Skip to main content
Log in

Hybrid model with viscous relativistic hydrodynamics: a role of constraints on the shear-stress tensor

  • Regular Article - Theoretical Physics
  • Published:
The European Physical Journal A Aims and scope Submit manuscript

Abstract

We present the hybrid model connecting the parton–hadron–string dynamic model (PHSD) and a hydrodynamic model taking into account shear viscosity within the Israel–Stewart approach. The numerical scheme, initialization, and particlization procedure are discussed in detail. The performance of the code is tested on the pion and proton rapidity and transverse mass distributions calculated for Au + Au and Pb+Pb collisions at AGS–SPS energies. The influence of the switch time from transport to hydro models, the viscous parameter, and freeze-out time are discussed. Since the applicability of the Israel–Stewart hydrodynamics assumes the perturbative character of the viscous stress tensor, \(\pi ^{\mu \nu }\), which should not exceed the ideal energy–momentum tensor, \(T_{\mathrm{id}}^{\mu \nu }\), hydrodynamical codes usually rescale the shear stress tensor if the inequality \(\Vert \pi ^{\mu \nu }\Vert \ll \Vert T_{\mathrm{id}}^{\mu \nu }\Vert \) is not fulfilled in some sense. There are several conditions used in the literature and we analyze in detail the influence of different conditions and values of the cut-off parameter on observables. We show that the form of the corresponding condition plays an important role in the sensitivity of hydrodynamic calculations to the viscous parameter – a ratio of the shear viscosity to the entropy density, \(\eta /s\). It is shown that the constraints used in the vHLLE and MUSIC models give the same results for the observables. With these constraints, the rapidity distributions and transverse momentum spectra are most sensitive to a change of the \(\eta /s\) ratio. We demonstrate that these constraints do not guarantee that each element of the \(\pi ^{\mu \nu }\) tensor is smaller than the corresponding element \(T_{\mathrm{id}}^{\mu \nu }\). As an alternative, a strict condition is used. When applied it reduces the sensitivity of the proton and pion momentum distributions to the viscosity parameter. We performed global fits of the rapidity and transverse mass distributions of pion and protons. It was also found that \(\eta /s\) as a function of the collision energy monotonically increases from \(E_{\mathrm{lab}}=6\,A\text {GeV}\) up to \(E_{\mathrm{lab}}=40\,A\text {GeV}\) and saturates for higher SPS energies. We observe that it is difficult to reproduce simultaneously pion and proton rapidity distribution within our model with the present choice of the equation of state without a phase transition.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16
Fig. 17
Fig. 18

Similar content being viewed by others

Data Availability Statement

This manuscript has no associated data or the data will not be deposited. [Authors’ comment: This is a theoretical study, there are not external data associated with the manuscript.]

Notes

  1. Recall that in the standard implementation of SHASTA, e.g., for solving ideal hydrodynamics in UrQMD, the mid-point rule is used to achieve the second-order accuracy in time.

  2. Equation (11) is applied before the antidiffusion step.

  3. It was noted [36, 37] that the system created in high energy nuclear collisions reaches or at least comes close to equilibrium. In particular, it was realized that because of the expansion of the matter into the vacuum, the system would cool and thus freeze into a hadronic gas quickly. Thus, it became apparent that a fluid dynamic approximation to the system dynamics had to start early, on a time-scale of \(\tau \sim 1\) fm/c or less.

  4. Let us also mention that one has to be careful with conclusions of [37] since as is seen from Fig. 5 there, \(P_T\ne 0\) when \(P_{\mathrm{eq}}=0\). It is not the case of usual matter.

  5. Generally, one can use different mask coefficients, \(A_{\mathrm{ad}}^{x,y,z}\), see Eq. (A9), for every direction, x, y, and z, but for simplicity we take one value for all axes.

  6. Particularly we use version 1.0 of the Parton–Hadron String Dynamics model with the switched off partonic option.

  7. As one can see from the function QuestRevert of MUSIC code or [82], the developers use an energy-dependent cut-off parameter \(C=C(\varepsilon )\) in Eq. (13). We take just a constant value for simplicity. Our results do not changes if one takes \(C(\varepsilon )\propto \tanh \frac{\varepsilon }{\varepsilon _0}\) with small \(\varepsilon _0\).

