+ζ¯Lθ2Wzμz˙μ+λ¯WuLWzμDzuμ+λ¯WLWzμzν,μuν\displaystyle\,\bar{\zeta}^{L}_{z}\theta_{L}+\bar{\zeta}^{L}_{\perp}\theta_{\perp}-2W_{\perp z}^{\mu}\dot{z}_{\mu}+\bar{\lambda}^{L}_{Wu}W_{\perp z}^{\mu}D_{z}u_{\mu}+\bar{\lambda}^{L}_{W\perp}W_{\perp z}^{\mu}z_{\nu}\nabla_{\perp,\mu}u^{\nu} (19a)
λ¯πLπμνσ,μν,\displaystyle-\bar{\lambda}^{L}_{\pi}\pi_{\perp}^{{\mu\nu}}\sigma_{\perp,{\mu\nu}}\,,
=\displaystyle\mathcal{I}_{\perp}= ζ¯zθL+ζ¯θ+Wzμz˙μ+λ¯WuWzμDzuμλ¯WWzμzν,μuν\displaystyle\,\bar{\zeta}^{\perp}_{z}\theta_{L}+\bar{\zeta}^{\perp}_{\perp}\theta_{\perp}+W_{\perp z}^{\mu}\dot{z}_{\mu}+\bar{\lambda}^{\perp}_{Wu}W_{\perp z}^{\mu}D_{z}u_{\mu}-\bar{\lambda}^{\perp}_{W\perp}W_{\perp z}^{\mu}z_{\nu}\nabla_{\perp,\mu}u^{\nu} (19b)
+λ¯ππμνσ,μν\displaystyle+\bar{\lambda}^{\perp}_{\pi}\pi_{\perp}^{{\mu\nu}}\sigma_{\perp,{\mu\nu}}

are the gradient source terms for 𝒫L\mathcal{P}_{L} and 𝒫\mathcal{P}_{\perp}. Similarly Eqs. (16c-d) can be rewritten as

τWzμ+i(viWzμ)\displaystyle\partial_{\tau}W_{\perp z}^{\mu}+\partial_{i}(v^{i}W_{\perp z}^{\mu}) =Wzμivi+1uτ[Wzμτπ+Wμ𝒫Wμ𝒢Wμ],\displaystyle=W_{\perp z}^{\mu}\partial_{i}v^{i}+\frac{1}{u^{\tau}}\left[-\frac{W_{\perp z}^{\mu}}{\tau_{\pi}}+\mathcal{I}^{\mu}_{W}-\mathcal{P}^{\mu}_{W}-\mathcal{G}_{W}^{\mu}\right]\,, (20a)
τπμν+i(viπμν)\displaystyle\partial_{\tau}\pi_{\perp}^{{\mu\nu}}+\partial_{i}(v^{i}\pi_{\perp}^{{\mu\nu}}) =πμνivi+1uτ[πμντπ+πμν𝒫πμν𝒢πμν];\displaystyle=\pi_{\perp}^{{\mu\nu}}\partial_{i}v^{i}+\frac{1}{u^{\tau}}\left[-\frac{\pi_{\perp}^{{\mu\nu}}}{\tau_{\pi}}+\mathcal{I}^{\mu\nu}_{\pi}-\mathcal{P}^{\mu\nu}_{\pi}-\mathcal{G}_{\pi}^{\mu\nu}\right]\,; (20b)

here

Wμ=\displaystyle\mathcal{I}^{\mu}_{W}= Ξμν(2η¯uWDzuντ¯zWz˙ν)2η¯Wzνμuνπμνz˙νλ¯WuWWzμθL\displaystyle\,\Xi^{\mu\nu}\big{(}2\bar{\eta}^{W}_{u}D_{z}u_{\nu}{-}\bar{\tau}^{W}_{z}\dot{z}_{\nu}\big{)}-2\bar{\eta}^{W}_{\perp}z_{\nu}\nabla_{\perp}^{\mu}u^{\nu}-\pi_{\perp}^{{\mu\nu}}\dot{z}_{\nu}-\bar{\lambda}^{W}_{Wu}W_{\perp z}^{\mu}\theta_{L} (21a)
+δ¯WWWzμθ+λ¯WWσμνWz,ν+ωμνWz,ν+λ¯πuWπμνDzuν\displaystyle+\bar{\delta}^{W}_{W}W_{\perp z}^{\mu}\theta_{\perp}+\bar{\lambda}^{W}_{W\perp}\sigma_{\perp}^{\mu\nu}W_{\perp z,\nu}+\omega_{\perp}^{\mu\nu}W_{\perp z,\nu}+\bar{\lambda}^{W}_{\pi u}\pi_{\perp}^{{\mu\nu}}D_{z}u_{\nu}
λ¯πWπμνzλ,νuλ,\displaystyle-\bar{\lambda}^{W}_{\pi\perp}\pi_{\perp}^{{\mu\nu}}z_{\lambda}\nabla_{\perp,\nu}u^{\lambda}\,,
πμν=\displaystyle\mathcal{I}^{\mu\nu}_{\pi}= Ξαβμν(2πλ(αω,λβ)τ¯πππλ(ασ,λβ)2Wz(αz˙β)λ¯WuπWz(αDzuβ))\displaystyle\,\Xi^{\mu\nu}_{\alpha\beta}\big{(}2\pi_{\perp}^{\lambda(\alpha}\omega^{\beta)}_{\perp,\lambda}-\bar{\tau}^{\pi}_{\pi}\pi_{\perp}^{\lambda(\alpha}\sigma^{\beta)}_{\perp,\lambda}-2W_{\perp z}^{(\alpha}\dot{z}^{\beta)}-\bar{\lambda}^{\pi}_{Wu}W_{\perp z}^{(\alpha}D_{z}u^{\beta)}\big{)} (21b)
+2η¯σμν+λ¯πππμνθLδ¯πππμνθ\displaystyle+2\bar{\eta}_{\perp}\sigma_{\perp}^{\mu\nu}+\bar{\lambda}^{\pi}_{\pi}\pi_{\perp}^{{\mu\nu}}\theta_{L}-\bar{\delta}^{\pi}_{\pi}\pi_{\perp}^{{\mu\nu}}\theta_{\perp}

are the gradient source terms,

𝒫Wμ=\displaystyle\mathcal{P}^{\mu}_{W}= Wzα(uμaαzμz˙α),\displaystyle\,W_{\perp z}^{\alpha}(u^{\mu}a_{\alpha}{-}z^{\mu}\dot{z}_{\alpha})\,, (22a)
𝒫πμν=\displaystyle\mathcal{P}^{\mu\nu}_{\pi}= πμα(uνaαzνz˙α)+πνα(uμaαzμz˙α)\displaystyle\,\pi_{\perp}^{\mu\alpha}(u^{\nu}a_{\alpha}{-}z^{\nu}\dot{z}_{\alpha})+\pi_{\perp}^{\nu\alpha}(u^{\mu}a_{\alpha}{-}z^{\mu}\dot{z}_{\alpha}) (22b)

are the transverse projection source terms, and

𝒢Wμ=\displaystyle\mathcal{G}^{\mu}_{W}= uγΓγλμWzλ,\displaystyle\,u^{\gamma}\Gamma^{\mu}_{\gamma\lambda}W_{\perp z}^{\lambda}\,, (23a)
𝒢πμν=\displaystyle\mathcal{G}^{\mu\nu}_{\pi}= uγΓγλμπνλ+uγΓγλνπμλ\displaystyle\,u^{\gamma}\Gamma^{\mu}_{\gamma\lambda}\pi_{\perp}^{\nu\lambda}+u^{\gamma}\Gamma^{\nu}_{\gamma\lambda}\pi_{\perp}^{\mu\lambda} (23b)

are the geometric source terms for WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}}. The individual components that make up the source terms in the relaxation equations (18a-b) and (20a-b) are listed in Appendices A and B.

If we evolve anisotropic fluid dynamics with a QCD equation of state (see Secs. 2.5 and 2.6) we also propagate a mean field BB as a dynamical variable. Its relaxation equation is [79, 66]

B˙=BeqBτΠm˙m(2𝒫𝒫L4B)\dot{B}=\frac{B_{\text{eq}}{-}B}{\tau_{\Pi}}-\frac{\dot{m}}{m}(\mathcal{E}{-}2\mathcal{P}_{\perp}{-}\mathcal{P}_{L}{-}4B) (24)

or, in conservative flux form,

τB+i(viB)=Bivi+1uτ[BeqBτΠm˙m(2𝒫𝒫L4B)],\partial_{\tau}B+\partial_{i}(v^{i}B)=B\,\partial_{i}v^{i}+\frac{1}{u^{\tau}}\left[\frac{B_{\text{eq}}{-}B}{\tau_{\Pi}}-\frac{\dot{m}}{m}(\mathcal{E}{-}2\mathcal{P}_{\perp}{-}\mathcal{P}_{L}{-}4B)\right]\,, (25)

where Beq()B_{\text{eq}}(\mathcal{E}) is the equilibrium mean field and m()m(\mathcal{E}) is the quasiparticle mass.

2.5 Equation of state

Refer to caption
Figure 1: (Color online) Left: The QCD energy density (blue) and equilibrium pressure (red), normalized by their Stefan–Boltzmann limits (26). Right: The squared speed of sound (purple) as a function of temperature; the gray line indicates the conformal limit cs2=13c_{s}^{2}=\frac{1}{3}.

In the code there are two options for the quark-gluon plasma’s equation of state 𝒫eq()\mathcal{P}_{\text{eq}}(\mathcal{E}): conformal and QCD. The former assumes a non-interacting gas of massless quarks and gluons:

𝒫eq=3=gT4π2,\mathcal{P}_{\text{eq}}=\frac{\mathcal{E}}{3}=\frac{gT^{4}}{\pi^{2}}\,, (26)

where TT is the temperature and the degeneracy factor is

g=π490[2(Nc21)+72NcNf],g=\frac{\pi^{4}}{90}\left[2\left(N_{c}^{2}{-}1\right)+\frac{7}{2}N_{c}N_{f}\right]\,, (27)

with Nc=3N_{c}=3 colors and Nf=3N_{f}=3 massless quark flavors (i.e. up, down and strange). This equation of state is primarily used to test the fluid dynamical simulation subject to either conformal Bjorken expansion or conformal Gubser expansion (see Secs. 4.1 and 4.2).

The QCD equation of state that we employ for more realistic simulations interpolates between the lattice QCD calculations provided by the HotQCD collaboration [80] and a hadron resonance gas composed of the hadrons that can be propagated in the hadronic afterburner code SMASH [81].171717In hybrid model simulations of heavy-ion collisions, the equilibrium equation of state used in the hydrodynamic module must be consistent with that in the hadronic afterburner. Otherwise, serious violations of energy and momentum conservation can occur at the hadronization phase. Fig. 1 shows the energy density, equilibrium pressure and speed of sound as a function of temperature. The data table in the code covers the temperature range T[0.05,1.0]T\in[0.05,1.0] GeV.

2.6 Transport coefficients

Refer to caption
Figure 2: (Color online) The quasiparticle mass to temperature ratio (blue, left) and equilibrium mean field normalized to the QCD equilibrium pressure (orange, right) as functions of temperature (for conformal systems, m=0m=0 and Beq=0B_{\text{eq}}=0).

In this section we list the transport coefficients appearing in the relaxation equations in Sec. 2.4, for both QCD and conformal equations of state.

At the present moment, the transport coefficients of QCD matter (especially in the nonperturbative region T[0.15,0.5]T\in[0.15,0.5] GeV) are not explicitly known from first principles. Instead, we parametrize the shear and bulk viscosities (η/𝒮)(T)(\eta/\mathcal{S})(T) and (ζ/𝒮)(T)(\zeta/\mathcal{S})(T) as a function of temperature. In this work, we use the best-fit parametrization models from the JETSCAPE collaboration [1, 2]. The relaxation times and anisotropic transport coefficients are computed with a quasiparticle kinetic theory model, whose equation of state is fitted to the QCD one. The kinetic model contains quasiparticles with a temperature-dependent mass m(T)m(T) and an equilibrium mean-field Beq(T)B_{\text{eq}}(T); these are shown in Figure 2.

2.6.1 Shear and bulk viscosities

Refer to caption
Figure 3: (Color online) The temperature parametrization of (η/𝒮)(T)(\eta/\mathcal{S})(T) and (ζ/𝒮)(T)(\zeta/\mathcal{S})(T) used in this work. (For conformal systems, we set η/𝒮=0.2\eta/\mathcal{S}=0.2 and ζ/𝒮=0\zeta/\mathcal{S}=0.)

The shear viscosity is modeled as a linear piecewise function with a kink at temperature TηT_{\eta}:

(η/𝒮)(T)=(η/𝒮)kink+(TTη)(alowΘ(TηT)+ahighΘ(TTη)),(\eta/\mathcal{S})(T)=(\eta/\mathcal{S})_{\text{kink}}+(T{-}T_{\eta})\left(a_{\text{low}}\Theta(T_{\eta}{-}T)+a_{\text{high}}\Theta(T{-}T_{\eta})\right)\,, (28)

where (η/𝒮)kink(\eta/\mathcal{S})_{\text{kink}} is the value of η/𝒮\eta/\mathcal{S} at TηT_{\eta}, alowa_{\text{low}} and ahigha_{\text{high}} are the left and right slopes, respectively, and Θ\Theta is the Heaviside step function. The bulk viscosity is parametrized as a skewed Cauchy distribution:

(ζ/𝒮)(T)=(ζ/𝒮)maxΛζ(T)2Λζ(T)2+(TTζ)2,(\zeta/\mathcal{S})(T)=\frac{(\zeta/\mathcal{S})_{\text{max}}\,\Lambda_{\zeta}(T)^{2}}{\Lambda_{\zeta}(T)^{2}+(T{-}T_{\zeta})^{2}}\,, (29)

where (ζ/𝒮)max(\zeta/\mathcal{S})_{\text{max}} is the normalization factor, TζT_{\zeta} is the peak temperature and

Λζ(T)=wζ(1+λζsgn(TTζ)),\Lambda_{\zeta}(T)=w_{\zeta}\left(1+\lambda_{\zeta}\,\mathrm{sgn}(T{-}T_{\zeta})\right)\,, (30)

with wζw_{\zeta} and λζ\lambda_{\zeta} being the width and skewness parameters, respectively.

The best-fit values for the viscosity parameters181818They correspond to a hybrid model whose particlization phase uses the 14-moment approximation for the δf\delta f correction in the Cooper-Frye formula [74, 1, 2] (the viscosity parameters used for the code validation tests in Sec. 4 are slightly different because they were taken from an early draft of Ref. [2]). are (η/𝒮)kink=0.096(\eta/\mathcal{S})_{\text{kink}}=0.096, Tη=0.223T_{\eta}=0.223 GeV, alow=0.776a_{\text{low}}=-0.776 GeV-1, ahigh=0.37a_{\text{high}}=0.37 GeV-1, (ζ/𝒮)max=0.133(\zeta/\mathcal{S})_{\text{max}}=0.133, Tζ=0.12T_{\zeta}=0.12 GeV, wζ=0.072w_{\zeta}=0.072 GeV and λζ=0.122\lambda_{\zeta}=-0.122 (see Table II in Ref. [2]). The resulting temperature dependence of η/𝒮\eta/\mathcal{S} and ζ/𝒮\zeta/\mathcal{S} is shown in Figure 3.

For conformal systems, we fix the shear viscosity to η/𝒮=0.2\eta/\mathcal{S}=0.2 and the bulk viscosity to ζ/𝒮=0\zeta/\mathcal{S}=0.

2.6.2 Shear and bulk relaxation times

Refer to caption
Figure 4: (Color online) The dimensionless shear and bulk relaxation times computed with the quasiparticle kinetic model (solid color) and small-mass approximation (dashed color). (For conformal systems, τπT=5η/𝒮\tau_{\pi}T=5\eta/\mathcal{S} and τΠT=0\tau_{\Pi}T=0.)

In the quasiparticle kinetic model [79, 66], the shear and bulk relaxation times are proportional to the shear and bulk viscosity, respectively,

τπ=ηβπ,τΠ=ζβΠ,\tau_{\pi}=\frac{\eta}{\beta_{\pi}}\,,\qquad\tau_{\Pi}=\frac{\zeta}{\beta_{\Pi}}\,, (31)

where the viscosity to relaxation time ratios βπ\beta_{\pi} and βΠ\beta_{\Pi} are

βπ(T)\displaystyle\beta_{\pi}(T) =32T,\displaystyle=\frac{\mathcal{I}_{32}}{T}\,, (32a)
βΠ(T)\displaystyle\beta_{\Pi}(T) =53βπ+cs2(mdmdT11(+𝒫eq)),\displaystyle=\frac{5}{3}\beta_{\pi}+c_{s}^{2}\big{(}m\frac{dm}{dT}\mathcal{I}_{11}-(\mathcal{E}{+}\mathcal{P}_{\text{eq}})\big{)}\,, (32b)

and cs2c_{s}^{2} and 𝒫eq\mathcal{P}_{\text{eq}} are evaluated with the QCD equation of state. Here, we have defined the thermodynamic integrals

nq=gP(up)n2q(pΔp)qfeq(2q+1)!!,\mathcal{I}_{nq}=g\int_{P}\frac{(u\cdot p)^{n-2q}(-p\cdot\Delta\cdot p)^{q}f_{\text{eq}}}{(2q{+}1)!!}\,, (33)

where Pd3p/Ep\int_{P}\dots\equiv\int d^{3}p/E_{p}\dots indicates integration with the Lorentz-invariant momentum space measure, pμp^{\mu} is the quasiparticle momentum, pΔp-p\cdot\Delta\cdot p is the square of its spatial LRF momentum, up=m2(T)pΔpu\cdot p=\sqrt{m^{2}(T)-p\cdot\Delta\cdot p} is its LRF energy, and feq=exp[up/T]f_{\text{eq}}=\exp\left[-u\cdot p/T\right] is the local-equilibrium distribution function.

The normalized quasiparticle relaxation times τπT\tau_{\pi}T and τΠT\tau_{\Pi}T are shown in Figure 4. These are compared to the relaxation times in standard viscous hydrodynamic models, where the kinetic transport coefficients are computed in the small-mass approximation m¯=m/T1\bar{m}=m/T\ll 1 and dm/dT=0dm/dT=0:

τπT\displaystyle\tau_{\pi}T 5η/𝒮+O(m¯2),\displaystyle\approx 5\eta/\mathcal{S}+O\big{(}\bar{m}^{2}\big{)}\,, (34a)
τΠT\displaystyle\tau_{\Pi}T ζT15(+𝒫eq)(13cs2)2+O(m¯5).\displaystyle\approx\frac{\zeta\,T}{15(\mathcal{E}{+}\mathcal{P}_{\text{eq}})\big{(}\frac{1}{3}{-}c_{s}^{2}\big{)}^{2}}+O\big{(}\bar{m}^{5}\big{)}\,. (34b)

One sees that the shear relaxation times are very similar to each other, except at low temperatures T<0.2T<0.2 GeV. On the other hand, the bulk relaxation times differ by about an order of magnitude for T<0.2T<0.2 GeV; this is due to the breakdown of the small-mass approximation, even at high temperatures T1T\sim 1 GeV. As a result, the evolution of the bulk viscous pressure Π\Pi will be more greatly affected by critical slowing down in our anisotropic hydrodynamics model compared to standard viscous hydrodynamics.

For conformal kinetic plasmas (m=0m=0, dm/dT=0dm/dT=0), the shear relaxation time is τπ=5η/(𝒮T)\tau_{\pi}=5\eta/(\mathcal{S}T) and the bulk relaxation time is τΠ=0\tau_{\Pi}=0.

2.6.3 Anisotropic transport coefficients

Finally, we list the anisotropic transport coefficients [66, 65] that are coupled to the gradient source terms in the relaxation equations for the longitudinal pressure 𝒫L\mathcal{P}_{L},

ζ¯zL\displaystyle\bar{\zeta}^{L}_{z} =24003(𝒫L+B)+mdmd(+𝒫L)0200,\displaystyle=\mathcal{I}_{2400}-3(\mathcal{P}_{L}{+}B)+m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{L})\mathcal{I}_{0200}\,, (35a)
ζ¯L\displaystyle\bar{\zeta}^{L}_{\perp} =2210𝒫LB+mdmd(+𝒫)0200,\displaystyle=\mathcal{I}_{2210}-\mathcal{P}_{L}-B+m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{\perp})\mathcal{I}_{0200}\,, (35b)
λ¯WuL\displaystyle\bar{\lambda}^{L}_{Wu} =44104210+mdmd0200,\displaystyle=\frac{\mathcal{I}_{4410}}{\mathcal{I}_{4210}}+m\frac{dm}{d\mathcal{E}}\mathcal{I}_{0200}\,, (35c)
λ¯WL\displaystyle\bar{\lambda}^{L}_{W\perp} =1λ¯WuL,\displaystyle=1-\bar{\lambda}^{L}_{Wu}\,, (35d)
λ¯πL\displaystyle\bar{\lambda}^{L}_{\pi} =42204020+mdmd0200,\displaystyle=\frac{\mathcal{I}_{4220}}{\mathcal{I}_{4020}}+m\frac{dm}{d\mathcal{E}}\mathcal{I}_{0200}\,, (35e)

the transverse pressure 𝒫\mathcal{P}_{\perp},

ζ¯z\displaystyle\bar{\zeta}^{\perp}_{z} =2210𝒫B+mdmd(+𝒫L)0010,\displaystyle=\mathcal{I}_{2210}-\mathcal{P}_{\perp}-B+m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{L})\mathcal{I}_{0010}\,, (36a)
ζ¯\displaystyle\bar{\zeta}^{\perp}_{\perp} =2(2020𝒫B)+mdmd(+𝒫)0010,\displaystyle=2(\mathcal{I}_{2020}{-}\mathcal{P}_{\perp}{-}B)+m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{\perp})\mathcal{I}_{0010}\,, (36b)
λ¯W\displaystyle\bar{\lambda}^{\perp}_{W\perp} =242204210+mdmd0010,\displaystyle=\frac{2\,\mathcal{I}_{4220}}{\mathcal{I}_{4210}}+m\frac{dm}{d\mathcal{E}}\mathcal{I}_{0010}\,, (36c)
λ¯Wu\displaystyle\bar{\lambda}^{\perp}_{Wu} =λ¯W1,\displaystyle=\bar{\lambda}^{\perp}_{W\perp}-1\,, (36d)
λ¯π\displaystyle\bar{\lambda}^{\perp}_{\pi} =1340304020mdmd0010,\displaystyle=1-\frac{3\,\mathcal{I}_{4030}}{\mathcal{I}_{4020}}-m\frac{dm}{d\mathcal{E}}\mathcal{I}_{0010}\,, (36e)

the longitudinal momentum diffusion current WzμW_{\perp z}^{\mu},

η¯uW\displaystyle\bar{\eta}^{W}_{u} =12(𝒫L+B2210),\displaystyle=\frac{1}{2}\big{(}\mathcal{P}_{L}+B-\mathcal{I}_{2210})\,, (37a)
η¯W\displaystyle\bar{\eta}^{W}_{\perp} =12(𝒫+B2210),\displaystyle=\frac{1}{2}(\mathcal{P}_{\perp}+B-\mathcal{I}_{2210})\,, (37b)
τ¯zW\displaystyle\bar{\tau}^{W}_{z} =𝒫L𝒫,\displaystyle=\mathcal{P}_{L}-\mathcal{P}_{\perp}\,, (37c)
δ¯WW\displaystyle\bar{\delta}^{W}_{W} =λ¯WW12+mdmd(+𝒫)(22104210),\displaystyle=\bar{\lambda}^{W}_{W\perp}-\frac{1}{2}+m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{\perp})\left(\frac{\mathcal{I}_{2210}}{\mathcal{I}_{4210}}\right)\,, (37d)
λ¯WuW\displaystyle\bar{\lambda}^{W}_{Wu} =244104210mdmd(+𝒫L)(22104210),\displaystyle=2-\frac{\mathcal{I}_{4410}}{\mathcal{I}_{4210}}-m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{L})\left(\frac{\mathcal{I}_{2210}}{\mathcal{I}_{4210}}\right)\,, (37e)
λ¯WW\displaystyle\bar{\lambda}^{W}_{W\perp} =2422042101,\displaystyle=\frac{2\,\mathcal{I}_{4220}}{\mathcal{I}_{4210}}-1, (37f)
λ¯πuW\displaystyle\bar{\lambda}^{W}_{\pi u} =42204020,\displaystyle=\frac{\mathcal{I}_{4220}}{\mathcal{I}_{4020}}\,, (37g)
λ¯πW\displaystyle\bar{\lambda}^{W}_{\pi\perp} =λ¯πuW1,\displaystyle=\bar{\lambda}^{W}_{\pi u}-1\,, (37h)

and the transverse shear stress tensor πμν\pi_{\perp}^{{\mu\nu}}:

η¯\displaystyle\bar{\eta}_{\perp} =𝒫(k)2020,\displaystyle=\mathcal{P}_{\perp}^{(k)}-\mathcal{I}_{2020}, (38a)
δ¯ππ\displaystyle\bar{\delta}^{\pi}_{\pi} =34τ¯ππ+12mdmd(+𝒫)(20204020),\displaystyle=\frac{3}{4}\bar{\tau}^{\pi}_{\pi}+\frac{1}{2}-m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{\perp})\left(\frac{\mathcal{I}_{2020}}{\mathcal{I}_{4020}}\right), (38b)
τ¯ππ\displaystyle\bar{\tau}^{\pi}_{\pi} =2440304020,\displaystyle=2-\frac{4\,\mathcal{I}_{4030}}{\mathcal{I}_{4020}}, (38c)
λ¯ππ\displaystyle\bar{\lambda}^{\pi}_{\pi} =λ¯πuW1+mdmd(+𝒫L)(20204020),\displaystyle=\bar{\lambda}^{W}_{\pi u}-1+m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{L})\left(\frac{\mathcal{I}_{2020}}{\mathcal{I}_{4020}}\right), (38d)
λ¯Wuπ\displaystyle\bar{\lambda}^{\pi}_{Wu} =λ¯WW1,\displaystyle=\bar{\lambda}^{W}_{W\perp}-1, (38e)
λ¯Wπ\displaystyle\bar{\lambda}^{\pi}_{W\perp} =λ¯Wuπ+2.\displaystyle=\bar{\lambda}^{\pi}_{Wu}+2. (38f)

We define the anisotropic integrals

nrqs=g(2π)3P(up)nr2q(zp)r(pΞp)qEasfa(2q)!!,\mathcal{I}_{nrqs}=\frac{g}{(2\pi)^{3}}\int_{P}\frac{(u\cdot p)^{n-r-2q}(-z\cdot p)^{r}(-p\cdot\Xi\cdot p)^{q}E_{a}^{s}f_{a}}{(2q)!!}\,, (39)

where

Ea=m2(T)(pΞp)/α2+(zp)2/αL2E_{a}=\sqrt{m^{2}(T)-(p\cdot\Xi\cdot p)/\alpha_{\perp}^{2}+(-z\cdot p)^{2}/\alpha_{L}^{2}} (40)

and fa=eEa/Λf_{a}=e^{-E_{a}/\Lambda} is the leading-order anisotropic distribution function.