References

  1. E. Molnár, H. Niemi, D.H. Rischke, Numerical tests of causal relativistic dissipative fluid dynamics. Eur. Phys. J. C 65, 615 (2010)

    Article  ADS  Google Scholar 

  2. L.D. Landau, On the multiparticle production in high-energy collisions. Izv. Akad. Nauk Ser. Fiz. 17, 5164 (1953)

    Google Scholar 

  3. L.D. Landau, Collected Papers of L.D.Landau, Ed. D. Ter-Haar (Pergamon Press, Oxford, 1965), p. 74

    Google Scholar 

  4. P.F. Kolb, U.W. Heinz, Hydrodynamic Description of Ultrarelativistic Heavy Ion Collisions, Quark-Gluon Plasma 3, Ed. by R. Hwa and X.-N. Wang (World Scientific, Singapore, 2004), p. 634

    Google Scholar 

  5. U.W. Heinz, R. Snellings, Collective flow and viscosity in relativistic heavy-ion collisions. Annu. Rev. Nucl. Part. Sci. 63, 123 (2013)

    Article  ADS  Google Scholar 

  6. C. Gale, S. Jeon, B. Schenke, Hydrodynamic modeling of heavy-ion collisions. Int. J. Mod. Phys. A 28, 1340011 (2013)

    Article  ADS  Google Scholar 

  7. S. Jeon, U. Heinz, Introduction to hydrodynamics. Int. J. Mod. Phys. E 24, 1530010 (2015)

    Article  ADS  Google Scholar 

  8. S. Jeon, U. Heinz, Quark-Gluon plasma 5, Ed. by X.-N.Wang (World Scientific, Singapore, 2016) [arXiv: 1503.03931]

  9. R. Derradi de Souza, T. Koide, T. Kodama, Hydrodynamic approaches in relativistic heavy ion reactions. Prog. Part. Nucl. Phys. 86, 35 (2016)

    Article  ADS  Google Scholar 

  10. W. Florkowski, M.P. Heller, M. Spalinski, New theories of relativistic hydrodynamics in the LHC era. Rept. Prog. Phys. 81, 046001 (2018)

    Article  ADS  MathSciNet  Google Scholar 

  11. A.S. Khvorostukhin, V.D. Toneev, Rapidity distributions of hadrons in the HydHSD hybrid model. Phys. Atom. Nucl. 80, 285 (2017)

    Article  ADS  Google Scholar 

  12. A.S. Khvorostukhin, V.D. Toneev, Hadron rapidity spectra within a hybrid model, Phys. Part. Nucl. Lett. 14, 9 (2017) [arXiv: 1606.00987]

  13. H. Song, U. Heinz, Suppression of elliptic flow in a minimally viscous quark-gluon plasma. Phys. Lett. B 658, 279 (2008)

    Article  ADS  Google Scholar 

  14. H. Song, U. Heinz, Causal viscous hydrodynamics in 2 + 1 dimensions for relativistic heavy-ion collisions. Phys. Rev. C 77, 064901 (2008)

    Article  ADS  Google Scholar 

  15. L.D. Landau, E.M. Lifshitz, Fluid Mechanics (Pergamon Press, Oxford, 1987)

    Google Scholar 

  16. G.S. Denicol, E. Molnár, H. Niemi, D.H., Rischke Derivation of fluid dynamics from kinetic theory with the 14-moment approximation. Eur. Phys. J. A 48, 170 (2012)

  17. W. Israel, J.M. Stewart, Transient relativistic thermodynamics and kinetic theory. Ann. Phys. 118, 341 (1979)

    Article  ADS  MathSciNet  Google Scholar 

  18. Iu. Karpenko, P. Huovinen, M. Bleicher, 3+1 dimensional viscous hydrodynamic code for relativistic heavy ion collisions. Comput. Phys. Commun. 185, 3016 (2014)

    Article  ADS  MATH  Google Scholar 

  19. B. Schenke, S. Jeon, C. Gale, Elliptic and Triangular Flow in Event-by-Event D\(=3+1\) Viscous Hydrodynamics, Phys. Rev. Lett. 106, 042301 (2011) (www.physics.mcgill.ca/music/)

  20. F.S. Bemfica, M.M. Disconzi, V. Hoang, J. Noronha, M. Radosz, Nonlinear Constraints on Relativistic Fluids Far From Equilibrium. Phys. Rev. Lett. 126, 222301 (2021)

    Article  ADS  MathSciNet  Google Scholar 

  21. Ch. Plumberg, D. Almaalol, T. Dore, J. Noronha, J. Noronha-Hostler, Causality violations in realistic simulations of heavy-ion collisions, arXiv:2103.15889

  22. E. Molnár, H. Holopainen, P. Huovinen, H. Niemi, Influence of temperature-dependent shear viscosity on elliptic flow at backward and forward rapidities in ultrarelativistic heavy-ion collisions. Phys. Rev. C 90, 044904 (2014)

    Article  ADS  Google Scholar 

  23. A. Muronga, Relativistic dynamics of nonideal fluids: Viscous and heat-conducting fluids. I. General aspects and 3+1 formulation for nuclear collisions. Phys. Rev. C 76, 014909 (2007)

    Article  ADS  Google Scholar 

  24. J.P. Boris, D.L. Book, Flux-corrected transport. I. SHASTA, a fluid transport algorithm that works. J. Comp Phys. A 11, 38 (1973)

    Article  ADS  MATH  Google Scholar 

  25. D.L. Book, J.P. Boris, K. Hain, Flux-corrected transport II: Generalizations of the method. J. Comp. Phys. A 18, 248 (1975)

    Article  ADS  MATH  Google Scholar 

  26. D.H. Rischke, S. Bernard, J.A. Maruhn, Relativistic hydrodynamics for heavy-ion collisions. I. General aspects and expansion into vacuum. Nucl. Phys. A 595, 346 (1995)

    Article  ADS  Google Scholar 

  27. D.H. Rischke, Fluid dynamics for relativistic nuclear collisions. In: Cleymans J., Geyer H.B., Scholtz F.G. (eds) Hadrons in Dense Matter and Hadrosynthesis. Lecture Notes in Physics, vol 516. Springer, Berlin, Heidelberg (1999). https://doi.org/10.1007/BFb0107310, (arXiv:nucl-th/9809044)

  28. E. Molnár, Comparing the first- and second-order theories of relativistic dissipative fluid dynamics using the \(1+1\) dimensional relativistic flux corrected transport algorithm. Eur. Phys. J. C 60, 413 (2009)

    Article  ADS  MATH  Google Scholar 

  29. Iu.A. Karpenko, P. Huovinen, H. Petersen, M. Bleicher, Estimation of the shear viscosity at finite net-baryon density from A+A collision data at \(\sqrt{s_{NN}}\) = 7.7-200ĠeV, Phys. Rev. C 91, 064901 (2015)

  30. C. Shen, Z. Qiu, H. Song, J. Bernhard, S. Bass, U. Heinz, The iEBE-VISHNU code package for relativistic heavy-ion collisions. Comp. Phys. Commun 199, 61 (2016)

    Article  ADS  MathSciNet  Google Scholar 

  31. E. Shuryak, Why does the quark gluon plasma at RHIC behave as a nearly ideal fluid? Prog. Part. Nucl. Phys. 53, 273 (2004)

    Article  ADS  Google Scholar 

  32. E.V. Shuryak, What RHIC experiments and theory tell us about properties of quark-gluon plasma? Nucl. Phys. A 750, 64 (2005)

    Article  ADS  Google Scholar 

  33. U. W. Heinz, Thermalization at RHIC, AIP Conf. Proc. 739, 163 (2005) [arXiv: nucl-th/0407067]

  34. J.M. Maldacena, The large \(N\)-limit of superconformal field theories and supergravity. Adv. Theor. Math. Phys. 2, 231 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. E. Witten, Anti De Sitter Space And Holography. Adv. Theor. Math. Phys. 2, 253 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  36. P. Romatschke, Do nuclear collisions create a locally equilibrated quark-gluon plasma? Eur. Phys. J. C 77, 21 (2017)

    Article  ADS  Google Scholar 

  37. M. Attems, Y. Bea, J. Casalderrey-Solana, D. Mateos, M. Trianaa, M. Zilhão, Phase transitions, inhomogeneous horizons and second-order hydrodynamics. JHEP 2017, 129 (2017)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. M.P. Heller, R.A. Janik, P. Witaszczyk, Characteristics of thermalization of boost-invariant plasma from holography. Phys. Rev. Lett. 108, 201602 (2012)

    Article  ADS  Google Scholar 

  39. B. Wu, P. Romatschke, Shock wave collisions in AdS\(_5\): approximate numerical solutions. Int. J. Mod. Phys. C 22, 1317 (2011)