In addition to the effective temperature Λ\Lambda this anisotropic distribution faf_{a} depends (through EaE_{a}) on two momentum anisotropy parameters αL\alpha_{L} and α\alpha_{\perp}, which deform the longitudinal and transverse momentum space, respectively. In order to compute the non-conformal anisotropic transport coefficients at each time step of the simulation, we propagate (Λ,α,αL)(\Lambda,\alpha_{\perp},\alpha_{L}) as inferred variables. They are obtained from the kinetic part of \mathcal{E}, 𝒫L\mathcal{P}_{L} and 𝒫\mathcal{P}_{\perp} in the quasiparticle kinetic model:

(k)\displaystyle\mathcal{E}^{(k)} =B,\displaystyle=\mathcal{E}-B\,, (41a)
𝒫L(k)\displaystyle\mathcal{P}_{L}^{(k)} =𝒫L+B,\displaystyle=\mathcal{P}_{L}+B\,, (41b)
𝒫(k)\displaystyle\mathcal{P}_{\perp}^{(k)} =𝒫+B,\displaystyle=\mathcal{P}_{\perp}+B\,, (41c)

where (k)=2000\mathcal{E}^{(k)}=\mathcal{I}_{2000} is the kinetic contribution to the energy density (i.e. the total energy density minus the mean field contribution [66]), 𝒫L(k)=2200\mathcal{P}_{L}^{(k)}=\mathcal{I}_{2200} is the kinetic longitudinal pressure, and 𝒫(k)=2010\mathcal{P}_{\perp}^{(k)}=\mathcal{I}_{2010} is the kinetic transverse pressure. The numerical method which we use to solve these equations will be discussed in Sec. 3.4.

The anisotropic transport coefficients in the conformal limit are listed in Appendix C.

3 Numerical scheme

In this section we discuss the numerical implementation of the hydrodynamic equations in the code. The dynamical and inferred variables are evolved on an (Nx+4)×(Ny+4)×(Nη+4)(N_{x}{+}4)\times(N_{y}{+}4)\times(N_{\eta}{+}4) Eulerian grid, where NxN_{x}, NyN_{y} and NηN_{\eta} are the number of physical grid points along each spatial direction.191919For longitudinally boost-invariant systems, the number of spacetime rapidity points is set to Nη=1N_{\eta}=1. A grid point with cell index (i,j,k)(i,j,k) that corresponds to the lower left front corner of a fluid cell, has a spatial position

xi\displaystyle x_{i} =[i212(Nx1)]Δx,\displaystyle=\big{[}i-2-\frac{1}{2}(N_{x}{-}1)\big{]}\Delta x\,, (42a)
yj\displaystyle y_{j} =[j212(Ny1)]Δy,\displaystyle=\big{[}j-2-\frac{1}{2}(N_{y}{-}1)\big{]}\Delta y\,, (42b)
ηs,k\displaystyle\eta_{s,k} =[k212(Nη1)]Δηs,\displaystyle=\big{[}k-2-\frac{1}{2}(N_{\eta}{-}1)\big{]}\Delta\eta_{s}\,, (42c)

where Δx\Delta x, Δy\Delta y and Δηs\Delta\eta_{s} are the lattice spacings. Physical fluid cells have indices i[2,Nx+1]i\in[2,N_{x}{+}1], j[2,Ny+1]j\in[2,N_{y}{+}1] and k[2,Nη+1]k\in[2,N_{\eta}{+}1]. The numerical algorithm also requires six sets of ghost cells with depth two, which neighbor the physical grid’s faces (for an illustration, see Fig. 3 in Ref. [20]). The boundary conditions that we impose on the ghost cells neighboring the two (y,ηs)(y,\eta_{s}) faces are202020In conformal anisotropic hydrodynamics, the energy density \mathcal{E} also requires ghost cell boundary conditions.

𝒒0,j,k\displaystyle\bm{q}_{0,j,k} =𝒒1,j,k=𝒒2,j,k,\displaystyle=\bm{q}_{1,j,k}=\bm{q}_{2,j,k}\,, (43a)
𝒖0,j,k\displaystyle\bm{u}_{0,j,k} =𝒖1,j,k=𝒖2,j,k,\displaystyle=\bm{u}_{1,j,k}=\bm{u}_{2,j,k}\,, (43b)
𝒒Nx+2,j,k\displaystyle\bm{q}_{N_{x}{+}2,j,k} =𝒒Nx+3,j,k=𝒒Nx+1,j,k,\displaystyle=\bm{q}_{N_{x}{+}3,j,k}=\bm{q}_{N_{x}{+}1,j,k}\,, (43c)
𝒖Nx+2,j,k\displaystyle\bm{u}_{N_{x}{+}2,j,k} =𝒖Nx+3,j,k=𝒖Nx+1,j,k,\displaystyle=\bm{u}_{N_{x}{+}3,j,k}=\bm{u}_{N_{x}{+}1,j,k}\,, (43d)

and similarly for (x,ηs)(x,\eta_{s}) and (x,y)(x,y) faces after permuting the grid indices and replacing NxNyN_{x}\to N_{y} or NηN_{\eta}.

The dynamical variables 𝒒\bm{q} are updated using a two-stage Runge–Kutta (RK2) scheme, where the time derivatives are evaluated with the Kurganov–Tadmor (KT) algorithm [16, 20, 75]. After each intermediate Euler step in the RK2 scheme, we reconstruct the inferred variables from the dynamical variables. In addition, we regulate the residual shear stresses WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}} and the mean-field BB. The code also provides the user with the option of using an adaptive time step to capture the fluid’s longitudinal dynamics at very early times; this will be discussed at the end of the section.

3.1 Kurganov-Tadmor algorithm

The time derivative of the dynamical variables τ𝒒\partial_{\tau}\bm{q} in the partial differential equations (13) can be computed on the Eulerian grid at any time τ\tau using the KT algorithm [75]:

(τ𝒒)ijk=𝑯i+12,j,kx𝑯i12,j,kxΔx𝑯i,j+12,ky𝑯i,j12,kyΔy𝑯i,j,k+12η𝑯i,j,k12ηΔηs+𝑺ijk(τ,𝒒ijk,𝒖ijk,ijk,(m𝒒)ijk,(μ𝒖)ijk),\begin{split}(\partial_{\tau}\bm{q})_{ijk}=&-\frac{\bm{H}^{x}_{i{+}\frac{1}{2},j,k}{-}\bm{H}^{x}_{i{-}\frac{1}{2},j,k}}{\Delta x}-\frac{\bm{H}^{y}_{i,j{+}\frac{1}{2},k}{-}\bm{H}^{y}_{i,j{-}\frac{1}{2},k}}{\Delta y}-\frac{\bm{H}^{\eta}_{i,j,k{+}\frac{1}{2}}{-}\bm{H}^{\eta}_{i,j,k{-}\frac{1}{2}}}{\Delta\eta_{s}}\\ &+\bm{S}_{ijk}\big{(}\tau,\bm{q}_{ijk},\bm{u}_{ijk},\mathcal{E}_{ijk},(\partial_{m}\bm{q})_{ijk},(\partial_{\mu}\bm{u})_{ijk}\big{)}\,,\end{split} (44)

where the numerical fluxes evaluated at the left and right faces of a staggered cell centered around the grid point (i,j,k)(i,j,k) are

𝑯i±12,j,kx=12[𝑭i±12,j,kx++𝑭i±12,j,kxsi±12,j,kx(𝒒i±12,j,k+𝒒i±12,j,k)],\bm{H}^{x}_{i\pm\frac{1}{2},j,k}=\frac{1}{2}\Big{[}\bm{F}^{x+}_{i\pm\frac{1}{2},j,k}+\bm{F}^{x-}_{i\pm\frac{1}{2},j,k}-s^{x}_{i\pm\frac{1}{2},j,k}\big{(}\bm{q}^{+}_{i\pm\frac{1}{2},j,k}{-\,}\bm{q}^{-}_{i\pm\frac{1}{2},j,k}\big{)}\Big{]}, (45)

and similarly for 𝑯i,j±12,ky\bm{H}^{y}_{i,j\pm\frac{1}{2},k} and 𝑯i,j,k±12η\bm{H}^{\eta}_{i,j,k\pm\frac{1}{2}} after permuting the ±12\pm\frac{1}{2} in the grid indices and the corresponding spatial component (i.e. xyx\to y or η\eta).

The first two terms in Eq. (45) take the average of the currents extrapolated to the staggered cell face (i+12,j,k)(i{+}\frac{1}{2},j,k) (or (i12,j,k)(i{-}\frac{1}{2},j,k)) from the left (-) and right (++) sides. A first-order expression for the extrapolated currents can be computed using the chain rule:

𝑭i+12,j,kx\displaystyle\bm{F}^{x-}_{i+\frac{1}{2},j,k} =𝑭ijkx+Δx2[(xvx)ijk𝒒ijk+vijkx(x𝒒)ijk],\displaystyle=\bm{F}^{x}_{ijk}+\frac{\Delta x}{2}\Big{[}(\partial_{x}v^{x})_{ijk}\bm{q}_{ijk}+v^{x}_{ijk}(\partial_{x}\bm{q})_{ijk}\Big{]}\,, (46a)
𝑭i+12,j,kx+\displaystyle\bm{F}^{x+}_{i+\frac{1}{2},j,k} =𝑭i+1,j,kxΔx2[(xvx)i+1,j,k𝒒i+1,j,k+vi+1,j,kx(x𝒒)i+1,j,k],\displaystyle=\bm{F}^{x}_{i+1,j,k}-\frac{\Delta x}{2}\Big{[}(\partial_{x}v^{x})_{i+1,j,k}\bm{q}_{i+1,j,k}+v^{x}_{i+1,j,k}(\partial_{x}\bm{q})_{i+1,j,k}\Big{]}\,, (46b)
𝑭i12,j,kx\displaystyle\bm{F}^{x-}_{i-\frac{1}{2},j,k} =𝑭i1,j,kx+Δx2[(xvx)i1,j,k𝒒i1,j,k+vi1,j,kx(x𝒒)i1,j,k],\displaystyle=\bm{F}^{x}_{i-1,j,k}+\frac{\Delta x}{2}\Big{[}(\partial_{x}v^{x})_{i-1,j,k}\bm{q}_{i-1,j,k}+v^{x}_{i-1,j,k}(\partial_{x}\bm{q})_{i-1,j,k}\Big{]}\,, (46c)
𝑭i12,j,kx+\displaystyle\bm{F}^{x+}_{i-\frac{1}{2},j,k} =𝑭ijkxΔx2[(xvx)ijk𝒒ijk+vijkx(x𝒒)ijk],\displaystyle=\bm{F}^{x}_{ijk}-\frac{\Delta x}{2}\Big{[}(\partial_{x}v^{x})_{ijk}\bm{q}_{ijk}+v^{x}_{ijk}(\partial_{x}\bm{q})_{ijk}\Big{]}\,, (46d)

where 𝑭ijkx=vijkx𝒒ijk\bm{F}^{x}_{ijk}=v^{x}_{ijk}\bm{q}_{ijk}. Similarly, the extrapolated currents 𝑭i,j±12,ky+\bm{F}^{y+}_{i,j\pm\frac{1}{2},k}, 𝑭i,j±12,ky\bm{F}^{y-}_{i,j\pm\frac{1}{2},k}, 𝑭i,j,k±12η+\bm{F}^{\eta+}_{i,j,k\pm\frac{1}{2}} and 𝑭i,j,k±12η\bm{F}^{\eta-}_{i,j,k\pm\frac{1}{2}} at the remaining faces of the staggered cell are obtained by permuting the ±12\pm\frac{1}{2} (or ±1\pm 1) in the grid indices, the spatial components and derivatives, and the lattice spacing.

The final term in Eq. (45) takes into account the wave propagation of the discontinuities 𝒒i±12,j,k+𝒒i±12,j,k\bm{q}^{+}_{i\pm\frac{1}{2},j,k}-\bm{q}^{-}_{i\pm\frac{1}{2},j,k} at a finite speed [75]. We define the local propagation speed component si±12,j,kxs^{x}_{i\pm\frac{1}{2},j,k} at the staggered cell faces (i±12,j,k)(i{\pm}\frac{1}{2},j,k) as

si±12,j,kx=max(|vi±12,j,kx|,|vi±12,j,kx+|),s^{x}_{i\pm\frac{1}{2},j,k}=\max\big{(}|v^{x-}_{i\pm\frac{1}{2},j,k}|,|v^{x+}_{i\pm\frac{1}{2},j,k}|\big{)}\,, (47)

where the extrapolated velocities are

vi+12,j,kx\displaystyle v^{x-}_{i+\frac{1}{2},j,k} =vijkx+Δx2(xvx)ijk,\displaystyle=v^{x}_{ijk}+\frac{\Delta x}{2}(\partial_{x}v^{x})_{ijk}\,, (48a)
vi+12,j,kx+\displaystyle v^{x+}_{i+\frac{1}{2},j,k} =vi+1,j,kxΔx2(xvx)i+1,j,k,\displaystyle=v^{x}_{i+1,j,k}-\frac{\Delta x}{2}(\partial_{x}v^{x})_{i+1,j,k}\,, (48b)
vi12,j,kx\displaystyle v^{x-}_{i-\frac{1}{2},j,k} =vi1,j,kx+Δx2(xvx)i1,j,k,\displaystyle=v^{x}_{i-1,j,k}+\frac{\Delta x}{2}(\partial_{x}v^{x})_{i-1,j,k}\,, (48c)
vi12,j,kx+\displaystyle v^{x+}_{i-\frac{1}{2},j,k} =vijkxΔx2(xvx)ijk.\displaystyle=v^{x}_{ijk}-\frac{\Delta x}{2}(\partial_{x}v^{x})_{ijk}\,. (48d)

The discontinuities 𝒒i±12,j,k+𝒒i±12,j,k\bm{q}^{+}_{i\pm\frac{1}{2},j,k}-\bm{q}^{-}_{i\pm\frac{1}{2},j,k} propagating from the staggered cell faces (i±12,j,k)(i{\pm}\frac{1}{2},j,k) depend on the extrapolated dynamical variables

𝒒i+12,j,k\displaystyle\bm{q}^{-}_{i+\frac{1}{2},j,k} =𝒒ijk+Δx2(x𝒒)ijk,\displaystyle=\bm{q}_{ijk}+\frac{\Delta x}{2}(\partial_{x}\bm{q})_{ijk}\,, (49a)
𝒒i+12,j,k+\displaystyle\bm{q}^{+}_{i+\frac{1}{2},j,k} =𝒒i+1,j,kΔx2(x𝒒)i+1,j,k,\displaystyle=\bm{q}_{i+1,j,k}-\frac{\Delta x}{2}(\partial_{x}\bm{q})_{i+1,j,k}\,, (49b)
𝒒i12,j,k\displaystyle\bm{q}^{-}_{i-\frac{1}{2},j,k} =𝒒i1,j,k+Δx2(x𝒒)i1,j,k,\displaystyle=\bm{q}_{i-1,j,k}+\frac{\Delta x}{2}(\partial_{x}\bm{q})_{i-1,j,k}\,, (49c)
𝒒i12,j,k+\displaystyle\bm{q}^{+}_{i-\frac{1}{2},j,k} =𝒒ijkΔx2(x𝒒)ijk.\displaystyle=\bm{q}_{ijk}-\frac{\Delta x}{2}(\partial_{x}\bm{q})_{ijk}\,. (49d)

The formulae for the local propagation speed components si,j±12,kys^{y}_{i,j\pm\frac{1}{2},k} and si,j,k±12ηs^{\eta}_{i,j,k\pm\frac{1}{2}}, extrapolated velocities vi,j±12,ky+v^{y+}_{i,j\pm\frac{1}{2},k}, vi,j±12,kyv^{y-}_{i,j\pm\frac{1}{2},k}, vi,j,k±12η+v^{\eta+}_{i,j,k\pm\frac{1}{2}} and vi,j,k±12ηv^{\eta-}_{i,j,k\pm\frac{1}{2}}, and extrapolated dynamical variables 𝒒i,j±12,k+\bm{q}^{+}_{i,j\pm\frac{1}{2},k}, 𝒒i,j±12,k\bm{q}^{-}_{i,j\pm\frac{1}{2},k}, 𝒒i,j,k±12+\bm{q}^{+}_{i,j,k\pm\frac{1}{2}} and 𝒒i,j,k±12\bm{q}^{-}_{i,j,k\pm\frac{1}{2}} are analogous.

The numerical spatial derivatives appearing in the extrapolated quantities (46a-d), (48a-d) and (49a-d) are computed with a minmod flux limiter [75]:

(x𝒒)ijk\displaystyle(\partial_{x}\bm{q})_{ijk} =(Θ𝒒ijk𝒒i1,j,kΔx,𝒒i+1,j,k𝒒i1,j,k2Δx,Θ𝒒i+1,j,k𝒒ijkΔx),\displaystyle=\mathcal{M}\Big{(}\Theta\,\frac{\bm{q}_{ijk}-\bm{q}_{i-1,j,k}}{\Delta x},\frac{\bm{q}_{i+1,j,k}-\bm{q}_{i-1,j,k}}{2\Delta x},\Theta\,\frac{\bm{q}_{i+1,j,k}-\bm{q}_{ijk}}{\Delta x}\Big{)}\,, (50a)
(xvx)ijk\displaystyle(\partial_{x}v^{x})_{ijk} =(Θvijkxvi1,j,kxΔx,vi+1,j,kxvi1,j,kx2Δx,Θvi+1,j,kxvijkxΔx),\displaystyle=\mathcal{M}\Big{(}\Theta\,\frac{v^{x}_{ijk}-v^{x}_{i-1,j,k}}{\Delta x},\frac{v^{x}_{i+1,j,k}-v^{x}_{i-1,j,k}}{2\Delta x},\Theta\,\frac{v^{x}_{i+1,j,k}-v^{x}_{ijk}}{\Delta x}\Big{)}\,, (50b)

where

(a,b,c)=minmod(a,minmod(b,c)),\mathcal{M}(a,b,c)=\mathrm{minmod}(a,\mathrm{minmod}(b,c))\,, (51)

with

minmod(a,b)=sgn(a)+sgn(b)2×min(|a|,|b|);\mathrm{minmod}(a,b)=\frac{\mathrm{sgn}(a)+\mathrm{sgn}(b)}{2}\times\min(|a|,|b|)\,; (52)

the flux limiter parameter is set to Θ=1.8\Theta=1.8 [82]. The flux limiter derivatives (y𝒒)ijk(\partial_{y}\bm{q})_{ijk}, (yvy)ijk(\partial_{y}v^{y})_{ijk}, (η𝒒)ijk(\partial_{\eta}\bm{q})_{ijk} and (ηvη)ijk(\partial_{\eta}v^{\eta})_{ijk} are analogous.

In contrast, the numerical spatial derivatives appearing in the source terms 𝑺ijk\bm{S}_{ijk} are approximated with second-order central differences [20, 76]:

(x𝒒)ijk\displaystyle(\partial_{x}\bm{q})_{ijk} =𝒒i+1,j,k𝒒i1,j,k2Δx,\displaystyle=\frac{\bm{q}_{i+1,j,k}-\bm{q}_{i-1,j,k}}{2\Delta x}\,, (53a)
(x𝒖)ijk\displaystyle(\partial_{x}\bm{u})_{ijk} =𝒖i+1,j,k𝒖i1,j,k2Δx,\displaystyle=\frac{\bm{u}_{i+1,j,k}-\bm{u}_{i-1,j,k}}{2\Delta x}\,, (53b)

and similarly for (y𝒒)ijk(\partial_{y}\bm{q})_{ijk}, (y𝒖)ijk(\partial_{y}\bm{u})_{ijk}, (η𝒒)ijk(\partial_{\eta}\bm{q})_{ijk} and (η𝒖)ijk(\partial_{\eta}\bm{u})_{ijk}. The fluid velocity’s time derivative (τ𝒖)ijk(\partial_{\tau}\bm{u})_{ijk} also appears in the source terms; its evaluation will be discussed the next subsection.

3.2 Two-stage Runge–Kutta scheme

Because the KT algorithm (44) admits a semi-discrete form [75] (i.e. the time derivative is continuous while the spatial derivatives are discrete), we can combine it with an RK2 ODE solver to evolve the system in time [16, 20]. Given the dynamical variables 𝒒n,ijk𝒒ijk(τn)\bm{q}_{n,ijk}\equiv\bm{q}_{ijk}(\tau_{n}) at time τ=τn\tau=\tau_{n} (discrete times are labeled with index nn), we evolve the system one time step Δτn\Delta\tau_{n} with an intermediate Euler step (omitting the spatial indices):

𝒒I,n+1=𝒒n+Δτn𝑬(τn,𝒒n,𝒖n,n;𝒖n1,Δτn1),\bm{q}_{\,\text{I},n+1}=\bm{q}_{n}+\Delta\tau_{n}\bm{E}(\tau_{n},\bm{q}_{n},\bm{u}_{n},\mathcal{E}_{n};\bm{u}_{n-1},\Delta\tau_{n-1})\,, (54)

where 𝑬=τ𝒒\bm{E}=\partial_{\tau}\bm{q} is evaluated with r.h.s of Eq. (44). The time derivative τ𝒖\partial_{\tau}\bm{u} in the source terms is approximated with a first-order backward difference [20, 17]:

(τ𝒖)n=𝒖n𝒖n1Δτn1,(\partial_{\tau}\bm{u})_{n}=\frac{\bm{u}_{n}-\bm{u}_{n-1}}{\Delta\tau_{n-1}}\,, (55)

where 𝒖n1\bm{u}_{n-1} is the previous fluid velocity and Δτn1\Delta\tau_{n-1} is the previous time step.212121At the start of the hydrodynamic simulation, n=0n=0 or τ=τ0\tau=\tau_{0}, the previous time step Δτn1\Delta\tau_{n-1} is set to the current time step Δτn\Delta\tau_{n}. Unless stated otherwise, we also initialize the previous fluid velocity as 𝒖n1=𝒖n\bm{u}_{n-1}=\bm{u}_{n}. From the intermediate variables (54), we reconstruct the inferred variables (I,n+1,𝒖I,n+1)(\mathcal{E}_{\text{I},n+1},\bm{u}_{\text{I},n+1}) as well as the anisotropic variables (ΛI,n+1\Lambda_{\text{I},n+1} α,I,n+1\alpha_{\perp,\text{I},n+1}, αL,I,n+1\alpha_{L,\text{I},n+1}), as described in Secs. 3.3 and 3.4 below. Afterwards, we regulate the mean field BB and residual shear stresses WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}} in 𝒒I,n+1\bm{q}_{\,\text{I},n+1} (see Sec. 3.5) and set the ghost cell boundary conditions for 𝒒I,n+1\bm{q}_{\,\text{I},n+1} and 𝒖I,n+1\bm{u}_{\text{I},n+1}.

Next, we evolve the system with a second intermediate Euler step

𝑸n+2=𝒒I,n+1+Δτn𝑬(τn+Δτn,𝒒I,n+1,𝒖I,n+1,I,n+1;𝒖n,Δτn),\bm{Q}_{n+2}=\bm{q}_{\,\text{I},n+1}+\Delta\tau_{n}\bm{E}(\tau_{n}{+}\Delta\tau_{n},\bm{q}_{\,\text{I},n+1},\bm{u}_{\text{I},n+1},\mathcal{E}_{\text{I},n+1};\bm{u}_{n},\Delta\tau_{n})\,, (56)

where the fluid velocity’s time derivative is now evaluated as

(τ𝒖)I,n+1=𝒖I,n+1𝒖nΔτn.(\partial_{\tau}\bm{u})_{\text{I},n+1}=\frac{\bm{u}_{\text{I},n+1}-\bm{u}_{n}}{\Delta\tau_{n}}\,. (57)

In the RK2 scheme, we average the two intermediate Euler steps in 𝑸n+2\bm{Q}_{n+2} to update the dynamical variables at τ=τn+Δτn\tau=\tau_{n}+\Delta\tau_{n}:

𝒒n+1=𝒒n+𝑸n+22.\bm{q}_{n+1}=\frac{\bm{q}_{n}+\bm{Q}_{n+2}}{2}\,. (58)

From this, we update the inferred variables (n+1\mathcal{E}_{n+1}, 𝒖n+1\bm{u}_{n+1}) and (Λn+1\Lambda_{n+1}, α,n+1\alpha_{\perp,n+1}, αL,n+1\alpha_{L,n+1}) and regulate the residual shear stresses and mean field. Finally, we set the ghost cell boundary conditions for 𝒒n+1\bm{q}_{n+1} and 𝒖n+1\bm{u}_{n+1} and proceed with the next RK2 iteration.

3.3 Reconstructing the energy density and fluid velocity

Given the hydrodynamic variables TτμT^{\tau\mu}, along with 𝒫L\mathcal{P}_{L}, 𝒫\mathcal{P}_{\perp}, WzμW_{\perp z}^{\mu} and πτμ\pi_{\perp}^{\tau\mu}, we can reconstruct the energy density and fluid velocity from Eq. (4). The solution for the energy density is [66]

=Mτττ(Mx)2+(My)2Mτ+𝒫ττ(τMη)2(Mτ+𝒫ττ)(Mτ+𝒫L)2,\mathcal{E}=M^{\tau}-\mathcal{L}^{\tau\tau}-\frac{(M^{x})^{2}{+}(M^{y})^{2}}{M^{\tau}{+}\mathcal{P}_{\perp}{-}\mathcal{L}^{\tau\tau}}-\frac{(\tau M^{\eta})^{2}(M^{\tau}{+}\mathcal{P}_{\perp}{-}\mathcal{L}^{\tau\tau})}{(M^{\tau}{+}\mathcal{P}_{L})^{2}}\,, (59)

where ττ=Δ𝒫(zτ)2\mathcal{L}^{\tau\tau}=\Delta\mathcal{P}(z^{\tau})^{2} was defined earlier and

Mμ=Tτμ2Wz(τzμ)πτμ.M^{\mu}=T^{\tau\mu}-2W^{(\tau}_{\perp z}z^{\mu)}-\pi_{\perp}^{\tau\mu}\,. (60)

The reconstruction formula (59) also requires the components zτz^{\tau}, zηz^{\eta}, which can be expressed in terms of hydrodynamic variables as follows:222222Writing F=uη/uτF=u^{\eta}/u^{\tau} implies that the argument 1(τF)21-(\tau F)^{2} in Eq. (61) is always positive.

zτ=τF1(τF)2,zη=1τ1(τF)2.z^{\tau}=\frac{\tau F}{\sqrt{1-(\tau F)^{2}}}\,,\qquad z^{\eta}=\frac{1}{\tau\sqrt{1-(\tau F)^{2}}}\,. (61)

Here

F=AB1+τ2(B2A2)1+(τB)2,F=\frac{A-B\sqrt{1+\tau^{2}(B^{2}{-}A^{2})}}{1+(\tau B)^{2}}\,, (62)

with A=Kη/(Kτ+𝒫L)A=K^{\eta}/(K^{\tau}{+}\mathcal{P}_{L}) and B=Wzτ/(τ(Kτ+𝒫L))B=W_{\perp z}^{\tau}/(\tau(K^{\tau}{+}\mathcal{P}_{L})) where Kμ=TτμπτμK^{\mu}=T^{\tau\mu}{-}\pi_{\perp}^{\tau\mu}.