    Article  ADS  MATH  Google Scholar 

  40. L. Keegan, A. Kurkela, P. Romatschke, W. van der Scheee, Y. Zhuf, Weak and strong coupling equilibration in nonabelian gauge theories. JHEP 2016, 31 (2016)

    Article  Google Scholar 

  41. V.V. Skokov, V.D. Toneev, Hydrodynamics of an Expanding Fireball. Phys. Atom. Nucl. 70, 109 (2007)

    Article  ADS  Google Scholar 

  42. L. Del Zanna, V. Chandra, G. Inghirami, V. Rolando, A. Beraudo, A. De Pace, G. Pagliara, A. Drago, F. Becattini, Relativistic viscous hydrodynamics for heavy-ion collisions with ECHO-QGP. Eur. Phys. J. C 73, 2524 (2013)

    Article  ADS  Google Scholar 

  43. H. Niemi, G.S. Donicol, P. Huovinen, E. Molnar, D.H. Rischke, Influence of a temperature-dependent share viscosity on the azimuthal asymmertries of transverse momentum spectra in ultrarelativistic heavy-ion collisions. Phys. Rev. C 86, 014909 (2012)

    Article  ADS  Google Scholar 

  44. L. Ahle et al., (E802 Collaboration) Particle production at high baryon density in central Au+Au reactions at 11.6 \(A\) GeV/\(c\). Phys. Rev. C 57, R466 (1988)

    Article  ADS  Google Scholar 

  45. J. Barrette et al., (E877 Collaboration), Proton and pion production in Au+Au collisions at 10.8 \(A\)GeV/\(c\). Phys. Rev. C 62, 024901 (2000)

    Article  ADS  Google Scholar 

  46. B.B. Back et al., (E917 Collaboration), Baryon rapidity loss in relativistic Au\(+\)Au collisions. Phys. Rev. Lett. 86, 1970 (2001)

    Article  ADS  Google Scholar 

  47. W. Ehehalt, W. Cassing, Relativistic transport approach for nucleus-nucleus collisions from SIS to SPS energies. Nucl. Phys. A 602, 449 (1996)

    Article  ADS  Google Scholar 

  48. J. Geiss, W. Cassing, C. Greiner, Strangeness production in the HSD transport approach from SIS to SPS energies. Nucl. Phys. A 644, 107 (1998)

    Article  ADS  Google Scholar 

  49. W. Cassing, E.L. Bratkovskaya, Hadronic and electromagnetic probes of hot and dense nuclear matter. Phys. Rept. 308, 65 (1999)

    Article  ADS  Google Scholar 

  50. D. Oliinychenko, H. Petersen, Deviations of the energy-momentum tensor from equilibrium in the initial State for hydrodynamics from transport approaches. Phys. Rev. C 93, 034905 (2016)

    Article  ADS  Google Scholar 

  51. F.G. Gardim, F. Grassi, Y. Hama, M. Luzum, J.-Y. Ollitrault, Directed flow at midrapidity in event-by-event hydrodynamics. Phys. Rev. C 83, 064901 (2011)

    Article  ADS  Google Scholar 

  52. P. Huovinen, H. Petersen, Particlization in hybrid models. Eur. Phys. J. A 48, 171 (2006)

    Article  ADS  Google Scholar 

  53. P. Arnold, G.D. Moore, L.G. Yaffe, Transport coeffcients in high temperature gauge theories, 1. Leading-log results. J. High. Energy Phys. 11, 001 (2000)

    Article  ADS  Google Scholar 

  54. D. Teaney, Effect of shear viscosity on spectra, elliptic flow, and Hanbury Brown-Twiss radii. Phys. Rev. C 68, 034913 (2003)

    Article  ADS  Google Scholar 

  55. M. McNelis, D. Everett, U. Heinz, Particlization in fluid dynamical simulations of heavy-ion collisions. The iS3D module, arXiv:1912.08271

  56. H. Petersen, J. Steinheimer, G. Burau, M. Bleicher, H., Stöcker Fully integrated transport approach to heavy ion reactions with an intermediate hydrodynamic stage. Phys. Rev. C 78, 044901 (2008)

  57. N.S. Amelin, R. Lednicky, T.A. Pocheptsov, I.P. Lokhtin, L.V. Malinina, A.M. Snigirev, Iu.A. Karpenko, M. Yu, Sinyukov, Fast hadron freeze-out generator. Phys. Rev. C 74, 064901 (2006)