In the cold, dilute regions surrounding the fireball (and, occasionally, in cold spots within the fluctuating fireball), the energy density can become much smaller than the freezeout energy density sw\mathcal{E}_{\text{sw}}.232323In this work, we construct a particlization hypersurface of constant energy density sw=0.116\mathcal{E}_{\text{sw}}=0.116 GeV/fm3, which corresponds to the switching temperature Tsw=0.136T_{\text{sw}}=0.136 GeV. The lowest value used for the switching temperature in the JETSCAPE SIMS analysis is Tsw=0.135T_{\text{sw}}=0.135 GeV [1, 2]. While these regions are not phenomenologically important, hydrodynamic simulations are susceptible to crashing there without intervention [5, 20, 83]. To prevent this, we regulate the energy density with the formula

++mine+/min,\mathcal{E}\leftarrow\mathcal{E}_{+}+\mathcal{E}_{\text{min}}\,e^{-\mathcal{E}_{+}/\mathcal{E}_{\text{min}}}\,, (63)

where min\mathcal{E}_{\text{min}} is the minimum energy density allowed in the Eulerian grid and +=max(0,)\mathcal{E}_{+}=\max(0,\mathcal{E}). For min\mathcal{E}\gg\mathcal{E}_{\text{min}}, the regulation has virtually no effect on the energy density. As 0\mathcal{E}\to 0, however, the energy density is smoothly regulated to min\mathcal{E}_{\text{min}}. Ideally, min\mathcal{E}_{\text{min}} should be the lower limit of our QCD equation of state table low=3.5×104\mathcal{E}_{\text{low}}=3.5\times 10^{-4} GeV/fm3. However, we find that the code evolving non-conformal anisotropic hydrodynamics with fluctuating initial conditions encounters fewer technical difficulties if we instead use the larger value min=0.02\mathcal{E}_{\text{min}}=0.02 GeV/fm3, which is still about six times smaller than our choice for sw\mathcal{E}_{\text{sw}}.242424The constraint min\mathcal{E}\geq\mathcal{E}_{\text{min}} is not imposed for conformal systems. We checked that the regulation scheme (63) has little to no impact on the fluid’s dynamics in regions where sw\mathcal{E}\sim\mathcal{E}_{\text{sw}} or larger.

After regulating the energy density, we reconstruct the fluid velocity’s spatial components:

ux\displaystyle u^{x} =Mx(+𝒫)(Mτ+𝒫ττ),\displaystyle=\frac{M^{x}}{\sqrt{(\mathcal{E}+\mathcal{P}_{\perp})(M^{\tau}+\mathcal{P}_{\perp}-\mathcal{L}^{\tau\tau})}}\,, (64a)
uy\displaystyle u^{y} =My(+𝒫)(Mτ+𝒫ττ),\displaystyle=\frac{M^{y}}{\sqrt{(\mathcal{E}+\mathcal{P}_{\perp})(M^{\tau}+\mathcal{P}_{\perp}-\mathcal{L}^{\tau\tau})}}\,, (64b)
uη\displaystyle u^{\eta} =FMτ+𝒫ττ+𝒫.\displaystyle=F\sqrt{\frac{M^{\tau}+\mathcal{P}_{\perp}-\mathcal{L}^{\tau\tau}}{\mathcal{E}+\mathcal{P}_{\perp}}}\,. (64c)

Because of the regulation (63), there will be inconsistencies between the inferred variables (\mathcal{E}, 𝒖\bm{u}) and the components TτμT^{\tau\mu} but only in the dilute cold regions.

3.4 Reconstructing the anisotropic variables

To compute the non-conformal anisotropic transport coefficients in Sec. 2.6.3, we need to reconstruct the anisotropic variables

𝑿=(ΛααL)\bm{X}=\left(\begin{array}[]{c}\Lambda\\ \alpha_{\perp}\\ \alpha_{L}\end{array}\right) (65)

from \mathcal{E}, 𝒫L\mathcal{P}_{L}, 𝒫\mathcal{P}_{\perp}, BB and m()m(\mathcal{E}), by solving the system of equations [66]

𝒇(𝑿)=𝟎,\bm{f}(\bm{X})=\bm{0}\,, (66)

where

𝒇(𝑿)=(2000(𝑿)+B2200(𝑿)𝒫LB2010(𝑿)𝒫B).\bm{f}(\bm{X})=\left(\begin{array}[]{c}\mathcal{I}_{2000}(\bm{X})-\mathcal{E}+B\\ \mathcal{I}_{2200}(\bm{X})-\mathcal{P}_{L}-B\\ \mathcal{I}_{2010}(\bm{X})-\mathcal{P}_{\perp}-B\end{array}\right). (67)

We solve these nonlinear equations numerically using Newton’s method. Taking the anisotropic variables prior to the intermediate Euler step (54) as our initial guess252525At τ=τ0\tau=\tau_{0} we initialize the anisotropic variables (Λ0\Lambda_{0}, α,0\alpha_{\perp,0}, αL,0\alpha_{L,0}) by iterating 𝑿g=(T,1,1)T\bm{X}_{g}=(T,1,1)^{T} (where TT inside the parentheses is the temperature). 𝑿g=(Λn,α,n,αL,n)T\bm{X}_{g}=(\Lambda_{n},\alpha_{\perp,n},\alpha_{L,n})^{T}, we iterate 𝑿\bm{X} along the direction given by the Newton step

Δ𝑿=𝑱1𝒇,\Delta\bm{X}=-\bm{J}^{-1}\bm{f}\,, (68)

where 𝑱1\bm{J}^{-1} is the inverse of the Jacobian

𝑱=𝒇𝑿=(2001(𝑿)Λ224011(𝑿)Λα34201(𝑿)ΛαL32201(𝑿)Λ224211(𝑿)Λα34401(𝑿)ΛαL32011(𝑿)Λ244021(𝑿)Λα34211(𝑿)ΛαL3).\bm{J}=\frac{\partial\bm{f}}{\partial\bm{X}}=\left(\begin{array}[]{c c c}\dfrac{\mathcal{I}_{2001}(\bm{X})}{\Lambda^{2}}&\,\dfrac{2\,\mathcal{I}_{401-1}(\bm{X})}{\Lambda\alpha_{\perp}^{3}}&\,\dfrac{\mathcal{I}_{420-1}(\bm{X})}{\Lambda\alpha_{L}^{3}}\\ \\ \dfrac{\mathcal{I}_{2201}(\bm{X})}{\Lambda^{2}}&\,\dfrac{2\,\mathcal{I}_{421-1}(\bm{X})}{\Lambda\alpha_{\perp}^{3}}&\,\dfrac{\mathcal{I}_{440-1}(\bm{X})}{\Lambda\alpha_{L}^{3}}\\ \\ \dfrac{\mathcal{I}_{2011}(\bm{X})}{\Lambda^{2}}&\,\dfrac{4\,\mathcal{I}_{402-1}(\bm{X})}{\Lambda\alpha_{\perp}^{3}}&\,\dfrac{\mathcal{I}_{421-1}(\bm{X})}{\Lambda\alpha_{L}^{3}}\\ \end{array}\right). (69)

Specifically, we iterate the solution as

𝑿𝑿+λΔ𝑿,\bm{X}\leftarrow\bm{X}+\lambda\,\Delta\bm{X}\,, (70)

where λ[0,1]\lambda\in[0,1] is a partial step that is optimized with a line-backtracking algorithm to improve the global convergence of Newton’s method [84]. This procedure is repeated until 𝑿\bm{X} has converged to the solution of Eq. (66).262626In our simulation tests, we found several instances when Newton’s method failed to converge near the edges of the Eulerian grid. Whenever that happens we simply set the anisotropic variables to the initial guess 𝑿g\bm{X}_{g}.

For conformal systems (B=0B=0, m=0m=0, α=1\alpha_{\perp}=1), the anisotropic variables Λ\Lambda and αL\alpha_{L} are much easier to solve. The system of equations (67) reduces to

\displaystyle\mathcal{E} =3gΛ4200(αL)2π2,\displaystyle=\frac{3g\Lambda^{4}\mathcal{R}_{200}(\alpha_{L})}{2\pi^{2}}\,, (71a)
𝒫L\displaystyle\mathcal{P}_{L} =gΛ4220(αL)2π2,\displaystyle=\frac{g\Lambda^{4}\mathcal{R}_{220}(\alpha_{L})}{2\pi^{2}}\,, (71b)

where the functions 200\mathcal{R}_{200} and 220\mathcal{R}_{220} are listed in Appendix C. We numerically invert the longitudinal pressure to energy density ratio for αL\alpha_{L}:

220(αL)200(αL)=𝒫L.\frac{\mathcal{R}_{220}(\alpha_{L})}{\mathcal{R}_{200}(\alpha_{L})}=\frac{\mathcal{P}_{L}}{\mathcal{E}}\,. (72)

From this, we can evaluate Λ\Lambda as

Λ=(π23g200(αL))1/4.\Lambda=\bigg{(}\frac{\pi^{2}\mathcal{E}}{3g\mathcal{R}_{200}(\alpha_{L})}\bigg{)}^{1/4}\,. (73)

Because the solutions (72) and (73) do not require the previous values for Λ\Lambda and αL\alpha_{L} as input, we do not need to store them in memory during runtime.

3.5 Regulating the residual shear stresses and mean-field

After reconstructing the inferred variables, we regulate the residual shear stress components WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}} such that they satisfy the orthogonality and traceless conditions

uμWzμ\displaystyle u_{\mu}W_{\perp z}^{\mu} =0,\displaystyle=0\,, (74a)
zμWzμ\displaystyle z_{\mu}W_{\perp z}^{\mu} =0,\displaystyle=0\,, (74b)
uμπμν\displaystyle u_{\mu}\pi_{\perp}^{{\mu\nu}} =0,\displaystyle=0\,, (74c)
zμπμν\displaystyle z_{\mu}\pi_{\perp}^{{\mu\nu}} =0,\displaystyle=0\,, (74d)
π,μμ\displaystyle\pi^{\mu}_{\perp,\mu} =0,\displaystyle=0\,, (74e)

and that their overall magnitude is smaller than the longitudinal and transverse pressures:

π,μνπμν2Wz,μWzμ𝒫L2+2𝒫2.\sqrt{\pi_{\perp,{\mu\nu}}\pi_{\perp}^{\mu\nu}-2W_{\perp z,\mu}W_{\perp z}^{\mu}}\leq\sqrt{\mathcal{P}_{L}^{2}+2\mathcal{P}_{\perp}^{2}}\,. (75)

The former ensures that WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}} maintain their orthogonal and tracelessness properties within numerical accuracy during the hydrodynamic simulation. The latter prevents the residual shear stresses from overwhelming the anisotropic part of the energy-momentum tensor [49].

First, we update these components in the following order:

Wzτ\displaystyle W_{\perp z}^{\tau} uτ(Wzxux+Wzyuy)1+u2,\displaystyle\leftarrow\frac{u^{\tau}(W_{\perp z}^{x}u^{x}+W_{\perp z}^{y}u^{y})}{1+u_{\perp}^{2}}\,, (76a)
Wzη\displaystyle W_{\perp z}^{\eta} Wzτuηuτ,\displaystyle\leftarrow\frac{W_{\perp z}^{\tau}u^{\eta}}{u^{\tau}}\,, (76b)
πyy\displaystyle\pi_{\perp}^{yy} 2πxyuxuyπxx(1+(uy)2)1+(ux)2,\displaystyle\leftarrow\frac{2\pi_{\perp}^{xy}u^{x}u^{y}-\pi_{\perp}^{xx}\left(1{+}(u^{y})^{2}\right)}{1+(u^{x})^{2}}\,, (76c)
πτx\displaystyle\pi_{\perp}^{\tau x} uτ(πxxux+πxyuy)1+u2,\displaystyle\leftarrow\frac{u^{\tau}(\pi_{\perp}^{xx}u^{x}+\pi_{\perp}^{xy}u^{y})}{1+u_{\perp}^{2}}\,, (76d)
πτy\displaystyle\pi_{\perp}^{\tau y} uτ(πxyux+πyyuy)1+u2,\displaystyle\leftarrow\frac{u^{\tau}(\pi_{\perp}^{xy}u^{x}+\pi_{\perp}^{yy}u^{y})}{1+u_{\perp}^{2}}\,, (76e)
πxη\displaystyle\pi_{\perp}^{x\eta} πτxuηuτ,\displaystyle\leftarrow\frac{\pi_{\perp}^{\tau x}u^{\eta}}{u^{\tau}}\,, (76f)
πyη\displaystyle\pi_{\perp}^{y\eta} πτyuηuτ,\displaystyle\leftarrow\frac{\pi_{\perp}^{\tau y}u^{\eta}}{u^{\tau}}\,, (76g)
πτη\displaystyle\pi_{\perp}^{\tau\eta} uτ(πxηux+πyηuy)1+u2,\displaystyle\leftarrow\frac{u^{\tau}(\pi_{\perp}^{x\eta}u^{x}+\pi_{\perp}^{y\eta}u^{y})}{1+u_{\perp}^{2}}\,, (76h)
πηη\displaystyle\pi_{\perp}^{\eta\eta} πτηuηuτ,\displaystyle\leftarrow\frac{\pi_{\perp}^{\tau\eta}u^{\eta}}{u^{\tau}}\,, (76i)
πττ\displaystyle\pi_{\perp}^{\tau\tau} πτxux+πτyuy+τ2πτηuηuτ,\displaystyle\leftarrow\frac{\pi_{\perp}^{\tau x}u^{x}+\pi_{\perp}^{\tau y}u^{y}+\tau^{2}\pi_{\perp}^{\tau\eta}u^{\eta}}{u^{\tau}}\,, (76j)

while leaving the components WzxW_{\perp z}^{x}, WzyW_{\perp z}^{y}, πxx\pi_{\perp}^{xx} and πxy\pi_{\perp}^{xy} unchanged. Next, we rescale all the components by the same factor ρreg[0,1]\rho_{\text{reg}}\in[0,1]:

Wzμ\displaystyle W_{\perp z}^{\mu} ρregWzμ,\displaystyle\leftarrow\rho_{\text{reg}}W_{\perp z}^{\mu}\,, (77a)
πμν\displaystyle\pi_{\perp}^{{\mu\nu}} ρregπμν,\displaystyle\leftarrow\rho_{\text{reg}}\pi_{\perp}^{{\mu\nu}}\,, (77b)

where272727For longitudinally boost-invariant systems, we replace the second argument in Eq. (78) by 2𝒫2/ππ\sqrt{2\mathcal{P}_{\perp}^{2}/\pi_{\perp}{\cdot\,}\pi_{\perp}}.

ρreg=min(1,𝒫L2+2𝒫2ππ2WzWz),\rho_{\text{reg}}=\min\Bigg{(}1,\sqrt{\frac{\mathcal{P}_{L}^{2}+2\mathcal{P}_{\perp}^{2}}{\pi_{\perp}{\cdot\,}\pi_{\perp}-2W_{\perp z}{\,\cdot\,}W_{\perp z}}}\Bigg{)}\,, (78)

with ππ=π,μνπμν\pi_{\perp}{\cdot\,}\pi_{\perp}=\pi_{\perp,{\mu\nu}}\pi_{\perp}^{\mu\nu}. In the validation tests discussed in Sec. 4, we did not encounter a situation where the residual shear stresses are suppressed by Eq. (78). This indicates that the residual shear stresses are naturally smaller than the leading-order anisotropic pressures during the simulation. Nevertheless, we keep this procedure in place as a precaution.

For hydrodynamic simulations with two or three spatial dimensions, we find that the relaxation equation (24) for the mean-field BB can become unstable in certain spacetime regions (T0.150.16T\sim 0.15-0.16 GeV), causing the bulk viscous pressure Π\Pi to grow positive without bound. The origin of this issue is not fully understood, and we will address it in future work. For now, we remove this instability by regulating the nonequilibrium mean-field component δB=BBeq\delta B=B-B_{\text{eq}} as

δBκregδB,\delta B\leftarrow\kappa_{\text{reg}}\delta B, (79)

where

κreg=min(1,|Beq|δB)δB<0.\kappa_{\text{reg}}=\min\Big{(}1,-\frac{|B_{\text{eq}}|}{\delta B}\Big{)}\indent\forall\,\,\delta B<0\,. (80)

In practice, we only find it necessary to regulate the mean-field when δB<0\delta B<0 (i.e. in regions with Π>0\Pi>0).

3.6 Adaptive time step

Most relativistic hydrodynamic codes that simulate heavy-ion collisions evolve the system with a fixed time step Δτn=Δτ\Delta\tau_{n}=\Delta\tau [16, 5]. Any choice for Δτ\Delta\tau must satisfy the CFL condition so that the hydrodynamic simulation is at least dynamically stable [75]. However, the time step must also be small enough to resolve the fluid’s evolution rate (in particular, the large longitudinal expansion rate θL1/τ\theta_{L}\sim 1/\tau at early times) but, on the other hand, large enough to finish the simulation within a reasonable runtime. This balancing act places a practical limit on how early the user can start the hydrodynamic simulation (in practice, the smallest value typically used is τ00.2\tau_{0}\sim 0.2 fm/cc [26]). This is not much of a concern for second-order viscous hydrodynamics since it is anyhow prone to breaking down for very early initialization times, as discussed in Sec. 1. Anisotropic hydrodynamics, on the other hand, can handle the large pressure anisotropies occurring at early times much better; to realize its full potential it should be initialized at earlier times, but for that we need to move away from a fixed time step. In this section, we introduce a new adaptive stepsize method, which automatically adjusts the successive time step Δτn+1\Delta\tau_{n+1} to be larger or smaller than Δτn\Delta\tau_{n} in such a way that we can push back our fluid dynamical simulation to very early times (τ00.010.05fm/c\tau_{0}\sim 0.01-0.05\,\text{fm}/c) without sacrificing numerical accuracy nor computational efficiency.282828This is also useful for constructing the particlization hypersurface in peripheral heavy-ion collisions or small collision systems (e.g. p+p), whose fireball lifetimes are not that much longer than the hydrodynamization time τhydro\tau_{\text{hydro}} (see Sec. 1).

For a system with no source terms (𝑺=𝟎\bm{S}=\bm{0}) the KT algorithm is dynamically stable as long as the time step satisfies the CFL condition [75]

ΔτnΔτCFL=18min(Δxsmaxx(τn),Δysmaxy(τn),Δηssmaxη(τn)),\Delta\tau_{n}\leq\Delta\tau_{\text{CFL}}=\frac{1}{8}\min\left(\frac{\Delta x}{s^{x}_{\text{max}}(\tau_{n})},\frac{\Delta y}{s^{y}_{\text{max}}(\tau_{n})},\frac{\Delta\eta_{s}}{s^{\eta}_{\text{max}}(\tau_{n})}\right)\,, (81)

where smaxi(τn)s^{i}_{\text{max}}(\tau_{n}) (i=x,y,ηi=x,y,\eta) are the maximum local propagation speed components on the Eulerian grid at time τn\tau_{n}. In Milne spacetime, the quark-gluon plasma’s flow profile is not ultrarelativistic throughout most of its evolution (as long as one uses a QCD equation of state). Thus, the criterium (81) allows one to maintain dynamical stability with an adaptive time step that is generally larger than the fixed time step obtained by taking the limit smaxi1s^{i}_{\text{max}}\to 1:

Δτn=18min(Δx,Δy,Δηs).\Delta\tau_{n}=\frac{1}{8}\min\left(\Delta x,\Delta y,\Delta\eta_{s}\right)\,. (82)

For 𝑺𝟎\bm{S}\neq\bm{0}, however, the time step ΔτCFL\Delta\tau_{\text{CFL}} from (81) is too coarse to resolve the fluid’s gradients and relaxation rates at early times τ<0.5\tau<0.5 fm/cc when the flow profile is very nonrelativistic. Therefore, at early times when the source terms are strongest, the adaptive time step should primarily depend on those source terms. The following implementation ensures that, as the source terms relax and the fluid velocity grows more relativistic over time, the adaptive time step naturally approaches the CFL bound (81).

First, we consider a homogeneous fluid undergoing Bjorken expansion (i.e. =Tττ\mathcal{E}=T^{\tau\tau} and 𝒖=𝟎\bm{u}=\bm{0}). The system of differential equations (13) for the dynamical variables 𝒒(τ)\bm{q}(\tau) reduces to

τ𝒒(τ)=𝑺(τ,𝒒),\partial_{\tau}\bm{q}(\tau)=\bm{S}(\tau,\bm{q}), (83)

and we are given the initial values 𝒒n\bm{q}_{n} and time step Δτn\Delta\tau_{n} at time τn\tau_{n}. We evolve the system one time step Δτn\Delta\tau_{n} using the RK2 scheme (58):

𝒒n+1=𝒒n+Δτn2(𝑺(τn,𝒒n)+𝑺(τn+Δτn,𝒒n+Δτn𝑺(τn,𝒒n)).\bm{q}_{n+1}=\bm{q}_{n}+\frac{\Delta\tau_{n}}{2}\left(\bm{S}(\tau_{n},\bm{q}_{n})+\bm{S}(\tau_{n}{+}\Delta\tau_{n},\bm{q}_{n}{+}\Delta\tau_{n}\bm{S}(\tau_{n},\bm{q}_{n})\right)\,. (84)

For the next iteration, we determine how much we need to adjust the time step Δτn+1\Delta\tau_{n+1}. Adaptive stepsize methods do this by estimating the local truncation error of each time step [84]. In our method, we approximate the local truncation error of the next intermediate Euler step

𝒒I,n+2=𝒒n+1+Δτn+1𝑺(τn+Δτn,𝒒n+1),\bm{q}_{\,\text{I},n+2}=\bm{q}_{n+1}+\Delta\tau_{n+1}\bm{S}(\tau_{n}{+}\Delta\tau_{n},\bm{q}_{n+1})\,, (85)

which is

ϵn+2=12(τ2𝒒)n+1Δτn+12+O(Δτn+13),\epsilon_{n+2}=\frac{1}{2}\mathinner{\!\left\lVert(\partial_{\tau}^{2}\bm{q})_{n+1}\right\rVert}\Delta\tau_{n+1}^{2}+O(\Delta\tau_{n+1}^{3})\,, (86)

where \mathinner{\!\left\lVert...\right\rVert} denotes the 2\ell^{2}-norm. If the second time derivative (τ2𝒒)n+1(\partial^{2}_{\tau}\bm{q})_{n+1} at τ=τn+Δτn\tau=\tau_{n}+\Delta\tau_{n} is known, we can set the local truncation error to the desired error tolerance (after dropping higher-order terms)

12(τ2𝒒)n+1Δτn+12=δ0×max(Nq1/2,𝒒I,n+2),\frac{1}{2}\mathinner{\!\left\lVert(\partial_{\tau}^{2}\bm{q})_{n+1}\right\rVert}\Delta\tau_{n+1}^{2}=\delta_{0}\times\max(N_{q}^{1/2},\,\mathinner{\!\left\lVert\bm{q}_{\,\text{I},n+2}\right\rVert})\,, (87)

where δ0\delta_{0} is the error tolerance parameter and NqN_{q} is the number of dynamical variables. It is reasonable to use absolute or relative errors when 𝒒I,n+2\mathinner{\!\left\lVert\bm{q}_{\,\text{I},n+2}\right\rVert} is small or large, respectively [84]. In this work, we set the error tolerance parameter to δ0=0.004\delta_{0}=0.004.

To obtain an expression for the second time derivative, we compute the next intermediate Euler step (85) using the old time step (denoted by \star):

𝒒I,n+2=𝒒n+1+Δτn𝑺(τn+Δτn,𝒒n+1).\bm{q}^{\star}_{\,\text{I},n+2}=\bm{q}_{n+1}+\Delta\tau_{n}\bm{S}(\tau_{n}{+}\Delta\tau_{n},\bm{q}_{n+1})\,. (88)

This allows us to approximate (τ2𝒒)n+1(\partial_{\tau}^{2}\bm{q})_{n+1} with central differences:

(τ2𝒒)n+1=2(𝒒I,n+22𝒒n+1+𝒒n)Δτn2+O(Δτn).(\partial_{\tau}^{2}\bm{q})_{n+1}=\frac{2(\bm{q}^{\star}_{\,\text{I},n+2}{-}2\bm{q}_{n+1}{+}\bm{q}_{n})}{\Delta\tau_{n}^{2}}+O(\Delta\tau_{n})\,. (89)

Compared to the usual central difference formula there is an additional factor of 22 that accounts for the local truncation error present in 𝒒I,n+2\bm{q}_{\,\text{I},n+2}^{\star}. Furthermore, the expression (89) is numerically accurate to O(Δτn)O(\Delta\tau_{n}) rather than O(Δτn2)O(\Delta\tau_{n}^{2}). After substituting Eqs. (85) and (89) in Eq. (87), one has for the next time step

Δτn+1=max(Δτn+1(abs),Δτn+1(rel)),\Delta\tau_{n+1}=\max\big{(}\Delta\tau_{n+1}^{\text{(abs)}},\Delta\tau_{n+1}^{\text{(rel)}}\big{)}\,, (90)

where

Δτn+1(abs)=Δτnδ0Nq1/2𝒒I,n+22𝒒n+1+𝒒n\Delta\tau_{n+1}^{\text{(abs)}}=\Delta\tau_{n}\sqrt{\dfrac{\delta_{0}\,N_{q}^{1/2}}{\mathinner{\!\left\lVert\bm{q}^{\star}_{\,\text{I},n+2}{-}2\bm{q}_{n+1}{+}\bm{q}_{n}\right\rVert}}} (91)

and Δτn+1(rel)\Delta\tau_{n+1}^{\text{(rel)}} is the numerical solution to the algebraic equation

Nq1/2(Δτn+1(rel))2(Δτn+1(abs))2=𝒒n+12+2𝒒n+1𝑺n+1Δτn+1(rel)+𝑺n+12(Δτn+1(rel))2,\frac{N_{q}^{1/2}\big{(}\Delta\tau_{n+1}^{\text{(rel)}}\big{)}^{2}}{\big{(}\Delta\tau_{n+1}^{\text{(abs)}}\big{)}^{2}}=\sqrt{\mathinner{\!\left\lVert\bm{q}_{n+1}\right\rVert}^{2}+2\bm{q}_{n+1}{\cdot\,}\bm{S}_{n+1}\Delta\tau_{n+1}^{\text{(rel)}}+\mathinner{\!\left\lVert\bm{S}_{n+1}\right\rVert}^{2}\big{(}\Delta\tau_{n+1}^{\text{(rel)}}\big{)}^{2}}\,, (92)

with 𝑺n+1=𝑺(τn+Δτn,𝒒n+1)\bm{S}_{n+1}=\bm{S}(\tau_{n}{+}\Delta\tau_{n},\bm{q}_{n+1}).

For the general case without Bjorken symmetry, 𝒒(x)\bm{q}(x) varies across the grid. We then perform the calculation (90) (replacing 𝑺\bm{S} by 𝑬\bm{E}) for all spatial grid points and take the minimum value. Afterwards, we place safety bounds to prevent the time step from changing too rapidly:

(1α)ΔτnΔτn+1(1+α)Δτn,\left(1{-}\alpha\right)\Delta\tau_{n}\leq\Delta\tau_{n+1}\leq(1{+}\alpha)\Delta\tau_{n}\,, (93)

where we set the control parameter to α=0.5\alpha=0.5. Finally, we impose the CFL bound (81) on Δτn+1\Delta\tau_{n+1}. With the new time step at hand, we resume computing the next RK2 iteration. Notice that there are no additional numerical evaluations of the flux and source terms in the adaptive RK2 scheme since we can recompute the next intermediate Euler step 𝒒I,n+2\bm{q}_{\,\text{I},n+2} simply by adjusting the time step in 𝒒I,n+2\bm{q}_{\,\text{I},n+2}^{\star}. This allows our adaptive time step algorithm to be readily integrated into our numerical scheme.

When we start the hydrodynamic simulation, we initialize the time step Δτ0\Delta\tau_{0} to be 20 times smaller than the initial time τ0\tau_{0}.292929We do not allow the adaptive time step to become any smaller than this (i.e. we impose Δτn0.05τ0\Delta\tau_{n}\geq 0.05\tau_{0}). At first, the adaptive time step Δτn\Delta\tau_{n} tends to increase with time since the fluid’s longitudinal expansion rate decreases. As the transverse flow builds up, Δτn\Delta\tau_{n} eventually becomes bounded by the CFL condition (81).

3.7 Program summary

Refer to caption
Figure 5: (Color online) Program flowchart of VAH.

We close this section by summarizing the workflow of the anisotropic fluid dynamical simulation with a QCD equation of state. Figure 5 shows the flowchart of the code’s primary mode,303030The secondary (test) mode outputs the hydrodynamic evolution from the simulation and their corresponding semi-analytic solutions, if they exist. which constructs a hypersurface of constant energy density sw\mathcal{E}_{\text{sw}} for a particlization module:

  1. 1.

    We read in the runtime parameters and configure the Eulerian grid.

  2. 2.

    We allocate memory to store the dynamical and inferred variables in the Eulerian grid at a given time τn\tau_{n}. Altogether, there are three dynamical variables (q, qI, Q), two fluid velocity variables (u, up), one energy density variable e and three anisotropic variables (lambda, aT, aL). They store one of the following variables during the RK2 iteration:

    1. (a)

      q holds the current or updated dynamical variables 𝒒n\bm{q}_{n} or 𝒒n+1\bm{q}_{n+1}.

    2. (b)

      qI holds the intermediate dynamical variables 𝒒I,n+1\bm{q}_{\,\text{I},n+1} (or 𝒒I,n+1\bm{q}^{\star}_{\,\text{I},n+1}).

    3. (c)

      Q holds the previous, updated or current dynamical variables 𝒒n1\bm{q}_{n-1}, 𝒒n+1\bm{q}_{n+1} or 𝒒n\bm{q}_{n}.

    4. (d)

      u holds the current, intermediate or updated fluid velocity 𝒖n\bm{u}_{n}, 𝒖I,n+1\bm{u}_{\text{I},n+1} or 𝒖n+1\bm{u}_{n+1}.

    5. (e)

      up holds the previous or current fluid velocity 𝒖n1\bm{u}_{n-1} or 𝒖n\bm{u}_{n}.

    6. (f)

      e holds the current, intermediate or updated energy density n\mathcal{E}_{n}, I,n+1\mathcal{E}_{\text{I},n+1} or n+1\mathcal{E}_{n+1}.

    7. (g)

      lambda, aT and aL hold the current, intermediate or updated anisotropic variables (Λn\Lambda_{n}, α,n\alpha_{\perp,n}, αL,n\alpha_{L,n}), (ΛI,n+1\Lambda_{\text{I},n+1}, α,I,n+1\alpha_{\perp,\text{I},n+1}, αL,I,n+1\alpha_{L,\text{I},n+1}) or (Λn+1\Lambda_{n+1}, α,n+1\alpha_{\perp,n+1}, αL,n+1\alpha_{L,n+1})

  3. 3.

    We initialize the variables q, u, up, e, lambda, aT and aL as follows:313131If the code is run with Gubser initial conditions, the hydrodynamic variables are initialized differently (see Sec. 4.2). first, we compute (or read in) the energy density profile given by an initial-state model (e.g. TR{}_{\text{\sc R}}ENTo [24]). Then we initialize the longitudinal and transverse pressures as

    𝒫L\displaystyle\mathcal{P}_{L} =3R𝒫eq2+R,\displaystyle=\frac{3R\,\mathcal{P}_{\text{eq}}}{2+R}\,, (94a)
    𝒫\displaystyle\mathcal{P}_{\perp} =3𝒫eq2+R,\displaystyle=\frac{3\mathcal{P}_{\text{eq}}}{2+R}\,, (94b)

    where R[0,1]R\in[0,1] is a pressure anisotropy ratio parameter set by the user (typically a small value R0.3R\leq 0.3). This model assumes that only the pressure anisotropy 𝒫L𝒫\mathcal{P}_{L}-\mathcal{P}_{\perp} enters in the initial 𝒫L\mathcal{P}_{L} and 𝒫\mathcal{P}_{\perp} while the initial bulk viscous pressure is Π=0\Pi=0. The initial fluid velocity is static (i.e. u=up=𝟎\texttt{u}=\texttt{up}=\bm{0}) and the residual shear stresses WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}} are set to zero. From this, we compute the components TτμT^{\tau\mu} with Eq. (4). Finally, we set the mean-field to B=BeqB=B_{\text{eq}} and initialize the anisotropic variables by solving Eq. (66).323232This initialization scheme is similar to how conformal free-streaming modules are initialized at τ00\tau_{0}\to 0 [50, 6, 7, 1, 2].

  4. 4.

    We set the ghost cell boundary conditions for q, e and u.333333The ghost cells of e are only used in conformal anisotropic hydrodynamics to approximate the source terms i\propto\partial_{i}\mathcal{E} on the grid’s faces. We also configure the freezeout finder and set the initial time step to Δτ0=0.05τ0\Delta\tau_{0}=0.05\,\tau_{0}.

  5. 5.

    We start the hydrodynamic evolution at τ=τ0\tau=\tau_{0} (or n=0n=0):

    1. (a)

      We call the freezeout finder every two time steps343434The user can adjust the number of time steps between freezeout finder calls. (i.e. nmod2=0n\mod 2=0) to search for freezeout cells between the Eulerian grids from the current and previous calls (if n=0n=0, we only load the initial grid to the freezeout finder). The hypersurface volume elements d3σμd^{3}\sigma_{\mu} and their spacetime positions are constructed with the code CORNELIUS [85]. The hydrodynamic variables at the hypersurface elements’ centroid are approximated with a 4d linear interpolation.

    2. (b)

      We compute the intermediate Euler step (54) and store the results 𝒒I,n+1\bm{q}_{\,\text{I},n+1} in qI (if n>0n>0, we compute 𝒒I,n+1\bm{q}^{\star}_{\,\text{I},n+1} and the adaptive time step Δτn\Delta\tau_{n} using the algorithm in Sec. 3.6 before evaluating 𝒒I,n+1\bm{q}_{\,\text{I},n+1}). Afterwards, we swap the variables uup\texttt{u}\leftrightarrow\texttt{up} so that up𝒖n\texttt{up}\leftarrow\bm{u}_{n}.

    3. (c)

      We reconstruct the intermediate inferred variables I,n+1\mathcal{E}_{\text{I},n+1}, 𝒖I,n+1\bm{u}_{\text{I},n+1}, ΛI,n+1\Lambda_{\text{I},n+1}, α,I,n+1\alpha_{\perp,\text{I},n+1} and αL,I,n+1\alpha_{L,\text{I},n+1} from qI and store the results in e, u, lambda, aT and aL, respectively. We regulate the residual shear stresses and mean-field in qI and set the ghost cell boundary conditions for qI, e and u.

    4. (d)

      We compute the second intermediate Euler step (56) and update the dynamical variables 𝒒n+1\bm{q}_{n+1} via Eq. (58), which is stored in Q. Afterwards, we swap the variables qQ\texttt{q}\leftrightarrow\texttt{Q} so that q𝒒n+1\texttt{q}\leftarrow\bm{q}_{n+1} and Q𝒒n\texttt{Q}\leftarrow\bm{q}_{n}.

    5. (e)

      We update the inferred variables n+1\mathcal{E}_{n+1}, 𝒖n+1\bm{u}_{n+1}, Λn+1\Lambda_{n+1}, α,n+1\alpha_{\perp,n+1} and αL,n+1\alpha_{L,n+1} from q and store the results in e, u, lambda, aT and aL, respectively. We regulate the residual shear stresses and mean-field in q and set the ghost cell boundary conditions for q, e and u.

    6. (f)

      Steps (a) – (e) are repeated until the temperature of all fluid cells in the grid are below the switching temperature (i.e. min(Tijk)<Tsw\min(T_{ijk})<T_{\text{sw}}).

  6. 6.

    After the hydrodynamic evolution, we deallocate the hydrodynamic variables and store the freezeout surface in memory, which can either be passed to another program or written to file.

4 Validation tests and hydrodynamic model comparisons

In this section we test the validity of our anisotropic fluid dynamical simulation for various initial-state configurations, using either the conformal or QCD equation of state. First, we run anisotropic hydrodynamics with conformal Bjorken and Gubser initial conditions [86, 87, 88], whose semi-analytic solutions can be computed accurately using a fourth-order Runge–Kutta ODE solver [69, 48]. After these two validation tests, we compare (3+1)–dimensional conformal anisotropic hydrodynamics and second-order viscous hydrodynamics353535The VAH code can also run second-order viscous hydrodynamics, where the kinetic transport coefficients are computed with either quasiparticle masses m(T)m(T) (see Fig. 2a) [79, 66] or light masses (i.e. m/T1m/T\ll 1[67, 20] (we use the same shear and bulk viscosities as in Fig. 3). In this work, we label the former model as quasiparticle viscous hydrodynamics (or VH) and the latter as standard viscous hydrodynamics (or VH2); for conformal systems, these two models are equivalent. For details on the numerical implementation of second-order viscous hydrodynamics, see Appendix E. for a central Pb+Pb collision with a smooth TR{}_{\text{\sc R}}ENTo initial condition [24].

Next, we run non-conformal anisotropic hydrodynamics and viscous hydrodynamics with Bjorken initial conditions and compare them to their semi-analytic solutions [66]. Finally, we study the differences between (3+1)–dimensional non-conformal anisotropic hydrodynamics and viscous hydrodynamics in central Pb+Pb collisions with smooth or fluctuating TR{}_{\text{\sc R}}ENTo initial conditions, with the goal of identifying situations where VAH offers definitive advantages in reliability over standard viscous hydrodynamics.

4.1 Conformal Bjorken flow test

For the first test, we evolve a conformal plasma undergoing Bjorken expansion [86]. In Bjorken flow, the system is longitudinally boost-invariant and homogeneous in the transverse plane. As a result, the fluid velocity uμ=(1,0,0,0)u^{\mu}=(1,0,0,0) is static and the residual shear stresses WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}} vanish. The only two degrees of freedom are the energy density =Tττ\mathcal{E}=T^{\tau\tau} and longitudinal pressure 𝒫L\mathcal{P}_{L} (the transverse pressure is 𝒫=12(𝒫L)\mathcal{P}_{\perp}=\frac{1}{2}(\mathcal{E}{-}\mathcal{P}_{L})). The anisotropic hydrodynamic equations simplify to [69]

τ\displaystyle\partial_{\tau}\mathcal{E} =+𝒫Lτ,\displaystyle=-\frac{\mathcal{E}+\mathcal{P}_{L}}{\tau}\,, (95a)
τ𝒫L\displaystyle\partial_{\tau}\mathcal{P}_{L} =3𝒫L3τπ+ζ¯zLτ,\displaystyle=\frac{\mathcal{E}{-}3\mathcal{P}_{L}}{3\tau_{\pi}}+\frac{\bar{\zeta}_{z}^{L}}{\tau}\,, (95b)

where the conformal transport coefficient ζ¯zL\bar{\zeta}_{z}^{L} is given in Eq. (35a).

Refer to caption
Figure 6: (Color online) Conformal Bjorken evolution of the normalized energy density /0\mathcal{E}/\mathcal{E}_{0} (a) and pressure ratio 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} (b) from the anisotropic hydrodynamic simulation (dashed red) and semi-analytic solution (transparent red). The gray lines in (a) show two different power laws for comparison. The subpanels at the bottom show the ratio between the numerical simulation and the semi-analytic solutions (solid red).
Refer to caption
Figure 7: (Color online) The evolution of the adaptive time step Δτn\Delta\tau_{n} (dashed red) in the conformal Bjorken simulation. The semi-analytic method (transparent red) uses a fixed time step of Δτn=5×104\Delta\tau_{n}=5\times 10^{-4} fm/cc.

We start the simulation363636Bjorken flow can be simulated on either one fluid cell (i.e. Nx=Ny=Nη=1)N_{x}=N_{y}=N_{\eta}=1) or a larger grid with homogeneous initial conditions. at τ0=0.01\tau_{0}=0.01 fm/cc with an initial temperature T0=1.05T_{0}=1.05 GeV, pressure ratio R=103R=10^{-3} and shear viscosity η/𝒮=0.2\eta/\mathcal{S}=0.2. We evolve the system with an adaptive time step, initially set to Δτ0=5×104\Delta\tau_{0}=5\times 10^{-4} fm/cc, until the temperature drops below Tsw=0.136T_{\text{sw}}=0.136 GeV at τf20\tau_{f}\sim 20 fm/cc. Figure 6 shows the evolution of the energy density normalized to its initial value 0\mathcal{E}_{0} and the pressure ratio 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp}. At very early times, the system is approximately free-streaming due to the rapid longitudinal expansion, resulting in a large Knudsen number and causing the energy density to decrease like 0τ0/τ\mathcal{E}\approx\mathcal{E}_{0}\tau_{0}/\tau. Over time, the longitudinal expansion rate θL=1/τ\theta_{L}=1/\tau decreases and the system approaches local equilibrium, i.e. τ4/3\mathcal{E}\propto\tau^{-4/3} and 𝒫L/𝒫1\mathcal{P}_{L}/\mathcal{P}_{\perp}\to 1.

We compare the (0+1)–d VAH simulation to the semi-analytic solution of (95), which uses a fixed time step Δτ=5×104\Delta\tau=5\times 10^{-4} fm/cc. The simulation is in excellent agreement with the semi-analytic solution, with numerical errors staying below 0.5%0.5\%. We also checked the convergence of the simulation curves as we decrease the error tolerance parameter δ0\delta_{0} in the adaptive time step algorithm. Figure 7 shows the evolution of the adaptive time step Δτn\Delta\tau_{n} for the parameters δ0=0.004\delta_{0}=0.004 and α=0.5\alpha=0.5. One sees that the time step increases linearly with time and is not bounded by the CFL condition (81) since the flow is stationary (i.e. ΔτCFL)\Delta\tau_{\text{CFL}}\to\infty). As a result, the adaptive RK2 scheme quickly reaches the switching temperature in 145 time steps, compared to about 40000 steps for the semi-analytic method.

4.2 Conformal Gubser flow test

In the next test, we evolve a conformal fluid subject to Gubser flow [88, 87], which is longitudinally boost-invariant and azimuthally symmetric. The fluid velocity profile takes the following analytic form:

uτ=coshκ,ux=sinhκcosϕ,uy=sinhκsinϕ,uη=0,u^{\tau}=\cosh\kappa\,,\quad u^{x}=\sinh\kappa\cos\phi\,,\quad u^{y}=\sinh\kappa\sin\phi\,,\quad u^{\eta}=0\,, (96)

where

κ(τ,r)=tanh1[2q2τr1+q2(τ2+r2)],\kappa(\tau,r)={\tanh^{-1}}\bigg{[}\frac{2q^{2}\tau r}{1+q^{2}(\tau^{2}{+}r^{2})}\bigg{]}\,, (97)

with r=x2+y2r{\,=\,}\sqrt{x^{2}{+}y^{2}} being the transverse radius, ϕ=tan1(y/x)\phi={\tan^{-1}}(y/x) the azimuthal angle, and qq an inverse length scale parameter that determines the fireball’s transverse size R1/qR_{\perp}\sim 1/q at τ=0+\tau=0^{+}.

One can transform the polar Milne coordinates x~μ=(τ,r,ϕ,ηs)\tilde{x}^{\mu}=(\tau,r,\phi,\eta_{s}) to the de Sitter coordinates x^μ=(ρ,θ,ϕ,ηs)\hat{x}^{\mu}=(\rho,\theta,\phi,\eta_{s}) where

ρ(τ,r)\displaystyle\rho(\tau,r) =sinh1[1q2(τ2r2)2qτ],\displaystyle=-{\sinh^{-1}}\bigg{[}\frac{1-q^{2}\left(\tau^{2}{-}r^{2}\right)}{2q\tau}\bigg{]}\,, (98a)
θ(τ,r)\displaystyle\theta(\tau,r) =tan1[2qr1+q2(τ2r2)].\displaystyle={\tan^{-1}}\bigg{[}\frac{2qr}{1+q^{2}\left(\tau^{2}{-}r^{2}\right)}\bigg{]}\,. (98b)

The line element in this de Sitter space possesses SO(3)qSO(1,1)Z2SO(3)_{q}\otimes SO(1,1)\otimes Z_{2} symmetry [88, 87]:

ds^2=dρ2+cosh2ρ(dθ2+sin2θdϕ2)+dηs2,d\hat{s}^{2}=-d\rho^{2}+{\cosh^{2}}\rho\,\bigl{(}d\theta^{2}{+}{\sin^{2}}\theta d\phi^{2}\bigr{)}+d\eta^{2}_{s}\,, (99)

which makes the fluid velocity u^μ=(1,0,0,0)\hat{u}^{\mu}=(1,0,0,0) stationary.373737A hat denotes a quantity in the de Sitter space x^μ=(ρ,θ,ϕ,ηs)\hat{x}^{\mu}=(\rho,\theta,\phi,\eta_{s}). All hatted quantities are made unitless by multiplying with appropriate powers of τ\tau. Hence, the hydrodynamic variables only depend on the de Sitter time ρ\rho, and the anisotropic hydrodynamic equations reduce to [48]

ρ^\displaystyle\partial_{\rho}\hat{\mathcal{E}} =(𝒫^L3^)tanhρ,\displaystyle=(\hat{\mathcal{P}}_{L}-3\hat{\mathcal{E}})\tanh\rho\,, (100a)
ρ𝒫^L\displaystyle\partial_{\rho}\hat{\mathcal{P}}_{L} =^3𝒫^L3τ^π(4𝒫^L+ζ¯^zL)tanhρ.\displaystyle=\frac{\hat{\mathcal{E}}{-}3\hat{\mathcal{P}}_{L}}{3\hat{\tau}_{\pi}}-\big{(}4\hat{\mathcal{P}}_{L}{+}\hat{\bar{\zeta}}_{z}^{L}\big{)}\tanh\rho\,. (100b)

The residual shear stresses W^zμ\hat{W}_{\perp z}^{\mu} and π^μν\hat{\pi}_{\perp}^{\mu\nu} vanish under Gubser symmetry [48]. In the semi-analytic solution (100), we start the evolution at ρ0=9.2\rho_{0}=-9.2, which corresponds to a corner in a 1414 fm ×14\times 14 fm transverse grid (r=72r=7\sqrt{2} fm) at the initial time τ0=0.01\tau_{0}=0.01 fm/cc, with an initial temperature T^0=0.0017\hat{T}_{0}=0.0017 and pressure ratio R^=103\hat{R}=10^{-3}. We set the inverse fireball size to q=1.0q=1.0 fm-1 and the shear viscosity to η/𝒮=0.2\eta/\mathcal{S}=0.2. We evolve the system with a fixed stepsize Δρ=104\Delta\rho=10^{-4} until ρf=1.1\rho_{f}=1.1, which corresponds to the center of our Eulerian grid (r=0r=0 fm) at τf=3.01\tau_{f}=3.01 fm/cc.

Next, we map the semi-analytic solution for the de Sitter energy density ^(ρ)\hat{\mathcal{E}}(\rho) and longitudinal pressure 𝒫^L(ρ)\hat{\mathcal{P}}_{L}(\rho) back to Milne coordinates [89]:

(τ,r)=^(ρ(τ,r))τ4,𝒫L(τ,r)=𝒫^L(ρ(τ,r))τ4.\mathcal{E}(\tau,r)=\frac{\hat{\mathcal{E}}\big{(}\rho(\tau,r)\big{)}}{\tau^{4}}\,,\qquad\mathcal{P}_{L}(\tau,r)=\frac{\hat{\mathcal{P}}_{L}\big{(}\rho(\tau,r)\big{)}}{\tau^{4}}\,. (101)

Refer to caption

Figure 8: (Color online) Conformal Gubser evolution of the energy density \mathcal{E} (GeV/fm3) (left column) and transverse fluid rapidity uu_{\perp} (right column) in the anisotropic hydrodynamic simulation.
Refer to caption
Figure 9: (Color online) Conformal Gubser evolution of (a) \mathcal{E} (GeV/fm3), (b) uxu^{x}, (c) 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp}, and (d) ππ/(𝒫2)\sqrt{\pi_{\perp}{\cdot\,}\pi_{\perp}}/(\mathcal{P}_{\perp}\sqrt{2}) along the xx–axis (y=0y=0), given by the anisotropic hydrodynamic simulation (dashed color) and semi-analytic solution (transparent color). The subpanels below panels (a-c) show the ratio between numerical simulation and semi-analytic solution (solid color).

We use this map to set the initial energy density and longitudinal pressure profiles (τ0,x,y)\mathcal{E}(\tau_{0},x,y) and 𝒫L(τ0,x,y)\mathcal{P}_{L}(\tau_{0},x,y); the initial temperature at the center of the fireball is T0,center=1.05T_{0,\text{center}}=1.05 GeV. We set the initial transverse shear stress to πμν=0\pi_{\perp}^{{\mu\nu}}=0. Finally, we use Eq. (96) to initialize the current fluid velocity uuμ(τ0,x,y)\texttt{u}\leftarrow u^{\mu}(\tau_{0},x,y) and previous fluid velocity upuμ(τ0Δτ0,x,y)\texttt{up}\leftarrow u^{\mu}(\tau_{0}-\Delta\tau_{0},x,y), where the initial time step is Δτ0=5×104\Delta\tau_{0}=5\times 10^{-4} fm/cc. We evolve the Gubser profile on a 1414 fm ×14\times 14 fm transverse grid with a lattice spacing of Δx=Δy=0.05\Delta x=\Delta y=0.05 fm.

Figure 8 shows the evolution of the energy density and transverse fluid velocity u=(ux)2+(uy)2u_{\perp}=\sqrt{(u^{x})^{2}+(u^{y})^{2}} in the transverse plane at various time frames.383838In order to output the hydrodynamic quantities at specific times, we readjust the adaptive time step whenever we are close to these output times (this is not shown in Fig. 10). At early times, the very hot and compact fireball expands rapidly along the longitudinal direction and quickly cools down without much change to its transverse shape. Over time, the transverse expansion overtakes the longitudinal expansion, pushing the medium radially outward as a ring of fire. One sees that the energy density and transverse fluid velocity maintain their azimuthal symmetry throughout the evolution. However, there are some numerical fluctuations around the lines y=±xy=\pm\,x at later times, especially near the peak of uu_{\perp}. This is a consequence of using a Cartesian grid with a finite spatial resolution.

Figure 9 shows the evolution of the energy density \mathcal{E}, fluid velocity component uxu^{x}, pressure anisotropy ratio 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} and transverse shear inverse Reynolds number ππ\sqrt{\pi_{\perp}{\cdot\,}\pi_{\perp}} / (𝒫2)(\mathcal{P}_{\perp}\sqrt{2}) along the xx–axis at multiple time frames. Overall, there is very good agreement between the simulation and semi-analytic solution, except near the local extrema of uxu^{x}, where the numerical errors are about 12%1{-}2\%; these errors can be reduced by using a finer lattice spacing. The transverse shear stress πμν\pi_{\perp}^{{\mu\nu}} should remain zero throughout the simulation. However, the transverse shear inverse Reynolds number from the VAH simulation is nonzero (albeit small, 2%\lesssim 2\%) since the Eulerian grid does not perfectly preserve Gubser symmetry.

Refer to caption
Figure 10: (Color online) The evolution of the adaptive time step Δτn\Delta\tau_{n} (dashed red) and CFL bound ΔτCFL\Delta\tau_{\text{CFL}} (transparent red) in the conformal Gubser simulation.

Figure 10 shows the evolution of the adaptive time step in the Gubser simulation. In contrast to Fig. 7, here Δτn\Delta\tau_{n} becomes bounded by the CFL condition (81) at τ0.4\tau\sim 0.4 fm/cc due to the increasing transverse expansion rate. As a result, the Gubser simulation finishes in 407 time steps, compared to 480 steps if we had used the fixed time step Δτ=Δx/8\Delta\tau=\Delta x/8. Although we only gain a slight 1.15×1.15\times speedup, the adaptive time step is able to resolve the fluid’s early-time dynamics more accurately than the fixed time step (82) because it is initially independent of the lattice spacing. This property is especially useful when simulating more realistic nuclear collisions, where the lattice spacing required to resolve the participant nucleons’ transverse energy deposition is several times coarser than the one used in this test (e.g. Δx=Δy0.10.3\Delta x=\Delta y\sim 0.1-0.3 fm).

4.3 (3+1)–d conformal hydrodynamic models in central Pb+Pb collisions

Here we compare (3+1)–dimensional conformal anisotropic hydrodynamics VAH to conformal second-order viscous hydrodynamics VH (see Appendix E), for a central Pb+Pb collision at zero impact parameter (b=0b=0 fm) with static (in Milne coordinates) and smooth initial conditions. To generate the latter we use an azimuthally symmetric TR{}_{\text{\sc R}}ENTo energy density profile averaged over 2000 fluctuating events [24] and extended along the ηs\eta_{s}–direction with a smooth rapidity plateau [76] (see Appendix D for more information on the TR{}_{\text{\sc R}}ENTo energy deposition model for high-energy nuclear collisions). The initial temperature at the center of the fireball is T0,center=1.05T_{0,\text{center}}=1.05 GeV at a starting time of τ0=0.01fm/c\tau_{0}=0.01\,\text{fm}/c which is the same for both types of hydrodynamic simulations. The fluid is evolved from an initial pressure ratio R=103R=10^{-3} with a constant specific shear viscosity η/𝒮=0.2\eta/\mathcal{S}=0.2; the runs stop once all the fluid cells are below Tsw=0.136T_{\text{sw}}=0.136 GeV, which happens after τf78\tau_{f}\sim 7-8 fm/cc.

Refer to caption
Figure 11: (Color online) The evolution of (a) \mathcal{E} (GeV/fm3), (b) uxu^{x}, (c) 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} and (d) ππ/(𝒫2)\sqrt{\pi_{\perp}{\cdot\,}\pi_{\perp}}/(\mathcal{P}_{\perp}\sqrt{2}) along the xx–axis (y=ηs= 0y{\,=\,}\eta_{s}{\,=\,}0), given by conformal anisotropic hydrodynamics (VAH, dashed color) and second-order viscous hydrodynamics (VH, transparent color), for the smooth (3+1)–dimensional TR{}_{\text{\sc R}}ENTo initial condition described in the text.

Figure 11 shows the evolution of \mathcal{E}, uxu^{x}, 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} and ππ\sqrt{\pi_{\perp}{\cdot\,}\pi_{\perp}} /  (𝒫2)(\mathcal{P}_{\perp}\sqrt{2}) along the xx–axis (y=ηs=0)y=\eta_{s}=0). Initially, the pressure anisotropy 𝒫L𝒫\mathcal{P}_{L}{-}\mathcal{P}_{\perp} in second-order viscous hydrodynamics is so large that the longitudinal pressure turns negative, especially near the grid’s edges. In comparison, the pressure ratio 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} in anisotropic hydrodynamics is, at early times, only moderately larger in the central fireball region but quite dramatically different near its transverse edge, staying positive everywhere. The resulting differences in early-time viscous heating create a disparity between the two models for the normalization of the energy density which persists to late times even after the 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} ratios have converged. On the other hand, the transverse flows uxu^{x} predicted by the two models are nearly identical even at very early times since they are primarily driven by the transverse pressure gradients x𝒫\partial_{x}\mathcal{P}_{\perp} which, in the smooth TR{}_{\text{\sc R}}ENTo profile, are smaller at early times than the longitudinal gradients 1/τ\sim 1/\tau. The transverse shear stress πμν\pi_{\perp}^{{\mu\nu}}, which tends to counteract the fluid’s transverse acceleration, is larger in viscous hydrodynamics than in anisotropic hydrodynamics, especially along the edges of the fireball.393939The transverse shear stress πμν\pi_{\perp}^{{\mu\nu}} is generally nonzero even for azimuthally symmetric flow profiles as long as they do not possess Gubser symmetry. Overall, however, the hydrodynamic variables are not substantially different along the transverse directions.

Refer to caption
Figure 12: (Color online) The evolution of (a) \mathcal{E} (GeV/fm3), (b) τuη\tau u^{\eta}, (c) 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} and (d) 2WzWz\sqrt{2W_{\perp z}{\,\cdot\,}W_{\perp z}} /𝒫L2+2𝒫2\sqrt{\mathcal{P}_{L}^{2}{+}2\mathcal{P}_{\perp}^{2}} along the ηs\eta_{s}–axis (x=y= 0x{\,=\,}y{\,=\,}0), computed with conformal anisotropic hydrodynamics (VAH, dashed color) and second-order viscous hydrodynamics (VH, transparent color), for the smooth (3+1)–d TR{}_{\text{\sc R}}ENTo initial condition described in the text. Note that in panel (d) the ηs\eta_{s}–axis is shifted transversely to (x,y)=(7.04,0)(x,y)=(7.04,0) fm.

The longitudinal profile, however, evolves very differently in anisotropic hydrodynamics (VAH) compared to second-order viscous hydrodynamics (VH). This is shown in Fig. 12 where we plot the evolution of the dimensionless longitudinal velocity τuη\tau u^{\eta} (as well as \mathcal{E} and 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp}) along the ηs\eta_{s}–axis (x=y=0x=y=0). Similar to Fig. 11, the shear stress (2) in viscous hydrodynamics quickly overpowers the equilibrium pressure, making 𝒫L\mathcal{P}_{L} negative initially. Along the longitudinal direction this causes a strong reversal of the longitudinal flow τuη\tau u^{\eta} at around |ηs|7|\eta_{s}|\sim 7. This unphysical feature (primarily driven by the strong longitudinal expansion rate at early times) results in a longitudinal contraction (in ηs\eta_{s}) of the fireball near its forward and backward edges at the beginning of the simulation. In anisotropic hydrodynamics, the longitudinal pressure remains always positive, allowing for a stronger longitudinal flow profile whose signs are consistent with the gradients of the rapidity distribution of the energy density (142) (and thus of the thermal pressure). This gives anisotropic hydrodynamics a clear advantage over second-order viscous hydrodynamics when simulating the longitudinal dynamics of a heavy-ion collision.

Refer to caption
Figure 13: (Color online) The evolution of the adaptive time step Δτn\Delta\tau_{n} (dashed color) and CFL bound ΔτCFL\Delta\tau_{\text{CFL}} (transparent color) in conformal anisotropic hydrodynamics (red) and second-order viscous hydrodynamics (blue) for the smooth (3+1)–d TR{}_{\text{\sc R}}ENTo initial condition.

In panel (d) of Fig. 12 we also plot the spacetime rapidity dependence of the inverse Reynolds number 2WzWz/(𝒫L2+2𝒫2)\sqrt{-2W_{\perp z}{\,\cdot\,}W_{\perp z}/(\mathcal{P}_{L}^{2}+2\mathcal{P}_{\perp}^{2})}. The longitudinal momentum diffusion current WzμW_{\perp z}^{\mu} is nonzero only in regions that have both longitudinal and transverse gradients. For this reason we shift the ηs\eta_{s}–axis transversely to (x,y)=(7.04,0)(x,y)=(7.04,0) fm, which corresponds to the mid-right region of the grid. As expected, the momentum diffusion current is weakest around mid-rapidity, ηs0\eta_{s}\sim 0, where the fireball profile is approximately longitudinally boost-invariant, and strongest along the sloping edges of the rapidity plateau (142). However, this longitudinal edge region is also the place where its inverse Reynolds number differs most strongly between anisotropic and standard viscous hydrodynamics. Overall, WzμW_{\perp z}^{\mu} only makes up a tiny fraction of the total shear stress (2) in anisotropic hydrodynamics. The situation may change for initial conditions whose longitudinal and transverse gradients are larger than the ones used in this test comparison.

Finally, we plot in Fig. 13 the adaptive time step for each of the two hydrodynamic simulations. One sees that in VH Δτn\Delta\tau_{n} stagnates until τ 0.02\tau{\,\sim\,}0.02 fm/cc (see footnote 29) while the one in VAH starts increasing immediately. This indicates that at early times viscous hydrodynamics generally has a faster evolution rate and therefore requires a smaller initial time step compared to anisotropic hydrodynamics. Nevertheless, both adaptive time steps remain below their CFL bound (initially dominated by the longitudinal velocity uηu^{\eta}) until τ 0.81\tau{\,\sim\,}0.8{-}1 fm/cc. Notice that the CFL bounds for VH and VAH become virtually identical since they have almost the same transverse velocity profile (see Fig. 11b).

4.4 Non-conformal Bjorken flow test

Refer to caption
Figure 14: (Color online) Non-conformal Bjorken evolution of (a) /0\mathcal{E}/\mathcal{E}_{0}, (b) the pressure ratio 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp}, (c) the pressure anisotropy 23(𝒫L𝒫)\frac{2}{3}(\mathcal{P}_{L}{-}\mathcal{P}_{\perp}), and (d) the bulk viscous pressure Π\Pi (the latter two normalized to the equilibrium pressure 𝒫eq\mathcal{P}_{\text{eq}}), given by anisotropic hydrodynamics (VAH, dashed red), quasiparticle viscous hydrodynamics (VH, dashed blue) and standard viscous hydrodynamics (VH2, dashed green), along with their semi-analytic solutions (transparent color). The lower gray curve in panel (a) is from an ideal hydrodynamic calculation (i.e. 𝒫L=𝒫=𝒫eq()\mathcal{P}_{L}=\mathcal{P}_{\perp}=\mathcal{P}_{\text{eq}}(\mathcal{E}) and η/𝒮=ζ/𝒮=0\eta/\mathcal{S}=\zeta/\mathcal{S}=0).

Now we turn to testing our non-conformal hydrodynamic simulation with the QCD equation of state from Fig. 1. First, we run (3+1)–d anisotropic hydrodynamics with Bjorken initial conditions and compare it to the semi-analytic solution of the equations [66]

τ\displaystyle\partial_{\tau}\mathcal{E} =+𝒫Lτ,\displaystyle=-\frac{\mathcal{E}+\mathcal{P}_{L}}{\tau}\,, (102a)
τ𝒫L\displaystyle\partial_{\tau}\mathcal{P}_{L} =𝒫eq𝒫¯τΠ𝒫L𝒫3τπ/2+ζ¯zLτ,\displaystyle=\frac{\mathcal{P}_{\text{eq}}{-}\bar{\mathcal{P}}}{\tau_{\Pi}}-\frac{\mathcal{P}_{L}{-}\mathcal{P}_{\perp}}{3\tau_{\pi}/2}+\frac{\bar{\zeta}_{z}^{L}}{\tau}\,, (102b)
τ𝒫\displaystyle\partial_{\tau}\mathcal{P}_{\perp} =𝒫eq𝒫¯τΠ+𝒫L𝒫3τπ+ζ¯zτ,\displaystyle=\frac{\mathcal{P}_{\text{eq}}{-}\bar{\mathcal{P}}}{\tau_{\Pi}}+\frac{\mathcal{P}_{L}{-}\mathcal{P}_{\perp}}{3\tau_{\pi}}+\frac{\bar{\zeta}_{z}^{\perp}}{\tau}\,, (102c)
τB\displaystyle\partial_{\tau}B =BeqBτΠ++𝒫Lmτdmd(2𝒫𝒫L4B);\displaystyle=\frac{B_{\text{eq}}{-}B}{\tau_{\Pi}}+\frac{\mathcal{E}{+}\mathcal{P}_{L}}{m\tau}\frac{dm}{d\mathcal{E}}(\mathcal{E}{-}2\mathcal{P}_{\perp}{-}\mathcal{P}_{L}{-}4B)\,; (102d)

here we use the shear and bulk relaxation times (31)–(32), with the viscosities given by Eqs. (28)–(29) (see Fig. 3). The non-conformal anisotropic transport coefficients ζ¯zL\bar{\zeta}_{z}^{L} and ζ¯z\bar{\zeta}_{z}^{\perp} are given by Eqs. (35a) and (36a), respectively. We start the simulation at τ0=0.05\tau_{0}=0.05 fm/cc with initial temperature T0=0.718T_{0}=0.718 GeV and initial pressure ratio R=0.3R=0.3, and evolve the system until τf80\tau_{f}\sim 80 fm/cc when the temperature falls below Tsw=0.136T_{\text{sw}}=0.136 GeV (we plot results only up to τ=20\tau=20 fm/cc). We also repeat this for quasiparticle viscous hydrodynamics (VH) and standard viscous hydrodynamics (VH2) (see footnote 35).404040VH and VH2 use the same equation of state 𝒫eq()\mathcal{P}_{\text{eq}}(\mathcal{E}) as VAH but different transport coefficients, see Appendix E and Ref. [66]. Ideally, we would have preferred using smaller values for τ0\tau_{0} and RR to better match the longitudinally free-streaming initial condition used in the conformal hydrodynamic simulations (see Secs. 4.14.3 and footnote 9). However, we found that at earlier times the VAH simulation has greater difficulty reconstructing the anisotropic variables (Λ\Lambda, α\alpha_{\perp}, αL\alpha_{L}), especially directly at initialization. This indicates that our anisotropic hydrodynamic model [66] which integrates the QCD equation of state consistently with quasiparticle kinetic transport coefficients, has its limitations.

Figures 14a,b show the evolution of the normalized energy density /0\mathcal{E}/\mathcal{E}_{0} and pressure ratio 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} from the three hydrodynamic simulations. We also disentangle the longitudinal and transverse pressures’ viscous components 𝒫L𝒫eq=23Δ𝒫+Π\mathcal{P}_{L}{-}\mathcal{P}_{\text{eq}}=\frac{2}{3}\Delta\mathcal{P}{+}\Pi and 𝒫𝒫eq=13Δ𝒫+Π\mathcal{P}_{\perp}{-}\mathcal{P}_{\text{eq}}=-\frac{1}{3}\Delta\mathcal{P}{+}\Pi into the pressure anisotropy Δ𝒫=𝒫L𝒫\Delta\mathcal{P}=\mathcal{P}_{L}{-}\mathcal{P}_{\perp} and bulk viscous pressure Π=13(𝒫L+2𝒫)𝒫eq\Pi=\frac{1}{3}(\mathcal{P}_{L}{+}2\mathcal{P}_{\perp})-\mathcal{P}_{\text{eq}}; their evolutions relative to 𝒫eq\mathcal{P}_{\text{eq}} are shown in Figs. 14c,d. Although our initial condition for the pressure ratio is somewhat ad hoc, it quickly reaches its minimum value at τ0.1\tau\sim 0.1 fm/c, allowing for the energy density to closely follow its free-streaming limit at the beginning of the simulation. The 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} ratio is typically larger in anisotropic hydrodynamics compared to the two viscous hydrodynamic models (until τ1\tau\sim 1 fm/cc). This is mainly due to differences in the higher-order transport coefficients associated with the pressure anisotropy (i.e. beyond η/𝒮\eta/\mathcal{S} and τπ\tau_{\pi}[62, 90, 63, 66]. Although in Bjorken flow the bulk viscous pressure Π\Pi is much smaller than the pressure anisotropy Δ𝒫\Delta\mathcal{P}, it evolves very differently in standard viscous hydrodynamics VH2 compared to quasiparticle viscous hydrodynamics VH and anisotropic hydrodynamics VAH. Since the dimensionless bulk relaxation time τΠT<1\tau_{\Pi}T<1 in standard viscous hydrodynamics VH2 (see the dashed curve in Fig. 4b), the bulk viscous pressure quickly reaches its Navier-Stokes solution ΠNS=ζ/τ\Pi_{\text{NS}}=-\zeta/\tau at τ2\tau\sim 2 fm/cc. In contrast, the bulk relaxation time used in anisotropic hydrodynamics VAH and quasiparticle viscous hydrodynamics VH is significantly larger (solid curve in Fig. 4b). As a consequence, it takes a much longer time for Π\Pi to relax to ΠNS\Pi_{\text{NS}} [79, 66].

One also sees that the simulations agree very well with their corresponding semi-analytic solutions (continuous lines in transparent color). Although we do not display them explicitly here, the numerical errors are small enough that we can unambiguously distinguish the different dynamics of the three hydrodynamic models in comparison studies discussed in the next subsections.

4.5 (3+1)–d non-conformal hydrodynamic models in central Pb+Pb collisions

4.5.1 Smooth TR{}_{\text{\sc R}}ENTo initial conditions

Next we run non-conformal hydrodynamics for a central Pb+Pb collision with smooth, azimuthally symmetric TR{}_{\text{\sc R}}ENTo initial conditions; we set the initial time to τ0=0.05\tau_{0}=0.05 fm/cc and the initial pressure ratio to R=0.3R=0.3. The initial energy density profile is almost identical to the one in Sec. 4.3, except the normalization 1/τ0\propto 1/\tau_{0} is five times smaller, making the initial temperature at the center of the fireball T0,center=0.718T_{0,\text{center}}=0.718 GeV. We evolve the system until, at τf 15 16\tau_{f}{\,\sim\,}15{\,-\,}16 fm/cc, all fluid cells are below Tsw=0.136T_{\text{sw}}=0.136 GeV.

Figure 15 shows, for the first Δτ=6\Delta\tau=6 fm/cc of the simulation, the evolution of \mathcal{E}, uxu^{x}, 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp}, ππ\sqrt{\pi_{\perp}\cdot\pi_{\perp}}/(𝒫2)(\mathcal{P}_{\perp}\sqrt{2}), 23(𝒫L𝒫)\frac{2}{3}(\mathcal{P}_{L}-\mathcal{P}_{\perp})/𝒫eq\mathcal{P}_{\text{eq}} and Π/𝒫eq\Pi/\mathcal{P}_{\text{eq}} given by the three hydrodynamic models along the xx–axis (y=ηs=0y=\eta_{s}=0). Here the fireball maintains higher temperatures for a longer duration than in conformal hydrodynamics since the QCD equilibrium pressure is much weaker than its Stefan–Boltzmann limit (26) (see Fig. 1a). But similar to Fig. 11a, the energy density in anisotropic hydrodynamics cools down at a slightly faster rate compared to viscous hydrodynamics. This is initially due to the moderately larger 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} ratio in the central fireball region, which closely follows the Bjorken evolution in Fig. 14b. The viscous corrections to the longitudinal pressure increase as we move towards the edges of the fireball, but 𝒫L\mathcal{P}_{L} still remains positive in anisotropic hydrodynamics. The longitudinal pressure in standard viscous hydrodynamics, however, is strongly negative at early times (and also, to a lesser extent, for the quasiparticle case), although it does not significantly alter the fluid’s transverse dynamics. The 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp} ratios start to overlap at later times, mainly driven by the pressure anisotropies’ convergence in Fig. 15e.

Refer to caption
Figure 15: (Color online) The evolution of (a) \mathcal{E} (GeV/fm3), (b) uxu^{x}, (c) 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp}, (d) ππ/(𝒫2)\sqrt{\pi_{\perp}{\cdot\,}\pi_{\perp}}/(\mathcal{P}_{\perp}\sqrt{2}), (e) 23(𝒫L𝒫)/𝒫eq\frac{2}{3}(\mathcal{P}_{L}{-}\mathcal{P}_{\perp})/\mathcal{P}_{\text{eq}} and (f) Π/𝒫eq\Pi/\mathcal{P}_{\text{eq}} along the xx–axis (y=ηs= 0y{\,=\,}\eta_{s}{\,=\,}0), given by non-conformal anisotropic hydrodynamics (VAH, dashed color), quasiparticle viscous hydrodynamics (VH, transparent color) and standard viscous hydrodynamics (VH2, dotted transparent color), for the smooth (3+1)–d TR{}_{\text{\sc R}}ENTo initial condition. The minimum energy density parameter is set to min=0.02\mathcal{E}_{\text{min}}=0.02 GeV/fm3.

Compared to Fig. 11b, the transverse flow uxu^{x} in non-conformal hydrodynamics is significantly weaker due to the softer QCD equation of state. The dip in the speed of sound at the quark-hadron phase transition (see Fig. 1b) also flattens the transverse velocity gradients near the edges of the fireball. This causes a significant decrease in the transverse shear stress relative to the conformal case in Fig. 11d. Among the hydrodynamic models, we see that VH2 produces the smallest uxu^{x} around the edges of the fireball since it has the largest negative bulk viscous pressure there. This is due to its ability to converge to the Navier-Stokes solution shortly after the simulation begins. In contrast, the bulk viscous pressure in VH hovers slightly above zero across the grid before falling down toward negative values at τ3\tau\sim 3 fm/cc. Even then, it hardly catches up to the bulk viscous pressure curves in VH2 since the differences between their bulk relaxation times grow as we move away from the center of the fireball. Overall, the bulk viscous pressure in VH is much weaker compared to VH2, which leads to a stronger transverse flow. Finally, the VAH simulation has a bulk viscous pressure that is qualitatively similar to VH (except at the edges of the grid) but slightly lags behind it. As a result, it generates the largest transverse flow out of the three simulations. The study here illustrates how the evolution of the bulk viscous pressure and transverse fluid velocity strongly depends on the choice of hydrodynamic model, specifically the temperature-dependent model for the bulk relaxation time τΠ\tau_{\Pi}. This is likely to have important implications for the phenomenological extraction of the bulk viscosity ζ/𝒮\zeta/\mathcal{S} [1, 2].

Next, we compare the fireballs’ longitudinal evolution along the ηs\eta_{s}–axis (x=y=0x=y=0) in Figure 16. When comparing the longitudinal evolution of the energy density and viscous pressures between the three hydrodynamic models, we note qualitatively similar features as already observed in Fig. 15 for their transverse evolution. While the transverse velocities shown in Fig. 15b varied between the models as a result of differences in the transverse pressures (mainly the bulk viscous pressures), here differences in their longitudinal pressures result in very different longitudinal flow profiles (Figure 16b). Most notably, the large negative 𝒫L\mathcal{P}_{L} in standard viscous hydrodynamics (VH2) initially causes τuη\tau u^{\eta} to reverse sign very sharply in the forward and backward rapidity regions. Only by imposing strong regulations on the VH2 simulation is it able to recover the correct direction of longitudinal flow at later times as the gradients relax. The same breakdown at early times can also be seen in quasiparticle viscous hydrodynamics (VH) although there the situation is not nearly as bad. With VAH we can maintain positive longitudinal pressures so that the fireball can expand without imploding in regions where the longitudinal gradients are large. This greatly reduces the amount of regulation needed for the viscous pressures.

Refer to caption
Figure 16: (Color online) The evolution of (a) \mathcal{E} (GeV/fm3), (b) τuη\tau u^{\eta}, (c) 𝒫L/𝒫\mathcal{P}_{L}/\mathcal{P}_{\perp}, (d) 2WzWz\sqrt{2W_{\perp z}{\,\cdot\,}W_{\perp z}} /𝒫L2+2𝒫2\sqrt{\mathcal{P}_{L}^{2}{+}2\mathcal{P}_{\perp}^{2}}, (e) 23(𝒫L𝒫)/𝒫eq\frac{2}{3}(\mathcal{P}_{L}{-}\mathcal{P}_{\perp})/\mathcal{P}_{\text{eq}}, and (f) Π/𝒫eq\Pi/\mathcal{P}_{\text{eq}} along the ηs\eta_{s}–axis (x=y=0x=y=0), given by non-conformal anisotropic hydrodynamics (VAH, dashed color), quasiparticle viscous hydrodynamics (VH, transparent color) and standard viscous hydrodynamics (VH2, dotted transparent color), for the smooth (3+1)–d TR{}_{\text{\sc R}}ENTo initial condition. Note that in panel (d) the ηs\eta_{s}–axis is shifted transversely to (x,y)=(7.04,0)(x,y)=(7.04,0) fm.

Finally, in Fig. 17 we plot the adaptive time step for each of the three hydrodynamic simulations. One sees that in anisotropic hydrodynamics (VAH) Δτn\Delta\tau_{n} increases steadily until hitting the CFL bound at τ1\tau\sim 1 fm/cc. Its behavior is similar in quasiparticle viscous hydrodynamics (VH). The two CFL bounds are slightly separated due to differences between their transverse velocity profiles (see Fig. 15b). On the other hand, the adaptive time step in standard viscous hydrodynamics (VH2) does not pick up until τ0.4\tau\sim 0.4 fm/cc (see footnote 29). Even then, it fluctuates up and down before reaching its CFL bound, which is higher than the other two bounds since its transverse flow is suppressed by a large bulk viscous pressure. Although the code runs faster per time step (see Table 1), standard viscous hydrodynamics takes considerably more time steps to evolve the fluid than the other two hydrodynamic models, suggesting that it has a harder time resolving the fireball evolution in the presence of large gradients.

Refer to caption
Figure 17: (Color online) The evolution of the adaptive time step Δτn\Delta\tau_{n} (dashed color) and CFL bound ΔτCFL\Delta\tau_{\text{CFL}} (transparent color) in non-conformal anisotropic hydrodynamics (VAH, red), quasiparticle viscous hydrodynamics (VH, blue) and standard viscous hydrodynamics (VH2, green), for the smooth (3+1)–d TR{}_{\text{\sc R}}ENTo initial condition.

Refer to caption

Figure 18: (Color online) The evolution of the QCD energy density profile (GeV/fm3) in the central transverse plane (ηs= 0)(\eta_{s}{\,=\,}0), given by non-conformal anisotropic hydrodynamics (VAH, left column), quasiparticle viscous hydrodynamics (VH, middle column) and standard viscous hydrodynamics (VH2, right column) for the fluctuating (3+1)–d TR{}_{\text{\sc R}}ENTo event described in the text. The white contour lines are ηs= 0\eta_{s}{\,=\,}0 slices at the shown time frames of a particlization hypersurface of constant energy density sw=0.116\mathcal{E}_{\text{sw}}=0.116 GeV/fm3.

4.5.2 Fluctuating TR{}_{\text{\sc R}}ENTo initial conditions

Finally we study the differences among the three non-conformal hydrodynamic models for a central Pb+Pb collision with a fluctuating initial energy density profile, using the same model parameters as in Sec. 4.5.1. In the TR{}_{\text{\sc R}}ENTo model used in this work, the energy density profile only has fluctuations in the transverse directions; the longitudinal profile is modeled with a finite rapidity plateau with smooth slopes at its longitudinal ends (see Appendix D).

Figure 18 shows the evolution of the fluctuating energy density profile in the transverse plane (ηs=0)(\eta_{s}=0), along with the transverse slice of a particlization hypersurface of constant energy density sw=0.116\mathcal{E}_{\text{sw}}=0.116 GeV/fm3, for the three hydrodynamic simulations. Since anisotropic hydrodynamics generally has a smaller shear stress than viscous hydrodynamics (especially at early times), the transverse fluctuations across its fireball show the least dissipation or smearing. As a result, it can convert the initial-state eccentricities into anisotropic flow slightly more efficiently. The pressure anisotropy 𝒫L𝒫\mathcal{P}_{L}{-}\mathcal{P}_{\perp} and bulk viscous pressure Π\Pi mainly influence the overall size of the fireball on the particlization hypersurface, affecting the final-state particle yields. The initially strong longitudinal expansion rapidly cools down the system and temporarily shrinks the fireball size at early times; a more positive longitudinal pressure helps cool the fireball even further. Compared to the two viscous hydrodynamic models, anisotropic hydrodynamics has a larger longitudinal pressure, which means its hypersurface has a slightly narrower waist. At later times, the transverse expansion overtakes the longitudinal expansion, increasing the fireball size. Although anisotropic hydrodynamics has the largest transverse flow, it also has the smallest bulk viscous pressure, enabling the fireball to cool faster and evaporate more quickly. This ultimately results in a smaller maximum fireball size. In contrast, standard viscous hydrodynamics has a much larger bulk viscous pressure. This extends the fireball’s lifetime by about 11 fm/c/c relative to the one in anisotropic hydrodynamics, allowing it to grow larger in size.

Refer to caption
Figure 19: (Color online) The τx\tau{-}x slice at y=ηs= 0y{\,=\,}\eta_{s}{\,=\,}0 of the residual shear and mean-field inverse Reynolds numbers in anisotropic hydrodynamics (VAH, top row) and the shear and bulk inverse Reynolds numbers in quasiparticle viscous hydrodynamics (VH, middle row) and standard viscous hydrodynamics (VH2, bottom row), for the fluctuating (3+1)–d TR{}_{\text{\sc R}}ENTo event. The white contour lines are τx\tau{-}x slices (y=ηs=0y=\eta_{s}=0) of the same particlization hypersurface as Fig. 18. Spacetime regions that are regulated according to Sec. 3.5 and App. E are circled in black.
Refer to caption
Figure 20: (Color online) The same as Fig. 19 but for the τηs\tau{-}\eta_{s} plane at x=y= 0x{\,=\,}y{\,=\,}0. The ηs\eta_{s}–axis in the upper left panel is shifted transversely to (x,y)=(7.04,0)(x,y)=(7.04,0) fm; the corresponding hypersurface slice (dashed white) is smaller than the one at x=y= 0x{\,=\,}y{\,=\,}0 (solid white).

In Figs. 19 and 20 we plot the inverse Reynolds numbers from the three hydrodynamic simulations in the τx\tau{-}x plane at y=ηs= 0y{\,=\,}\eta_{s}{\,=\,}0 and in the τηs\tau{-}\eta_{s} plane at x=y= 0x{\,=\,}y{\,=\,}0, respectively. Traditionally, the inverse Reynolds number measures the validity of the hydrodynamic expansion around local equilibrium [20, 5]. Since anisotropic hydrodynamics expands the energy-momentum tensor (1) around an anisotropic background (i.e. Taμν=uνuν+𝒫Lzμzν12𝒫ΞμνT^{\mu\nu}_{a}=\mathcal{E}u^{\nu}u^{\nu}+\mathcal{P}_{L}z^{\mu}z^{\nu}-\frac{1}{2}\mathcal{P}_{\perp}\Xi^{\mu\nu}), its validity is measured by the residual shear inverse Reynolds number [49]

ReπW1=\displaystyle\text{Re}^{-1}_{\pi W}= ππ2WzWz𝒫L2+2𝒫2,\displaystyle\,\sqrt{\frac{\pi_{\perp}{\cdot\,}\pi_{\perp}-2W_{\perp z}{\,\cdot\,}W_{\perp z}}{\mathcal{P}_{L}^{2}+2\mathcal{P}_{\perp}^{2}}}\,, (103a)

as opposed to the shear and bulk inverse Reynolds numbers in second-order viscous hydrodynamics [20, 5]:

Reπ1=ππ𝒫eq3,ReΠ1=|Π|𝒫eq,\text{Re}^{-1}_{\pi}=\,\frac{\sqrt{\pi\cdot\pi}}{\mathcal{P}_{\text{eq}}\sqrt{3}}\,,\qquad\text{Re}^{-1}_{\Pi}=\,\frac{|\Pi|}{\mathcal{P}_{\text{eq}}}\,, (104)

with ππ=πμνπμν\pi{\,\cdot\,}\pi=\pi_{\mu\nu}\pi^{\mu\nu}. One sees that the residual shear inverse Reynolds numbers stay much smaller than unity inside the fireball, indicating that a first-order expansion in O(ReπW1)O(\text{Re}^{-1}_{\pi W}) is sufficient to capture the residual shear corrections to the anisotropic hydrodynamic equations. While there is no need to regulate their strength here, the regulation scheme might be needed for collision events with sharper initial-state fluctuations.

We also plot the contribution of the nonequilibrium mean-field component δB\delta B to the bulk viscous pressure in anisotropic hydrodynamics:

ReδB1=|δB|𝒫eq.\text{Re}_{\delta B}^{-1}=\frac{|\delta B|}{\mathcal{P}_{\text{eq}}}\,. (105)

We find that δB\delta B is moderately large and positive near the edges of the fireball but small and negative inside the fireball. However, it is necessary to regulate the mean-field in the latter region via Eq. (80) to prevent unstable growth. Specifically, the instability seems to originate near the quark-hadron phase transition, where the driving force proportional to the kinetic trace anomaly (k)2𝒫(k)𝒫L(k)=2𝒫𝒫L4B\mathcal{E}^{(k)}-2\mathcal{P}_{\perp}^{(k)}-\mathcal{P}_{L}^{(k)}=\mathcal{E}-2\mathcal{P}_{\perp}-\mathcal{P}_{L}-4B is at its strongest while the bulk relaxation rate τΠ1\tau_{\Pi}^{-1} is near a minimum. After applying the regulation (80) for a brief period, δB\delta B evolves freely from that point onward. While this initial instability is unfortunate, the overall impact of the non-equilibrium mean-field component on the longitudinal and transverse pressures inside the fireball region is limited.

In second-order viscous hydrodynamics, the shear inverse Reynolds number is large for a short period of time Δτ1\Delta\tau\sim 1 fm/cc after the collision. The bulk inverse Reynolds number in standard viscous hydrodynamics is also very large near the peak of the bulk viscosity ζ/𝒮\zeta/\mathcal{S} during the same time frame. To prevent the code from crashing, our regulation scheme suppresses both the shear stress and bulk viscous pressure (see Appendix E). Although it is still technically possible to run the heavy-ion simulation with standard viscous hydrodynamics, we cannot escape the effects of regulation around the base of the particlization hypersurface. In contrast, the viscous pressures in quasiparticle viscous hydrodynamics often do not require regulation414141Even in the absence of regulation, the longitudinal pressure in quasiparticle viscous hydrodynamics can still be negative in some regions at early times (τ<0.6\tau<0.6 fm/cc) (e.g. see Figs. 15 and 16). since the bulk inverse Reynolds number is quite small.

We close this section with a comparison of the particlization hypersurfaces in the τx\tau{-}x and τηs\tau{-}\eta_{s} planes. In Fig. 19 one sees that standard viscous hydrodynamics has the largest hypersurface along the xx–direction since its large bulk viscous pressure generates the most viscous heating. It also has the longest lifetime τf17\tau_{f}\sim 17 fm/cc, for the same reason. In contrast, it has a pretty narrow waist along the ηs\eta_{s}–direction because its longitudinal flow profile uηu^{\eta} initially contracts the medium in the rapidity direction. In anisotropic hydrodynamics, the longitudinal flow begins transporting the medium away from the collision zone, following the direction of the longitudinal pressure gradients. As a result, its hypersurface has a wider waist than the one in standard viscous hydrodynamics (about Δηs0.5\Delta\eta_{s}\sim 0.5 larger for τ12\tau\sim 1-2 fm/cc) along the ηs\eta_{s}–direction.

5 Benchmarks

In this Section we benchmark the typical computational time needed to run (2+1)–d non-conformal hydrodynamic simulations of Pb+Pb collisions at LHC energies (sNN=2.76\sqrt{s_{\text{NN}}}=2.76 TeV) with fluctuating initial conditions and different values for the impact parameter bb and model parameters. We also benchmark the OpenMP–accelerated runtime of (3+1)–d non-conformal hydrodynamics for a central Pb+Pb collision (b=0b=0 fm) with the smooth TR{}_{\text{\sc R}}ENTo initial condition and model parameters from Sec. 4.5.1.

Mean time (s)(s) Max time (s)(s) Mean time per step (s)(s) Mean # steps
VAH 204 (136) 1840 (1630) 0.639 (0.410) 272
VH 75.0 (49.9) 561 (320) 0.183 (0.119) 350
VH2 89.7 (56.9) 463 (358) 0.175 (0.109) 461
Table 1: The mean runtime, maximum runtime, mean runtime per step and mean number of steps of the (2+1)–d non-conformal hydrodynamic simulations of Pb+Pb collisions on the fixed grid (auto–grid). (The number of time steps in the last column does not change for the auto–grid.) These statistics were generated by running a total of 1000010000 simulations (200200 test parameter samples times 5050 fluctuating TR{}_{\text{\sc R}}ENTo events) on single-core Intel Xeon E5-2680 v4 CPUs for each of the three hydrodynamic models.

5.1 (2+1)–d non-conformal hydrodynamics on a fixed grid

In this test, we generate 200200 random samples for the impact parameter bb and Bayesian model parameters [1, 2]. The impact parameter is sampled from a piecewise linear distribution,

P(b)={b/(2RA2)(0b2RA),0(b> 2RA),P(b)=\Bigg{\{}\begin{array}[]{ll}b/(2R_{A}^{2})&(0\leq b\leq 2R_{A}),\\ 0&(b{\,>\,}2R_{A}),\end{array} (106)

where we set RA=7R_{A}=7 fm for the lead nuclear radius. The model parameters

PB=[N,p,w,dmin,σk,Tsw,(η/𝒮)kink,Tη,alow,ahigh,(ζ/𝒮)max,Tζ,wζ,λζ]P_{B}=\left[N,p,w,d_{\text{min}},\sigma_{k},T_{\text{sw}},(\eta/\mathcal{S})_{\text{kink}},T_{\eta},a_{\text{low}},a_{\text{high}},(\zeta/\mathcal{S})_{\text{max}},T_{\zeta},w_{\zeta},\lambda_{\zeta}\right] (107)

are each sampled from a distribution that is uniform within the finite intervals used in the JETSCAPE SIMS Bayesian analysis of heavy-ion collisions [1, 2].424242The TR{}_{\text{\sc R}}ENTo initial condition model parameters (NN, pp, ww, dmind_{\text{min}}, σk\sigma_{k}), are defined in Appendix D [24] and the viscosity model parameters ((η/𝒮)kink(\eta/\mathcal{S})_{\text{kink}}, TηT_{\eta}, alowa_{\text{low}}, ahigha_{\text{high}}, (ζ/𝒮)max(\zeta/\mathcal{S})_{\text{max}}, TζT_{\zeta}, wζw_{\zeta}, λζ\lambda_{\zeta}) are discussed in Sec. 2.6.1. Finally, the switching temperature TswT_{\text{sw}} determines the particlization hypersurface of constant energy density sw=(Tsw)\mathcal{E}_{\text{sw}}=\mathcal{E}(T_{\text{sw}}).,434343We exclude the ratio between the dimensionless shear relaxation time and shear viscosity bπ=τπT/(η/𝒮)b_{\pi}=\tau_{\pi}T/(\eta/\mathcal{S}) as a continuous model parameter [1, 2] and instead allow it to vary between discrete hydrodynamic models. The rapidity plateau model parameters ηflat\eta_{\text{flat}} and ση\sigma_{\eta} (see Appendix D) are also not considered in this benchmark test.

Refer to caption
Figure 21: (Color online) The runtime distribution of the (2+1)–d non-conformal anisotropic hydrodynamic simulations on the fixed grid (blue) and auto-grid (red).

For each parameter sample {b,PB}\{b,P_{B}\} we run 5050 (2+1)–d non-conformal hydrodynamic simulations with fluctuating TR{}_{\text{\sc R}}ENTo initial conditions on a 3030 fm ×\times3030 fm transverse grid (ηs=0\eta_{s}=0).444444All other runtime parameters (e.g. the initial time τ0\tau_{0} and pressure ratio RR) are the same as the ones used in Sec. 4.5.,454545We found that 93%93\% of the anisotropic hydrodynamic simulations finish successfully (the second-order viscous hydrodynamic simulations had a success rate of 99+%99{+}\%). For certain model parameter sets and/or fluctuating initial conditions, the code fails to reconstruct the anisotropic variables in the cold dilute regions, either because the maximum number of Newton iterations was exhausted or the kinetic longitudinal pressure 𝒫L(k)\mathcal{P}_{L}^{(k)} turned negative. The resulting runtime statistics of the three hydrodynamic models are shown in Table 1. On average, it takes about 85s85s to run each (2+1)–d viscous hydrodynamic simulation on the fixed grid, compared to 200s200s for anisotropic hydrodynamics. The additional routine in Sec. 3.4 makes the runtime per step in anisotropic hydrodynamics about 3.5×3.5\times slower than in viscous hydrodynamics, but this is partially compensated by the fewer time steps required to finish each simulation.

Figure 21 shows the runtime distribution of the anisotropic hydrodynamic simulations on the fixed grid. One sees that some runs take longer than others, depending on the parameter values used. For example, a smaller impact parameter bb increases the participant nucleon multiplicity, which extends the fireball’s lifetime. The runtime is also very sensitive to the nucleon width parameter ww: halving the value for ww doubles the spatial resolution required464646In this work, we set the transverse lattice spacing to Δx=Δy=15w\Delta x=\Delta y=\frac{1}{5}w. to capture the spatial variations of the fluctuating TR{}_{\text{\sc R}}ENTo profile (see Appendix D) and therefore increases the runtime by roughly a factor of eight. As a result, the maximum runtime can be about an order of magnitude longer than the mean runtime (see Table 1).

5.2 (2+1)–d non-conformal hydrodynamics on an automated grid

The previous benchmark test was performed on a fixed transverse grid of 3030 fm ×\times3030 fm. Although this grid is large enough to fit all possible fireball sizes produced in Pb+Pb collisions, evolving peripheral collisions (b2RAb\lesssim 2R_{A}) on it is computationally inefficient. A large grid is also unnecessary for certain Bayesian model parameter combinations that further decrease the fireball size. In this section, we present a regression algorithm that predicts the fireball radius for a given set of parameters {b,PB}\{b,P_{B}\}. With this, we can automatically optimize the grid size to save computational time and memory.

Refer to caption
Figure 22: (Color online) A subset of the scattering matrix used to train the automated grid for (2+1)–d non-conformal anisotropic hydrodynamic simulations.

To train the regression model, we generate 10001000 training parameter samples arranged in the column vectors

𝑨=[𝒃,𝑷B],\bm{A}=\left[\bm{b},\bm{P}_{B}\right], (108)

to construct our feature matrix [91]. For each parameter sample, we run a single hydrodynamic simulation with a smooth, event-averaged TR{}_{\text{\sc R}}ENTo initial condition and compute the mean fireball radius 𝒓¯\bar{\bm{r}} (again arranged in vector); this will be our target variable474747We define the fireball radius rr as the maximum transverse radius of the particlization hypersurface in a given hydrodynamic simulation.

𝒚=𝒓¯.\bm{y}=\bm{\bar{r}}\,. (109)

A sample subset of the scattering matrix containing the five most important features is shown in Fig. 22. As expected, the fireball size strongly decreases with the impact parameter. There are also some moderately positive correlations between the fireball radius and initial condition parameters (e.g. increasing the nucleon width parameter ww increases the spread of the participant nucleons’ thickness function, see Appendix D, Eq. (139)). While the viscosity parameters have a much smaller effect on the fireball radius individually, the model’s fit to the train-validation data improves when we include all of them.

VAH VH VH2
δr¯RMSE\delta\bar{r}_{\text{RMSE}} 0.290.29 fm 0.190.19 fm 0.220.22 fm
α\alpha 0.0130.013 0.0090.009 0.0070.007
Table 2: The cross-validated root-mean-square error δr¯RMSE\delta\bar{r}_{\text{RMSE}} and regularization parameter α\alpha of the Lasso regression fit for each hydrodynamic model.

Next, we fit a cubic polynomial Lasso regression model with standardization feature scaling [91] to the training data (108)–(109). The regularization parameter α\alpha is chosen to minimize the root-mean-square error δr¯RMSE\delta\bar{r}_{\text{RMSE}}, averaged over a five-fold cross-validation [91]. The values for the cross-validated error and regularization parameter of the regression fit are listed in Table 2.

Finally, we rerun the simulations with fluctuating TR{}_{\text{\sc R}}ENTo initial conditions, using the 200 test parameter samples from earlier. For each parameter sample, we launch the regression model to predict the mean fireball radius r¯pred\bar{r}_{\text{pred}} and set the transverse grid lengths to

Lx=Ly=2(r¯pred+δr¯RMSE+),L_{x}=L_{y}=2\left(\bar{r}_{\text{pred}}+\delta\bar{r}_{\text{RMSE}}+\ell\right)\,, (110)

where the margin parameter \ell gives the fireball additional room to expand within the grid. Here we set =2.5\ell=2.5 fm not only for extra space but also to account for statistical fluctuations of the fireball radius in fluctuating TR{}_{\text{\sc R}}ENTo events, which were not considered in the training routine.484848With enough computing resources, the user has the option to run multiple fluctuating hydrodynamic simulations per parameter sample to produce a statistical distribution for the fireball radius rr, which can be characterized with a mean radius and standard deviation 𝒀=[𝒓¯,𝝈𝒓]\bm{Y}=[\bm{\bar{r}},\bm{\sigma_{r}}]. For more details, we refer the reader to https://github.com/mjmcnelis/fireball. We find that the automated grid has about a 99.6%99.6\% success rate of enclosing the fireball without touching its edges. Fig. 21 shows the anisotropic hydrodynamic runtime distribution on the automated grid. On average, the automated grid algorithm reduces the grid area by 3234%32-34\,\% relative to the fixed grid and provides the simulation with an additional 1.51.6×1.5-1.6\times speedup (see Table 1).

While (3+1)–d hydrodynamic simulations stand to benefit the most from an optimized grid volume, we would need to implement more realistic longitudinal initial conditions than the model used in this work before retraining the regression model to predict (in addition to its transverse radius) the fireball’s average elongation along the ηs\eta_{s}–direction. There may also be additional parameters that characterize such fluctuating (3+1)–d initial-state models, depending on the collision system of interest. Since it takes much more computational resources to produce such a training data set, we leave the (3+1)–d automated grid for future work.

Refer to caption
Figure 23: (Color online) The simulation runtime of (3+1)–d non-conformal anisotropic hydrodynamics (red dots), quasiparticle viscous hydrodynamics (blue dots) and standard viscous hydrodynamics (green dots) on an Intel Xeon E5-2680 v4 multi-core processor for the smooth TR{}_{\text{\sc R}}ENTo initial condition from Sec. 4.5.1. The ideal runtime is inversely proportional to the number of parallel CPU threads (same-colored continuous lines).

5.3 (3+1)–d non-conformal hydrodynamics with OpenMP acceleration

The code also includes the option to use OpenMP acceleration on a multi-core processor. This feature is especially useful for speeding up (3+1)–d hydrodynamic simulations, which run typically about two orders of magnitude longer than (2+1)–d simulations [20].494949For even faster runtimes, one can parallelize relativistic hydrodynamics on a graphics processing unit [20, 76]. The present version of VAH does not yet offer this option. Fig. 23 shows the runtime of the (3+1)–d non-conformal hydrodynamic simulations from Sec. 4.5.1. The anisotropic hydrodynamic simulation takes about two and a half hours to finish on a single-core CPU; the runtime of central Pb+Pb collisions can be about an order of magnitude longer or shorter than this depending on the values used for the nucleon-width parameter ww and rapidity plateau parameters ηflat\eta_{\text{flat}} and ση\sigma_{\eta} (see Appendix D). On an Intel Xeon E5-2680 multi-core processor, we can significantly reduce the simulation time to about nine minutes with 28 CPU threads. However, we do not achieve a perfect speedup because some parts of the routine are not easily parallelizable (e.g. the construction of the particlization hypersurface).

6 Summary and outlook

In this work we developed a (3+1)–dimensional anisotropic fluid dynamical simulation that evolves both the pre-equilibrium and viscous hydrodynamic stages of ultrarelativistic nuclear collisions. We validated the code’s performance by reproducing the semi-analytic solutions of conformal and non-conformal Bjorken flow and conformal Gubser flow on a Cartesian grid. Thanks to the adaptive time step algorithm derived in Sec. 3.6, we can accurately capture the early-time dynamics of Bjorken and Gubser flow while finishing the simulation within a reasonable number of iterations. We also compared anisotropic hydrodynamics to two second-order viscous hydrodynamic models in central Pb+Pb collisions. Apart from the apparent sensitivity of the bulk viscous pressure evolution to the bulk relaxation time, the three hydrodynamic models have similar transverse dynamics, at least in the mid-rapidity region ηs=0\eta_{s}=0. However, the fluid’s longitudinal evolution varies greatly near the longitudinal edges of the fireball. We showed that the longitudinal pressure in (3+1)–dimensional anisotropic hydrodynamics stays positive even in the presence of large gradients at very early times, unlike in viscous hydrodynamics. This causes the fluid to initially expand outward along the spacetime rapidity direction, as expected from the outward-pointing longitudinal gradients of the thermal pressure, reducing the risk of cavitation at the beginning of the simulation. With the new development presented here, we can for the first time model even the very early pre-equilibrium dynamics stage at τ0τhydro\tau_{0}{\,\ll\,}\tau_{\text{hydro}} with a QCD equation of state, as opposed to the conformal equation of state implicit in other pre-equilibrium models [92, 93, 50, 94, 95, 96, 44, 45].

In the near term, we plan to run VAH in the JETSCAPE framework with the Maximum a Posteriori (MAP) model parameters extracted in Refs. [1, 2].505050Several model parameters associated with the conformal free-streaming module would not be used in our code, which replaces the free-streaming stage by anisotropic hydrodynamics. Since the JETSCAPE SIMS hybrid model [1, 2] uses conformal free-streaming and standard viscous hydrodynamics to evolve the pre-equilibrium and hydrodynamic stages, we will replace these two modules with our code and study the changes to the hadronic observables, such as the transverse momentum spectra and pTp_{T}–differential anisotropic flows. A full analysis of all available data will need to wait for a Bayesian recalibration of the full evolution model using VAH as its hydrodynamic core.

Integration of VAH with the JETSCAPE framework also requires an updated version of the particlization module iS3D [74] with a new option that allows choosing the leading-order anisotropic distribution faf_{a} (plus residual shear corrections δf~\delta\tilde{f}) for the hadronic distribution in the Cooper–Frye formula [97, 74]. Once completed, this will allow us to investigate how different selections among a discrete set of hydrodynamic models affect both the theoretical description of heavy-ion experimental observables and the shear and bulk viscosity constraints inferred from such data–theory comparisons. In particular, it would be interesting to test the predictions from anisotropic hydrodynamics against a variety of initializations of second-order viscous hydrodynamics using different pre-hydrodynamic evolution models515151Anisotropic hydrodynamics (VAH) as presented here is not meant to be integrated with a pre-hydrodynamic module, especially one that uses a conformal approximation. This is mainly due to the technical difficulties in initializing the mean-field and anisotropic variables [66]. Available second-order viscous hydrodynamic codes can read in the energy-momentum tensor from a pre-hydrodynamic module and thus initialize the dynamical and inferred variables more easily, but this has not yet been implemented in the VAH code. (or no pre-hydrodynamic stage at all), such as the recent study done in Ref. [53].

The current code only evolves the fluid’s energy-momentum tensor components and ignores the effects of net-baryon density and diffusion. Second-order viscous hydrodynamic codes with nonzero nBn_{B} and VBμV_{B}^{\mu} have already been practically implemented [98, 83], and others that initialize viscous hydrodynamics with dynamical sources [10, 54, 55, 57, 58] are currently under development. The BEShydro code [57, 58], in particular, shares a common ancestry [49] and therefore a number of similar features with VAH. The integration of both codes into a single framework, in order to utilize the adaptive time step for capturing and resolving the dynamical production of energy and baryon sources at the onset of low-energy nuclear collisions, offers itself as an interesting project. On the theoretical side, the effects of non-zero chemical potentials for net charge, baryon number and strangeness have not yet been considered in our formulation of anisotropic hydrodynamics [66]. An upgrade of our code package that will include them is planned for the future.

7 Acknowledgements

M.M. would like to thank Seyed Sabok-Sayr from Rutgers University for assisting in the development of the automated grid algorithm during the Erdős Institute 2020 Data Science Boot Camp. Computational resources for the code validation, comparison and benchmark tests were provided by the Ohio Supercomputer Center under Project PAS0254 [99]. This work was supported by the National Science Foundation (NSF) within the framework of the JETSCAPE Collaboration under Award No. ACI-1550223. Additional partial support by the U.S. Department of Energy (DOE), Office of Science, Office for Nuclear Physics under Award No. DE-SC0004286 and within the framework of the BEST and JET Collaborations is also acknowledged.

Appendix A Gradient source terms

In this appendix, we list the gradients that appear in the source terms of the anisotropic hydrodynamic equations. The derivatives of the fluid velocity component uτu^{\tau} are

τuτ\displaystyle\partial_{\tau}u^{\tau} =vxτux+vyτuy+τ2vητuη+τvηuη,\displaystyle=v^{x}\partial_{\tau}u^{x}+v^{y}\partial_{\tau}u^{y}+\tau^{2}v^{\eta}\partial_{\tau}u^{\eta}+\tau v^{\eta}u^{\eta}\,, (111a)
xuτ\displaystyle\partial_{x}u^{\tau} =vxxux+vyxuy+τ2vηxuη,\displaystyle=v^{x}\partial_{x}u^{x}+v^{y}\partial_{x}u^{y}+\tau^{2}v^{\eta}\partial_{x}u^{\eta}\,, (111b)
yuτ\displaystyle\partial_{y}u^{\tau} =vxyux+vyyuy+τ2vηyuη,\displaystyle=v^{x}\partial_{y}u^{x}+v^{y}\partial_{y}u^{y}+\tau^{2}v^{\eta}\partial_{y}u^{\eta}\,, (111c)
ηuτ\displaystyle\partial_{\eta}u^{\tau} =vxηux+vyηuy+τ2vηηuη.\displaystyle=v^{x}\partial_{\eta}u^{x}+v^{y}\partial_{\eta}u^{y}+\tau^{2}v^{\eta}\partial_{\eta}u^{\eta}\,. (111d)

The derivatives of the longitudinal basis vector zμz^{\mu} are

τzτ\displaystyle\partial_{\tau}z^{\tau} =τ1+u2(τuηuη(uxτux+uyτuy)1+u2)+zττ,\displaystyle=\frac{\tau}{\sqrt{1{+}u_{\perp}^{2}}}\left(\partial_{\tau}u^{\eta}-\frac{u^{\eta}\left(u^{x}\partial_{\tau}u^{x}{+}u^{y}\partial_{\tau}u^{y}\right)}{1{+}u_{\perp}^{2}}\right)+\frac{z^{\tau}}{\tau}\,, (112a)
xzτ\displaystyle\partial_{x}z^{\tau} =τ1+u2(xuηuη(uxxux+uyxuy)1+u2),\displaystyle=\frac{\tau}{\sqrt{1{+}u_{\perp}^{2}}}\left(\partial_{x}u^{\eta}-\frac{u^{\eta}\left(u^{x}\partial_{x}u^{x}{+}u^{y}\partial_{x}u^{y}\right)}{1{+}u_{\perp}^{2}}\right)\,, (112b)
yzτ\displaystyle\partial_{y}z^{\tau} =τ1+u2(yuηuη(uxyux+uyyuy)1+u2),\displaystyle=\frac{\tau}{\sqrt{1{+}u_{\perp}^{2}}}\left(\partial_{y}u^{\eta}-\frac{u^{\eta}\left(u^{x}\partial_{y}u^{x}{+}u^{y}\partial_{y}u^{y}\right)}{1{+}u_{\perp}^{2}}\right)\,, (112c)
ηzτ\displaystyle\partial_{\eta}z^{\tau} =τ1+u2(ηuηuη(uxηux+uyηuy)1+u2),\displaystyle=\frac{\tau}{\sqrt{1{+}u_{\perp}^{2}}}\left(\partial_{\eta}u^{\eta}-\frac{u^{\eta}\left(u^{x}\partial_{\eta}u^{x}{+}u^{y}\partial_{\eta}u^{y}\right)}{1{+}u_{\perp}^{2}}\right)\,, (112d)
τzη\displaystyle\partial_{\tau}z^{\eta} =1τ1+u2(τuτuτ(uxτux+uyτuy)1+u2)zητ,\displaystyle=\frac{1}{\tau\sqrt{1{+}u_{\perp}^{2}}}\left(\partial_{\tau}u^{\tau}-\frac{u^{\tau}\left(u^{x}\partial_{\tau}u^{x}{+}u^{y}\partial_{\tau}u^{y}\right)}{1{+}u_{\perp}^{2}}\right)-\frac{z^{\eta}}{\tau}\,, (112e)
xzη\displaystyle\partial_{x}z^{\eta} =1τ1+u2(xuτuτ(uxxux+uyxuy)1+u2),\displaystyle=\frac{1}{\tau\sqrt{1{+}u_{\perp}^{2}}}\left(\partial_{x}u^{\tau}-\frac{u^{\tau}\left(u^{x}\partial_{x}u^{x}{+}u^{y}\partial_{x}u^{y}\right)}{1{+}u_{\perp}^{2}}\right)\,, (112f)
yzη\displaystyle\partial_{y}z^{\eta} =1τ1+u2(yuτuτ(uxyux+uyyuy)1+u2),\displaystyle=\frac{1}{\tau\sqrt{1{+}u_{\perp}^{2}}}\left(\partial_{y}u^{\tau}-\frac{u^{\tau}\left(u^{x}\partial_{y}u^{x}{+}u^{y}\partial_{y}u^{y}\right)}{1{+}u_{\perp}^{2}}\right)\,, (112g)
ηzη\displaystyle\partial_{\eta}z^{\eta} =1τ1+u2(ηuτuτ(uxηux+uyηuy)1+u2).\displaystyle=\frac{1}{\tau\sqrt{1{+}u_{\perp}^{2}}}\left(\partial_{\eta}u^{\tau}-\frac{u^{\tau}\left(u^{x}\partial_{\eta}u^{x}{+}u^{y}\partial_{\eta}u^{y}\right)}{1{+}u_{\perp}^{2}}\right)\,. (112h)

The divergence of the spatial fluid velocity vi=ui/uτv^{i}=u^{i}/u^{\tau} is

ivi=xuxvxxuτ+yuyvyyuτ+ηuηvηηuτuτ.\partial_{i}v^{i}=\frac{\partial_{x}u^{x}-v^{x}\partial_{x}u^{\tau}+\partial_{y}u^{y}-v^{y}\partial_{y}u^{\tau}+\partial_{\eta}u^{\eta}-v^{\eta}\partial_{\eta}u^{\tau}}{u^{\tau}}\,. (113)

The longitudinal expansion rate is

θL=zμDzuμ=zμzννuμzμzνΓνλμuλ=(zτ)2τuτ+zτzη(τ2τuηηuτ)+(τzη)2ηuη+τ(zη)2uτ.\begin{split}\theta_{L}&=z_{\mu}D_{z}u^{\mu}=-z_{\mu}z^{\nu}\partial_{\nu}u^{\mu}-z_{\mu}z^{\nu}\Gamma^{\mu}_{\nu\lambda}u^{\lambda}\\ &=-(z^{\tau})^{2}\partial_{\tau}u^{\tau}+z^{\tau}z^{\eta}(\tau^{2}\partial_{\tau}u^{\eta}-\partial_{\eta}u^{\tau})+(\tau z^{\eta})^{2}\partial_{\eta}u^{\eta}+\tau(z^{\eta})^{2}u^{\tau}\,.\end{split} (114)

The transverse expansion rate is

θ=μuμ=θθL,\theta_{\perp}=\nabla_{\perp\mu}u^{\mu}=\theta-\theta_{L}\,, (115)

where

θ=Dμuμ=μuμ+Γμνμuν=τuτ+xux+yuy+ηuη+uττ\begin{split}\theta&=D_{\mu}u^{\mu}=\partial_{\mu}u^{\mu}+\Gamma^{\mu}_{\mu\nu}u^{\nu}\\ &=\partial_{\tau}u^{\tau}+\partial_{x}u^{x}+\partial_{y}u^{y}+\partial_{\eta}u^{\eta}+\frac{u^{\tau}}{\tau}\end{split} (116)

is the scalar expansion rate.

The components of the fluid acceleration

aμ=Duμ=uννuμ+uνΓνλμuλa^{\mu}=Du^{\mu}=u^{\nu}\partial_{\nu}u^{\mu}+u^{\nu}\Gamma^{\mu}_{\nu\lambda}u^{\lambda} (117)

are

aτ\displaystyle a^{\tau} =uττuτ+uxxuτ+uyyuτ+uηηuτ+τ(uη)2,\displaystyle=u^{\tau}\partial_{\tau}u^{\tau}+u^{x}\partial_{x}u^{\tau}+u^{y}\partial_{y}u^{\tau}+u^{\eta}\partial_{\eta}u^{\tau}+\tau(u^{\eta})^{2}\,, (118a)
ax\displaystyle a^{x} =uττux+uxxux+uyyux+uηηux,\displaystyle=u^{\tau}\partial_{\tau}u^{x}+u^{x}\partial_{x}u^{x}+u^{y}\partial_{y}u^{x}+u^{\eta}\partial_{\eta}u^{x}\,, (118b)
ay\displaystyle a^{y} =uττuy+uxxuy+uyyuy+uηηuy,\displaystyle=u^{\tau}\partial_{\tau}u^{y}+u^{x}\partial_{x}u^{y}+u^{y}\partial_{y}u^{y}+u^{\eta}\partial_{\eta}u^{y}\,, (118c)
aη\displaystyle a^{\eta} =uττuη+uxxuη+uyyuη+uηηuη+2uτuητ.\displaystyle=u^{\tau}\partial_{\tau}u^{\eta}+u^{x}\partial_{x}u^{\eta}+u^{y}\partial_{y}u^{\eta}+u^{\eta}\partial_{\eta}u^{\eta}+\frac{2u^{\tau}u^{\eta}}{\tau}\,. (118d)

The components of the longitudinal vector’s comoving time derivative

z˙μ=Dzμ=uννzμ+uνΓνλμzλ\dot{z}^{\mu}=Dz^{\mu}=u^{\nu}\partial_{\nu}z^{\mu}+u^{\nu}\Gamma^{\mu}_{\nu\lambda}z^{\lambda} (119)

are

z˙τ\displaystyle\dot{z}^{\tau} =uττzτ+uxxzτ+uyyzτ+uηηzτ+τuηzη,\displaystyle=u^{\tau}\partial_{\tau}z^{\tau}+u^{x}\partial_{x}z^{\tau}+u^{y}\partial_{y}z^{\tau}+u^{\eta}\partial_{\eta}z^{\tau}+\tau u^{\eta}z^{\eta}\,, (120a)
z˙η\displaystyle\dot{z}^{\eta} =uττzη+uxxzη+uyyzη+uηηzη+uτzη+uηzττ.\displaystyle=u^{\tau}\partial_{\tau}z^{\eta}+u^{x}\partial_{x}z^{\eta}+u^{y}\partial_{y}z^{\eta}+u^{\eta}\partial_{\eta}z^{\eta}+\frac{u^{\tau}z^{\eta}{+}u^{\eta}z^{\tau}}{\tau}\,. (120b)

The components of the fluid velocity’s longitudinal derivative

Dzuμ=zννuμzνΓνλμuλD_{z}u^{\mu}=-z^{\nu}\partial_{\nu}u^{\mu}-z^{\nu}\Gamma^{\mu}_{\nu\lambda}u^{\lambda} (121)

are

Dzuτ\displaystyle D_{z}u^{\tau} =zττuτzηηuττuηzη,\displaystyle=-z^{\tau}\partial_{\tau}u^{\tau}-z^{\eta}\partial_{\eta}u^{\tau}-\tau u^{\eta}z^{\eta}\,, (122a)
Dzux\displaystyle D_{z}u^{x} =zττuxzηηux,\displaystyle=-z^{\tau}\partial_{\tau}u^{x}-z^{\eta}\partial_{\eta}u^{x}\,, (122b)
Dzuy\displaystyle D_{z}u^{y} =zττuyzηηuy,\displaystyle=-z^{\tau}\partial_{\tau}u^{y}-z^{\eta}\partial_{\eta}u^{y}\,, (122c)
Dzuη\displaystyle D_{z}u^{\eta} =zττuηzηηuηuτzη+uηzττ.\displaystyle=-z^{\tau}\partial_{\tau}u^{\eta}-z^{\eta}\partial_{\eta}u^{\eta}-\frac{u^{\tau}z^{\eta}{+}u^{\eta}z^{\tau}}{\tau}\,. (122d)

The transverse gradient of the fluid velocity projected along the longitudinal direction is

zνμuν=ΞαμzνDαuν,z_{\nu}\nabla_{\perp}^{\mu}u^{\nu}=\Xi^{\mu}_{\alpha}z_{\nu}D^{\alpha}u^{\nu}\,, (123)

where the components of

zνDαuν=gαβzνβuν+gαβzνΓβλνuλz_{\nu}D^{\alpha}u^{\nu}=g^{\alpha\beta}z_{\nu}\partial_{\beta}u^{\nu}+g^{\alpha\beta}z_{\nu}\Gamma^{\nu}_{\beta\lambda}u^{\lambda} (124)

are

zνDτuν\displaystyle z_{\nu}D^{\tau}u^{\nu} =zττuττ2zητuητuηzη,\displaystyle=z^{\tau}\partial_{\tau}u^{\tau}-\tau^{2}z^{\eta}\partial_{\tau}u^{\eta}-\tau u^{\eta}z^{\eta}\,, (125a)
zνDxuν\displaystyle z_{\nu}D^{x}u^{\nu} =zτxuτ+τ2zηxuη,\displaystyle=-z^{\tau}\partial_{x}u^{\tau}+\tau^{2}z^{\eta}\partial_{x}u^{\eta}\,, (125b)
zνDyuν\displaystyle z_{\nu}D^{y}u^{\nu} =zτyuτ+τ2zηyuη,\displaystyle=-z^{\tau}\partial_{y}u^{\tau}+\tau^{2}z^{\eta}\partial_{y}u^{\eta}\,, (125c)
zνDηuν\displaystyle z_{\nu}D^{\eta}u^{\nu} =1τ2(zτηuττ2zηηuη+τ(uηzτuτzη))\displaystyle=-\frac{1}{\tau^{2}}\left(z^{\tau}\partial_{\eta}u^{\tau}-\tau^{2}z^{\eta}\partial_{\eta}u^{\eta}+\tau\left(u^{\eta}z^{\tau}{-}u^{\tau}z^{\eta}\right)\right) (125d)

The transverse velocity-shear tensor is525252In the code we obtain σμν\sigma_{\perp}^{\mu\nu} by applying the projector Ξαβμν\Xi^{\mu\nu}_{\alpha\beta} onto D(αuβ)D^{(\alpha}u^{\beta)} directly, rather than simplifying the expression (126).

σμν=ΞαβμνD(αuβ),\sigma_{\perp}^{\mu\nu}=\Xi^{\mu\nu}_{\alpha\beta}D^{(\alpha}u^{\beta)}\,, (126)

where the components of

D(αuβ)=12(gαρρuβ+gβρρuα+gαρΓρλβuλ+gβρΓρλαuλ)D^{(\alpha}u^{\beta)}=\frac{1}{2}\left(g^{\alpha\rho}\partial_{\rho}u^{\beta}+g^{\beta\rho}\partial_{\rho}u^{\alpha}+g^{\alpha\rho}\Gamma^{\beta}_{\rho\lambda}u^{\lambda}+g^{\beta\rho}\Gamma^{\alpha}_{\rho\lambda}u^{\lambda}\right) (127)

are

D(τuτ)\displaystyle D^{(\tau}u^{\tau)} =τuτ,\displaystyle=\partial_{\tau}u^{\tau}\,, (128a)
D(τux)\displaystyle D^{(\tau}u^{x)} =12(τuxxuτ),\displaystyle=\frac{1}{2}\left(\partial_{\tau}u^{x}-\partial_{x}u^{\tau}\right)\,, (128b)
D(τuy)\displaystyle D^{(\tau}u^{y)} =12(τuyyuτ),\displaystyle=\frac{1}{2}\left(\partial_{\tau}u^{y}-\partial_{y}u^{\tau}\right)\,, (128c)
D(τuη)\displaystyle D^{(\tau}u^{\eta)} =12(τuηηuττ2),\displaystyle=\frac{1}{2}\left(\partial_{\tau}u^{\eta}-\frac{\partial_{\eta}u^{\tau}}{\tau^{2}}\right)\,, (128d)
D(xux)\displaystyle D^{(x}u^{x)} =xux,\displaystyle=-\partial_{x}u^{x}\,, (128e)
D(xuy)\displaystyle D^{(x}u^{y)} =12(xuy+yux),\displaystyle=-\frac{1}{2}\left(\partial_{x}u^{y}+\partial_{y}u^{x}\right)\,, (128f)
D(xuη)\displaystyle D^{(x}u^{\eta)} =12(xuη+ηuxτ2),\displaystyle=-\frac{1}{2}\left(\partial_{x}u^{\eta}+\frac{\partial_{\eta}u^{x}}{\tau^{2}}\right)\,, (128g)
D(yuy)\displaystyle D^{(y}u^{y)} =yuy,\displaystyle=-\partial_{y}u^{y}\,, (128h)
D(yuη)\displaystyle D^{(y}u^{\eta)} =12(yuη+ηuyτ2),\displaystyle=-\frac{1}{2}\left(\partial_{y}u^{\eta}+\frac{\partial_{\eta}u^{y}}{\tau^{2}}\right)\,, (128i)
D(ηuη)\displaystyle D^{(\eta}u^{\eta)} =1τ2(ηuη+uττ).\displaystyle=-\frac{1}{\tau^{2}}\left(\partial_{\eta}u^{\eta}+\frac{u^{\tau}}{\tau}\right)\,. (128j)

The transverse vorticity tensor is

ωμν=ΞαμΞβνD[αuβ]=D[μuν]uμaνuνaμ+zμ(Dzuν+zαDνuα)zν(Dzuμ+zαDμuα)2\begin{split}\omega_{\perp}^{\mu\nu}&=\Xi^{\mu}_{\alpha}\Xi^{\nu}_{\beta}D^{[\alpha}u^{\beta]}\\ &=D^{[\mu}u^{\nu]}-\frac{u^{\mu}a^{\nu}{-}u^{\nu}a^{\mu}{+}z^{\mu}(D_{z}u^{\nu}{+}z_{\alpha}D^{\nu}u^{\alpha}){-}z^{\nu}(D_{z}u^{\mu}{+}z_{\alpha}D^{\mu}u^{\alpha})}{2}\end{split} (129)

where the components of

D[αuβ]=12(gαρρuβgβρρuα+gαρΓρλβuλgβρΓρλαuλ)D^{[\alpha}u^{\beta]}=\frac{1}{2}\left(g^{\alpha\rho}\partial_{\rho}u^{\beta}-g^{\beta\rho}\partial_{\rho}u^{\alpha}+g^{\alpha\rho}\Gamma^{\beta}_{\rho\lambda}u^{\lambda}-g^{\beta\rho}\Gamma^{\alpha}_{\rho\lambda}u^{\lambda}\right) (130)

are

D[τuτ]\displaystyle D^{[\tau}u^{\tau]} =0,\displaystyle=0\,, (131a)
D[τux]\displaystyle D^{[\tau}u^{x]} =12(τux+xuτ),\displaystyle=\frac{1}{2}\left(\partial_{\tau}u^{x}+\partial_{x}u^{\tau}\right)\,, (131b)
D[τuy]\displaystyle D^{[\tau}u^{y]} =12(τuy+yuτ),\displaystyle=\frac{1}{2}\left(\partial_{\tau}u^{y}+\partial_{y}u^{\tau}\right)\,, (131c)
D[τuη]\displaystyle D^{[\tau}u^{\eta]} =12(τuη+ηuττ2)+uητ,\displaystyle=\frac{1}{2}\left(\partial_{\tau}u^{\eta}+\frac{\partial_{\eta}u^{\tau}}{\tau^{2}}\right)+\frac{u^{\eta}}{\tau}\,, (131d)
D[xux]\displaystyle D^{[x}u^{x]} =0,\displaystyle=0\,, (131e)
D[xuy]\displaystyle D^{[x}u^{y]} =12(xuyyux),\displaystyle=-\frac{1}{2}\left(\partial_{x}u^{y}-\partial_{y}u^{x}\right)\,, (131f)
D[xuη]\displaystyle D^{[x}u^{\eta]} =12(xuηηuxτ2),\displaystyle=-\frac{1}{2}\left(\partial_{x}u^{\eta}-\frac{\partial_{\eta}u^{x}}{\tau^{2}}\right)\,, (131g)
D[yuy]\displaystyle D^{[y}u^{y]} =0,\displaystyle=0\,, (131h)
D[yuη]\displaystyle D^{[y}u^{\eta]} =12(yuη+ηuyτ2),\displaystyle=-\frac{1}{2}\left(\partial_{y}u^{\eta}+\frac{\partial_{\eta}u^{y}}{\tau^{2}}\right)\,, (131i)
D[ηuη]\displaystyle D^{[\eta}u^{\eta]} =0.\displaystyle=0\,. (131j)

Appendix B Geometric source terms

Here we list the components of the geometric source terms GWμG_{W}^{\mu} and GπμνG_{\pi}^{\mu\nu} (Eq. (23)) that appear in the relaxation equations (20) for WzμW_{\perp z}^{\mu} and πμν\pi_{\perp}^{{\mu\nu}}, respectively:

GWτ\displaystyle G_{W}^{\tau} =τuηWzη,\displaystyle=\tau u^{\eta}W_{\perp z}^{\eta}\,, (132a)
GWx\displaystyle G_{W}^{x} =0,\displaystyle=0\,, (132b)
GWy\displaystyle G_{W}^{y} =0,\displaystyle=0\,, (132c)
GWη\displaystyle G_{W}^{\eta} =uτWzη+uηWzττ.\displaystyle=\frac{u^{\tau}W_{\perp z}^{\eta}+u^{\eta}W_{\perp z}^{\tau}}{\tau}\,. (132d)
Gπττ\displaystyle G_{\pi}^{\tau\tau} =2τuηπτη,\displaystyle=2\tau u^{\eta}\pi_{\perp}^{\tau\eta}\,, (133a)
Gπτx\displaystyle G_{\pi}^{\tau x} =τuηπxη,\displaystyle=\tau u^{\eta}\pi_{\perp}^{x\eta}\,, (133b)
Gπτy\displaystyle G_{\pi}^{\tau y} =τuηπyη,\displaystyle=\tau u^{\eta}\pi_{\perp}^{y\eta}\,, (133c)
Gπτη\displaystyle G_{\pi}^{\tau\eta} =τuηπηη+uτπτη+uηπτττ,\displaystyle=\tau u^{\eta}\pi_{\perp}^{\eta\eta}+\frac{u^{\tau}\pi_{\perp}^{\tau\eta}{+}u^{\eta}\pi_{\perp}^{\tau\tau}}{\tau}\,, (133d)
Gπxx\displaystyle G_{\pi}^{xx} =0,\displaystyle=0\,, (133e)
Gπxy\displaystyle G_{\pi}^{xy} =0,\displaystyle=0\,, (133f)
Gπxη\displaystyle G_{\pi}^{x\eta} =uτπxη+uηπτxτ,\displaystyle=\frac{u^{\tau}\pi_{\perp}^{x\eta}+u^{\eta}\pi_{\perp}^{\tau x}}{\tau}\,, (133g)
Gπyy\displaystyle G_{\pi}^{yy} =0,\displaystyle=0\,, (133h)
Gπyη\displaystyle G_{\pi}^{y\eta} =uτπyη+uηπτyτ,\displaystyle=\frac{u^{\tau}\pi_{\perp}^{y\eta}+u^{\eta}\pi_{\perp}^{\tau y}}{\tau}\,, (133i)
Gπηη\displaystyle G_{\pi}^{\eta\eta} =2(uτπηη+uηπτη)τ.\displaystyle=\frac{2\left(u^{\tau}\pi_{\perp}^{\eta\eta}+u^{\eta}\pi_{\perp}^{\tau\eta}\right)}{\tau}\,. (133j)

Appendix C Conformal anisotropic transport coefficients

In the conformal limit m=B=0m=B=0, the anisotropic transport coefficients (35) – (38) only depend the functions nrqs\mathcal{I}_{nrqs}, which reduce to

nrqs=g(n+s+1)!αLr+1Λn+s+2nrq4π2(2q)!!,\mathcal{I}_{nrqs}=\frac{g(n{+}s{+}1)!\,\alpha_{L}^{r+1}\Lambda^{n+s+2}\mathcal{R}_{nrq}}{4\pi^{2}(2q)!!}\,, (134)

where gg, αL\alpha_{L} and Λ\Lambda are given by Eqs. (27), (72) and (73), respectively. We list the functions nrq\mathcal{R}_{nrq} used in this work [66]:

200\displaystyle\mathcal{R}_{200} =αL(1+(1+ξL)tL),\displaystyle=\alpha_{L}\big{(}1+(1+\xi_{L})t_{L}\big{)}\,, (135a)
220\displaystyle\mathcal{R}_{220} =1+(1+ξL)tLξLαL,\displaystyle=\frac{-1+(1+\xi_{L})t_{L}}{\xi_{L}\alpha_{L}}\,, (135b)
201\displaystyle\mathcal{R}_{201} =1+(ξL1)tLξLαL,\displaystyle=\frac{1+(\xi_{L}-1)t_{L}}{\xi_{L}\alpha_{L}}\,, (135c)
240\displaystyle\mathcal{R}_{240} =3+2ξL3(1+ξL)tL)ξL2αL3,\displaystyle=\frac{3+2\xi_{L}-3(1+\xi_{L})t_{L})}{\xi_{L}^{2}\alpha_{L}^{3}}\,, (135d)
202\displaystyle\mathcal{R}_{202} =3+ξL+(1+ξL)(ξL3)tLξL2(1+ξL)αL3,\displaystyle=\frac{3+\xi_{L}+(1+\xi_{L})(\xi_{L}-3)t_{L}}{\xi_{L}^{2}(1+\xi_{L})\alpha_{L}^{3}}\,, (135e)
221\displaystyle\mathcal{R}_{221} =3+(3+ξL)tL)ξL2αL3,\displaystyle=\frac{-3+(3+\xi_{L})t_{L})}{\xi_{L}^{2}\alpha_{L}^{3}}\,, (135f)
441\displaystyle\mathcal{R}_{441} =15+13ξL+3(1+ξL)(5+ξL)tL4ξL3αL3,\displaystyle=\frac{-15+13\xi_{L}+3(1+\xi_{L})(5+\xi_{L})t_{L}}{4\xi_{L}^{3}\alpha_{L}^{3}}\,, (135g)
402\displaystyle\mathcal{R}_{402} =3(ξL1)+(ξL(3ξL2)+3)tL4ξL2αL,\displaystyle=\frac{3(\xi_{L}-1)+(\xi_{L}(3\xi_{L}-2)+3)t_{L}}{4\xi_{L}^{2}\alpha_{L}}\,, (135h)
421\displaystyle\mathcal{R}_{421} =3+ξL+(1+ξL)(ξL3)tL4ξL2αL,\displaystyle=\frac{3+\xi_{L}+(1+\xi_{L})(\xi_{L}-3)t_{L}}{4\xi_{L}^{2}\alpha_{L}}\,, (135i)
422\displaystyle\mathcal{R}_{422} =15+ξL+(ξL(ξL6)15)tL4ξL3αL3,\displaystyle=\frac{15+\xi_{L}+(\xi_{L}(\xi_{L}-6)-15)t_{L}}{4\xi_{L}^{3}\alpha_{L}^{3}}\,, (135j)
403\displaystyle\mathcal{R}_{403} =(ξL3)(5+3ξL)+3(1+ξL)(ξL(ξL2)+5)tL4ξL3(1+ξL)αL3,\displaystyle=\frac{(\xi_{L}-3)(5+3\xi_{L})+3(1+\xi_{L})(\xi_{L}(\xi_{L}-2)+5)t_{L}}{4\xi_{L}^{3}(1+\xi_{L})\alpha_{L}^{3}}\,, (135k)

where ξL=αL21\xi_{L}=\alpha_{L}^{-2}-1 and tL=arctanξL/ξLt_{L}=\mathrm{arctan}\sqrt{\xi_{L}}/\sqrt{\xi_{L}}.

Appendix D TR{}_{\text{\sc R}}ENTo energy deposition model

In the TR{}_{\text{\sc R}}ENTo model, the transverse energy deposition (GeV/fm2) of a single fluctuating nuclear collision event in the mid-rapidity region is [24]

dETdxdydηs|ηs=0=N×TR(x,y),\frac{dE_{T}}{dxdyd\eta_{s}}_{|{\eta_{s}=0}}=N\times T_{R}(x,y)\,, (136)

where NN is the normalization parameter and

TR(x,y)=(TAp(x,y)+TBp(x,y)2)1/pT_{R}(x,y)=\left(\frac{T^{p}_{A}(x,y)+T^{p}_{B}(x,y)}{2}\right)^{1/p} (137)

is the reduced nuclear thickness function, with pp being the geometric parameter. The nuclear thickness function of nucleus AA (BB) is535353In this work, we only consider Pb+Pb collisions (A=B=208A=B=208) at LHC energies sNN=2.76\sqrt{s_{\text{NN}}}=2.76 TeV; the inelastic nucleon–nucleon cross section is set to σNN=6.4\sigma_{\text{NN}}=6.4 fm-2.

TA,B(x,y)=n=1Npart,A,BγnTp(xxn,yyn),T_{A,B}(x,y)=\sum_{n=1}^{N_{\text{part},A,B}}\gamma_{n}\,T_{p}(x-x_{n},y-y_{n})\,, (138)

where Npart,A,BN_{\text{part},A,B} are the number of participant nucleons from nucleus AA (BB); the nucleon positions are sampled from a Woods–Saxon distribution under the constraint that each nucleon–nucleon pair in nucleus AA (BB) maintains a minimum separation dmind_{\text{min}} [24]. The participant nucleon’s thickness function is centered around its sampled transverse position (xnx_{n}, yny_{n}):

Tp(xxn,yyn)=12πw2×exp[(xxn)2+(yyn)22w2],T_{p}(x-x_{n},y-y_{n})=\frac{1}{2\pi w^{2}}\times\exp\left[-\frac{(x{-}x_{n})^{2}+(y{-}y_{n})^{2}}{2w^{2}}\right]\,, (139)

where ww is the nucleon width parameter.545454This parameter does not control the nucleon’s charge radius but rather the spread of thermal energy it deposits in the collision zone along the transverse directions. Furthermore, the multiplicity factor γn\gamma_{n} of each participant nucleon is sampled from the gamma distribution

P(γ)=kkγk1exp[kγ]Γ[k],P(\gamma)=\frac{k^{k}\gamma^{k-1}\exp\left[-k\gamma\right]}{\Gamma[k]}\,, (140)

where k=σk2k=\sigma_{k}^{-2} and σk\sigma_{k} is the standard deviation.

For a very brief period τ0\tau_{0} after the collision, we assume the system is longitudinally free-streaming and static in Milne coordinates (i.e. 𝒫L/𝒫eq0\mathcal{P}_{L}/\mathcal{P}_{\text{eq}}\sim 0 and 𝒖𝟎\bm{u}\sim\bm{0}) so that (x)1/τ0\mathcal{E}(x)\propto 1/\tau_{0}. Therefore, we initialize the energy density profile of the hydrodynamic simulation at the starting time τ0\tau_{0} as

(τ0,x,y,ηs)=1τ0×dETdxdydηs|ηs=0×fL(ηs).\mathcal{E}(\tau_{0},x,y,\eta_{s})=\frac{1}{\tau_{0}}\times\frac{dE_{T}}{dxdyd\eta_{s}}_{|\eta_{s}=0}\times f_{L}(\eta_{s})\,. (141)

Here we also extend the transverse energy density profile along the spacetime rapidity direction with a smooth plateau distribution (unitless) [76]:

fL(ηs)=exp[(|ηs|12ηflat)2Θ(|ηs|12ηflat)2ση2],f_{L}(\eta_{s})=\exp\left[-\frac{\big{(}|\eta_{s}|-\frac{1}{2}\eta_{\text{flat}}\big{)}^{2}\,\Theta\big{(}|\eta_{s}|-\frac{1}{2}\eta_{\text{flat}}\big{)}}{2\sigma_{\eta}^{2}}\right]\,, (142)

where ηflat\eta_{\text{flat}} is the plateau length, ση\sigma_{\eta} is the standard deviation of the half-Gaussian tails and Θ\Theta is the Heaviside step function. The initial condition parameter values used in this work are N=14.19N=14.19 GeV, p=0.06p=0.06, w=1.11w=1.11 fm, dmin=1.45d_{\text{min}}=1.45 fm, σk=1.03\sigma_{k}=1.03, ηflat=4.0\eta_{\text{flat}}=4.0 and ση=1.8\sigma_{\eta}=1.8 [76, 1, 2].

We set the lattice spacings to Δx=Δy=15w\Delta x=\Delta y=\frac{1}{5}w and Δηs=15ση\Delta\eta_{s}=\frac{1}{5}\sigma_{\eta} to resolve the fluctuating energy density profile (141) (or event-averaged profile). For the grid size, we set the longitudinal length to Lη=ηflat+10σηL_{\eta}=\eta_{\text{flat}}+10\sigma_{\eta} to fit the rapidity plateau (142). The transverse lengths LxL_{x} and LyL_{y} are automatically configured by the algorithm described in Sec. 5.2.

Appendix E Numerical implementation of second-order viscous hydrodynamics

In this appendix, we summarize how second-order viscous hydrodynamics is implemented the code. We evolve viscous hydrodynamics with the same numerical algorithm discussed in Sec. 3 except the energy-momentum tensor (1) is decomposed as

Tμν=uμuν(𝒫eq+Π)Δμν+πμν,T^{\mu\nu}=\mathcal{E}u^{\mu}u^{\nu}-(\mathcal{P}_{\text{eq}}{+}\Pi)\Delta^{\mu\nu}+\pi^{\mu\nu}\,, (143)

where πμν=ΔαβμνTαβ\pi^{\mu\nu}=\Delta^{\mu\nu}_{\alpha\beta}T^{\alpha\beta} is the shear stress tensor and Π=13ΔμνTμν𝒫eq\Pi=-\frac{1}{3}\Delta_{\mu\nu}T^{\mu\nu}-\mathcal{P}_{\text{eq}} is the bulk viscous pressure. We also define the spatial projector Δμν=gμνuμuν\Delta^{\mu\nu}=g^{\mu\nu}-u^{\mu}u^{\nu} and traceless double spatial projector Δαβμν=12(ΔαμΔβν+ΔβνΔαμ23ΔμνΔαβ)\Delta^{\mu\nu}_{\alpha\beta}=\frac{1}{2}(\Delta^{\mu}_{\alpha}\Delta^{\nu}_{\beta}+\Delta^{\nu}_{\beta}\Delta^{\mu}_{\alpha}-\frac{2}{3}\Delta^{\mu\nu}\Delta_{\alpha\beta}). The corresponding dynamical variables

𝒒=(Tτμ,πμν,Π)\bm{q}=(T^{\tau\mu},\pi^{\mu\nu},\Pi) (144)

are propagated along with \mathcal{E} and 𝒖\bm{u} (the mean-field BB and anisotropic variables (Λ\Lambda, α\alpha_{\perp}, αL\alpha_{L}) are not propagated). Although πμν\pi^{\mu\nu} has only five independent components, we propagate all ten components in the simulation [20].555555For longitudinally boost-invariant systems, we do not propagate the shear stress components πτη\pi^{\tau\eta}, πxη\pi^{x\eta} and πyη\pi^{y\eta}.

E.1 Hydrodynamic equations

Here we list the evolution equations for the dynamical variables (144) (we refer the reader to Refs. [20, 18] for details on their derivation):

τTττ+i(viTττ)=\displaystyle\partial_{\tau}T^{\tau\tau}+\partial_{i}(v^{i}T^{\tau\tau})= Tττ+τ2Tηητ+(πττ𝒫eqΠ)ivi\displaystyle-\frac{T^{\tau\tau}{+}\tau^{2}T^{\eta\eta}}{\tau}+(\pi^{\tau\tau}{-}\mathcal{P}_{\text{eq}}{-}\Pi)\partial_{i}v^{i} (145a)
+vii(πττ𝒫eqΠ)iπτi,\displaystyle+v^{i}\partial_{i}(\pi^{\tau\tau}{-}\mathcal{P}_{\text{eq}}{-}\Pi)-\partial_{i}\pi^{\tau i}\,,
τTτx+i(viTτx)=\displaystyle\partial_{\tau}T^{\tau x}+\partial_{i}(v^{i}T^{\tau x})= Tτxτx(𝒫eq+Π)+πτxivi+viiπτx\displaystyle-\frac{T^{\tau x}}{\tau}-\partial_{x}(\mathcal{P}_{\text{eq}}{+}\Pi)+\pi^{\tau x}\partial_{i}v^{i}+v^{i}\partial_{i}\pi^{\tau x} (145b)
iπxi,\displaystyle-\partial_{i}\pi^{xi}\,,
τTτy+i(viTτy)=\displaystyle\partial_{\tau}T^{\tau y}+\partial_{i}(v^{i}T^{\tau y})= Tτyτy(𝒫eq+Π)+πτyivi+viiπτy\displaystyle-\frac{T^{\tau y}}{\tau}-\partial_{y}(\mathcal{P}_{\text{eq}}{+}\Pi)+\pi^{\tau y}\partial_{i}v^{i}+v^{i}\partial_{i}\pi^{\tau y} (145c)
iπyi,\displaystyle-\partial_{i}\pi^{yi}\,,
τTτη+i(viTτη)=\displaystyle\partial_{\tau}T^{\tau\eta}+\partial_{i}(v^{i}T^{\tau\eta})= 3Tτητη(𝒫eq+Π)τ2+πτηivi+viiπτη\displaystyle-\frac{3T^{\tau\eta}}{\tau}-\frac{\partial_{\eta}(\mathcal{P}_{\text{eq}}{+}\Pi)}{\tau^{2}}+\pi^{\tau\eta}\partial_{i}v^{i}+v^{i}\partial_{i}\pi^{\tau\eta} (145d)
iπηi,\displaystyle-\partial_{i}\pi^{\eta i}\,,
τπμν+i(viπμν)=\displaystyle\partial_{\tau}\pi^{\mu\nu}+\partial_{i}(v^{i}\pi^{\mu\nu})= πμνivi+1uτ[πμντπ+πμν𝒫πμν𝒢πμν],\displaystyle\,\pi^{\mu\nu}\partial_{i}v^{i}+\frac{1}{u^{\tau}}\left[-\frac{\pi^{\mu\nu}}{\tau_{\pi}}+\mathcal{I}_{\pi^{\prime}}^{\mu\nu}-\mathcal{P}_{\pi^{\prime}}^{\mu\nu}-\mathcal{G}_{\pi^{\prime}}^{\mu\nu}\right]\,, (145e)
τΠ+i(viΠ)=\displaystyle\partial_{\tau}\Pi+\partial_{i}(v^{i}\Pi)= Πivi+1uτ[ΠτΠ+Π],\displaystyle\,\Pi\partial_{i}v^{i}+\frac{1}{u^{\tau}}\left[-\frac{\Pi}{\tau_{\Pi}}+\mathcal{I}_{\Pi}\right]\,, (145f)

where Tηη=(+𝒫eq+Π)(uη)2+(𝒫eq+Π)/τ2+πηηT^{\eta\eta}=(\mathcal{E}{+}\mathcal{P}_{\text{eq}}{+}\Pi)(u^{\eta})^{2}+(\mathcal{P}_{\text{eq}}{+}\Pi)/\tau^{2}+\pi^{\eta\eta},

πμν=\displaystyle\mathcal{I}^{\mu\nu}_{\pi^{\prime}}=  2βπσμν+Δαβμν(2πλ(αωλβ)τ¯πππλ(ασλβ))δ¯πππμνθ\displaystyle\,2\beta_{\pi}\sigma^{\mu\nu}+\Delta^{\mu\nu}_{\alpha\beta}\big{(}2\pi^{\lambda(\alpha}\omega^{\beta)}_{\,\,\,\lambda}-\bar{\tau}_{\pi\pi}\pi^{\lambda(\alpha}\sigma^{\beta)}_{\,\,\,\lambda}\big{)}-\bar{\delta}_{\pi\pi}\pi^{\mu\nu}\theta (146a)
+λ¯πΠΠσμν,\displaystyle+\bar{\lambda}_{\pi\Pi}\Pi\sigma^{\mu\nu}\,,
Π=\displaystyle\mathcal{I}_{\Pi}= βΠθδ¯ΠΠΠθ+λ¯Πππμνσμν,\displaystyle\,-\beta_{\Pi}\theta-\bar{\delta}_{\Pi\Pi}\Pi\theta+\bar{\lambda}_{\Pi\pi}\pi^{\mu\nu}\sigma_{\mu\nu}\,, (146b)

are the gradient source terms for πμν\pi^{\mu\nu} and Π\Pi and

𝒫πμν=\displaystyle\mathcal{P}^{\mu\nu}_{\pi^{\prime}}= (πμαuν+πναuμ)aα,\displaystyle\,\left(\pi^{\mu\alpha}u^{\nu}+\pi^{\nu\alpha}u^{\mu}\right)a_{\alpha}\,, (147a)
𝒢πμν=\displaystyle\mathcal{G}^{\mu\nu}_{\pi^{\prime}}= uγΓγλμπνλ+uγΓγλνπμλ,\displaystyle\,u^{\gamma}\Gamma^{\mu}_{\gamma\lambda}\pi^{\nu\lambda}+u^{\gamma}\Gamma^{\nu}_{\gamma\lambda}\pi^{\mu\lambda}\,, (147b)

are the spatial projection and geometric source terms for πμν\pi^{\mu\nu}. We also define the velocity-shear tensor σμν=ΔαβμνD(αuβ)\sigma^{\mu\nu}=\Delta^{\mu\nu}_{\alpha\beta}D^{(\alpha}u^{\beta)} and vorticity tensor ωμν=ΔαμΔβνD[αuβ]\omega^{\mu\nu}=\Delta^{\mu}_{\alpha}\Delta^{\nu}_{\beta}D^{[\alpha}u^{\beta]}.

The components of σμν\sigma^{\mu\nu} and 𝒢πμν\mathcal{G}^{\mu\nu}_{\pi^{\prime}} are the same as σμν\sigma_{\perp}^{\mu\nu} and 𝒢πμν\mathcal{G}^{\mu\nu}_{\pi} after replacing ΞαβμνΔαβμν\Xi^{\mu\nu}_{\alpha\beta}\to\Delta^{\mu\nu}_{\alpha\beta} and πμνπμν\pi_{\perp}^{{\mu\nu}}\to\pi^{\mu\nu} in Eqs. (126) and (133), respectively. The components of ωμν\omega^{\mu\nu} are

ωμν=D[μuν]uμaνuνaμ2,\omega^{\mu\nu}=D^{[\mu}u^{\nu]}-\frac{u^{\mu}a^{\nu}{-}u^{\nu}a^{\mu}}{2}\,, (148)

where D[μuν]D^{[\mu}u^{\nu]} is given by Eq. (130).

E.2 Transport coefficients

In quasiparticle viscous hydrodynamics [79] (i.e. m=m(T)m=m(T) from Fig. 2a), the relaxation times (τπ\tau_{\pi}, τΠ\tau_{\Pi}) and first-order transport coefficients (βπ\beta_{\pi}, βΠ\beta_{\Pi}) are given by Eqs. (31) – (32). The second-order transport coefficients are [66]

τππ=\displaystyle\tau_{\pi\pi}= 10+4c¯πm2227,\displaystyle\,\frac{10+4\bar{c}_{\pi}m^{2}\mathcal{I}_{22}}{7}\,, (149a)
δππ=\displaystyle\delta_{\pi\pi}= 43+c¯π22(m23mdmd(+𝒫eq)),\displaystyle\,\frac{4}{3}+\bar{c}_{\pi}\mathcal{I}_{22}\Big{(}\frac{m^{2}}{3}-m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{\text{eq}})\Big{)}\,, (149b)
λπΠ=\displaystyle\lambda_{\pi\Pi}= 652m415(c¯00+c¯Π01),\displaystyle\,\frac{6}{5}-\frac{2m^{4}}{15}\big{(}\bar{c}_{\mathcal{E}}\mathcal{I}_{00}+\bar{c}_{\Pi}\mathcal{I}_{01}\big{)}\,, (149c)
δΠΠ=\displaystyle\delta_{\Pi\Pi}=  1cs2m49(c¯00+c¯Π01)\displaystyle\,1-c_{s}^{2}-\frac{m^{4}}{9}\big{(}\bar{c}_{\mathcal{E}}\mathcal{I}_{00}+\bar{c}_{\Pi}\mathcal{I}_{01}\big{)} (149d)
mdmd(+𝒫eq)(c¯21+53c¯Π22+3m2),\displaystyle-m\frac{dm}{d\mathcal{E}}(\mathcal{E}{+}\mathcal{P}_{\text{eq}})\Big{(}\bar{c}_{\mathcal{E}}\mathcal{I}_{21}+\frac{5}{3}\bar{c}_{\Pi}\mathcal{I}_{22}+\frac{3}{m^{2}}\Big{)}\,,
λΠπ=\displaystyle\lambda_{\Pi\pi}= 13cs2+c¯πm2223,\displaystyle\,\frac{1}{3}-c_{s}^{2}+\frac{\bar{c}_{\pi}m^{2}\mathcal{I}_{22}}{3}\,, (149e)

where

c¯π=\displaystyle\bar{c}_{\pi}= 1242,\displaystyle\,\frac{1}{2\mathcal{I}_{42}}\,, (150a)
c¯=\displaystyle\bar{c}_{\mathcal{E}}= 41534042412,\displaystyle\,-\frac{\mathcal{I}_{41}}{\frac{5}{3}\mathcal{I}_{40}\mathcal{I}_{42}-\mathcal{I}_{41}^{2}}\,, (150b)
c¯Π=\displaystyle\bar{c}_{\Pi}= 40534042412,\displaystyle\,\frac{\mathcal{I}_{40}}{\frac{5}{3}\mathcal{I}_{40}\mathcal{I}_{42}-\mathcal{I}_{41}^{2}}\,, (150c)

and the function nq\mathcal{I}_{nq} are given by Eq. (33).

In standard viscous hydrodynamics [67, 20] (i.e. m¯=m/T1\bar{m}=m/T\ll 1), the relaxation times are given by Eq. (34) and

βπ\displaystyle\beta_{\pi}\approx +𝒫eq5+O(m¯2),\displaystyle\,\frac{\mathcal{E}{+}\mathcal{P}_{\text{eq}}}{5}+O(\bar{m}^{2})\,, (151a)
βΠ\displaystyle\beta_{\Pi}\approx  15(13cs2)2(+𝒫eq)+O(m¯5).\displaystyle\,15\Big{(}\frac{1}{3}-c_{s}^{2}\Big{)}^{2}(\mathcal{E}{+}\mathcal{P}_{\text{eq}})+O(\bar{m}^{5})\,. (151b)

The second-order transport coefficients (149) reduce to

τππ\displaystyle\tau_{\pi\pi}\approx 107+O(m¯2),\displaystyle\,\frac{10}{7}+O(\bar{m}^{2})\,, (152a)
δππ\displaystyle\delta_{\pi\pi}\approx 43+O(m¯2),\displaystyle\,\frac{4}{3}+O(\bar{m}^{2})\,, (152b)
λπΠ\displaystyle\lambda_{\pi\Pi}\approx 65+O(m¯2lnm¯),\displaystyle\,\frac{6}{5}+O(\bar{m}^{2}\ln\bar{m})\,, (152c)
δΠΠ\displaystyle\delta_{\Pi\Pi}\approx 23+O(m¯2lnm¯),\displaystyle\,\frac{2}{3}+O(\bar{m}^{2}\ln\bar{m})\,, (152d)
λΠπ\displaystyle\lambda_{\Pi\pi}\approx 85(13cs2)2+O(m¯4).\displaystyle\,\frac{8}{5}\Big{(}\frac{1}{3}-c_{s}^{2}\Big{)}^{2}+O(\bar{m}^{4})\,. (152e)

Both models use the same shear and bulk viscosities as anisotropic hydrodynamics (e.g. Fig. 3).

E.3 Reconstruction formulas for the energy density and fluid velocity

We reconstruct the energy density by solving the following nonlinear equation via Newton’s method [5, 20]:

f()=0,f(\mathcal{E})=0\,, (153)

where

f()=(M¯τ)(M¯τ+𝒫eq+Π)(M¯x)2(M¯y)2(τM¯η)2,f(\mathcal{E})=\left(\bar{M}^{\tau}{-}\mathcal{E}\right)\left(\bar{M}^{\tau}{+}\mathcal{P}_{\text{eq}}{+}\Pi\right)-(\bar{M}^{x})^{2}-(\bar{M}^{y})^{2}-(\tau\bar{M}^{\eta})^{2}\,, (154)

with M¯μ=Tτμπτμ\bar{M}^{\mu}=T^{\tau\mu}-\pi^{\tau\mu} and 𝒫eq=𝒫eq()\mathcal{P}_{\text{eq}}=\mathcal{P}_{\text{eq}}(\mathcal{E}) being the QCD equation of state. Using the previous energy density for the initial guess, we iterate the solution to Eq. (153) as

f()df/d,\mathcal{E}\leftarrow\mathcal{E}-\frac{f(\mathcal{E})}{df/d\mathcal{E}}\,, (155)

where

dfd=cs2(M¯τ)M¯τ𝒫eqΠ,\frac{df}{d\mathcal{E}}=c_{s}^{2}(\bar{M}^{\tau}{-}\mathcal{E})-\bar{M}^{\tau}-\mathcal{P}_{\text{eq}}-\Pi\,, (156)

with cs2=d𝒫eq/dc_{s}^{2}=d\mathcal{P}_{\text{eq}}/d\mathcal{E} being the QCD speed of sound squared. We repeat the iteration (155) until we achieve sufficient convergence or the energy density falls below min\mathcal{E}_{\text{min}}. If the bulk viscous pressure Π<𝒫eq\Pi<-\mathcal{P}_{\text{eq}}, we regulate it so that 𝒫eq+Π=0\mathcal{P}_{\text{eq}}+\Pi=0; this allows us to solve for \mathcal{E} explicitly [5]:

=M¯τ(M¯x)2+(M¯y)2+(τM¯η)2M¯τ.\mathcal{E}=\bar{M}^{\tau}-\frac{(\bar{M}^{x})^{2}{+}(\bar{M}^{y})^{2}{+}(\tau\bar{M}^{\eta})^{2}}{{\bar{M}}^{\tau}}\,. (157)

Afterwards, we regulate the energy density via Eq. (63) and evaluate the fluid velocity components as

ux\displaystyle u^{x} =M¯x(+𝒫eq+Π)(M¯τ+𝒫eq+Π),\displaystyle=\frac{\bar{M}^{x}}{\sqrt{\left(\mathcal{E}{+}\mathcal{P}_{\text{eq}}{+}\Pi\right)\left(\bar{M}^{\tau}{+}\mathcal{P}_{\text{eq}}{+}\Pi\right)}}\,, (158a)
uy\displaystyle u^{y} =M¯y(+𝒫eq+Π)(M¯τ+𝒫eq+Π),\displaystyle=\frac{\bar{M}^{y}}{\sqrt{\left(\mathcal{E}{+}\mathcal{P}_{\text{eq}}{+}\Pi\right)\left(\bar{M}^{\tau}{+}\mathcal{P}_{\text{eq}}{+}\Pi\right)}}\,, (158b)
uη\displaystyle u^{\eta} =M¯η(+𝒫eq+Π)(M¯τ+𝒫eq+Π).\displaystyle=\frac{\bar{M}^{\eta}}{\sqrt{\left(\mathcal{E}{+}\mathcal{P}_{\text{eq}}{+}\Pi\right)\left(\bar{M}^{\tau}{+}\mathcal{P}_{\text{eq}}{+}\Pi\right)}}\,. (158c)

E.4 Regulating the shear stress and bulk viscous pressure

In this regulation scheme, we first adjust the shear stress components

πηη\displaystyle\pi^{\eta\eta}\leftarrow 1τ2(1+u2)[πxx((ux)2(uτ)2)+πyy((uy)2(uτ)2)\displaystyle\,\frac{1}{\tau^{2}\left(1{+}u_{\perp}^{2}\right)}\Big{[}\pi^{xx}\big{(}(u^{x})^{2}{-}(u^{\tau})^{2}\big{)}+\pi^{yy}\big{(}(u^{y})^{2}{-}(u^{\tau})^{2}\big{)}
+2(πxyuxuy+τ2(πxηux+πyηuy)uη)],\displaystyle+2\big{(}\pi^{xy}u^{x}u^{y}+\tau^{2}(\pi^{x\eta}u^{x}{+}\pi^{y\eta}u^{y})u^{\eta}\big{)}\Big{]}\,, (159a)
πτx\displaystyle\pi^{\tau x}\leftarrow πxxux+πxyuy+τ2πxηuηuτ,\displaystyle\,\frac{\pi^{xx}u^{x}+\pi^{xy}u^{y}+\tau^{2}\pi^{x\eta}u^{\eta}}{u^{\tau}}\,, (159b)
πτy\displaystyle\pi^{\tau y}\leftarrow πxyux+πyyuy+τ2πyηuηuτ,\displaystyle\,\frac{\pi^{xy}u^{x}+\pi^{yy}u^{y}+\tau^{2}\pi^{y\eta}u^{\eta}}{u^{\tau}}\,, (159c)
πτη\displaystyle\pi^{\tau\eta}\leftarrow πxηux+πyηuy+τ2πηηuηuτ,\displaystyle\,\frac{\pi^{x\eta}u^{x}+\pi^{y\eta}u^{y}+\tau^{2}\pi^{\eta\eta}u^{\eta}}{u^{\tau}}\,, (159d)
πττ\displaystyle\pi^{\tau\tau}\leftarrow πτxux+πτyuy+τ2πτηuηuτ,\displaystyle\,\frac{\pi^{\tau x}u^{x}+\pi^{\tau y}u^{y}+\tau^{2}\pi^{\tau\eta}u^{\eta}}{u^{\tau}}\,, (159e)

so that πμν\pi^{\mu\nu} satisfies the orthogonality and tracelessness conditions

πμνuν\displaystyle\pi^{\mu\nu}u_{\nu} =0,\displaystyle=0\,, (160a)
πμμ\displaystyle\pi^{\mu}_{\mu} =0.\displaystyle=0\,. (160b)

Then we regulate the shear stress and bulk viscous pressure as

πμν\displaystyle\pi^{\mu\nu} γregπμν,\displaystyle\leftarrow\gamma_{\text{reg}}\pi^{\mu\nu}\,, (161a)
Π\displaystyle\Pi γregΠ,\displaystyle\leftarrow\gamma_{\text{reg}}\Pi\,, (161b)

where

γreg=min(1,3𝒫eq2ππ+3Π2),\gamma_{\text{reg}}=\min\Bigg{(}1,\sqrt{\frac{3\mathcal{P}_{\text{eq}}^{2}}{\pi{\,\cdot\,}\pi+3\Pi^{2}}}\Bigg{)}\,, (162)

with ππ=πμνπμν\pi{\,\cdot\,}\pi=\pi_{{\mu\nu}}\pi^{\mu\nu}. The regulation factor γreg\gamma_{\text{reg}} usually suppresses πμν\pi^{\mu\nu} and Π\Pi around the edges of the fireball at early times τ<1\tau<1 fm/cc, especially in standard viscous hydrodynamics (e.g. see Figs. (19c) and (20c)).

References