    Article  ADS  Google Scholar 

  58. L.M. Satarov, M.N. Dmitriev, I.N. Mishustin, Equation of state of hadron resonance gas and the phase diagram of strongly interacting matter. Phys. Atom. Nucl. 72, 1390 (2009)

    Article  ADS  Google Scholar 

  59. J.D. Ramshaw, Partial chemical equilibrium in fluid dynamics. Phys. Fluids 23, 675 (1980)

    Article  ADS  MATH  Google Scholar 

  60. A. Monnai, T. Hirano, Relativistic dissipative hydrodynamic equations at the second order for multi-component systems with multiple conserved currents. Nucl. Phys. A 847, 283 (2010)

    Article  ADS  Google Scholar 

  61. K. Peach, M. Reiter, A. Dumitru, H. Stöcker, W. Greiner, On the observation of phase transitions in collisions of elementary matter. Nucl. Phys. A 681, 41 (2001)

    Article  ADS  Google Scholar 

  62. T. Anticic et al., (NA49 Collaboration), Centrality dependence of proton and antiproton spectra in Pb+Pb collisions at 40 \(A\)GeV and 158 \(A\)GeV measured at the CERN Super Proton Synchrotron. Phys. Rev. C 83, 014901 (2011)

    Article  ADS  Google Scholar 

  63. T. Anticic et al., (NA49 Collaboration), Energy and centrality dependence of deuteron and proton production in Pb+Pb collisions at relativistic energies. Phys. Rev. C 69, 024902 (2004)

    Article  ADS  Google Scholar 

  64. S.V. Afanasiev et al., (The NA49 Collaboration), Energy dependence of pion and kaon production in central Pb+Pb collisions. Phys. Rev. C 66, 054902 (2002)

    Article  ADS  Google Scholar 

  65. L. Du, U. Heinz, \((3+1)\)-dimensional dissipative relativistic fluid dynamics at non-zero net baryon density. Comput. Phys. Commun. 251, 107090 (2020)

    Article  MathSciNet  Google Scholar 

  66. J.L. Klay et al., (E895 Collaboration), Longitudinal flow of protons from 2\(A\)-8\(A\) GeV central Au+Au collisions. Phys. Rev. Lett 88, 102301 (2002)

  67. J.L. Klay et al., (E895 Collaboration), Charged pion production in 2\(A\) to 8\(A\) GeV central Au+Au ollisions. Phys. Rev. C 68, 054905 (2003)

    Article  ADS  Google Scholar 

  68. L. Ahle et al., (E802 Collaboration), Simultaneous multiplicity and forward energy characterization of particle spectra in Au+Au collisions at 11.6 \(A\)GeV/\(c\). Phys. Rev. C 59, 2173 (1999)

    Article  ADS  Google Scholar 

  69. Y. Akiba et al., (E802 Collaboration), Particle Production in Au+Au collisions from BNL E866. Nucl. Phys. A 610, 139c (1996)

    Article  ADS  Google Scholar 

  70. C. Alt et al., (NA49 Collaboration), Energy and centrality dependence of \({\bar{p}}\) and \(p\) production and the \({\bar{\Lambda }}/{\bar{p}}\) ratio in Pb+Pb collisions between 20A GeV and 158A GeV. Phys. Rev. C 73, 044910 (2006)

    Article  ADS  Google Scholar 

  71. C. Blume, Recent results from the NA49 experiment. J. Phys. G 35, 044004 (2008)

    Article  ADS  Google Scholar 

  72. S.V. Afanasiev et al., Recent results on spectra and yields from NA49. Nucl. Phys. A 715, 161c (2003)

    Article  ADS  Google Scholar 

  73. A. Andronic, P. Braun-Munzinger, J. Stachel, Hadron production in central nucleus nucleus collisions at chemical freeze-out. Nucl. Phys. A 772, 167 (2006)

    Article  ADS  Google Scholar 

  74. G. Denicol, A. Monnai, B. Schenke, Moving forward to constrain the shear viscosity of QCD matter. Phys. Rev. Lett 116, 212301 (2016)

    Article  ADS  Google Scholar 

  75. Yu.B. Ivanov, V.N. Russkikh, V.D. Toneev, Relativistic heavy-ion collisions within three-fluid hydrodynamics: Hadronic scenario. Phys. Rev. C 73, 044904 (2006)

    Article  ADS  Google Scholar 

  76. Yu.B. Ivanov, Baryon stopping signal for mixed phase formation in HIC. J. Phys. Conf. Ser. 668, 012061 (2016)

    Article  Google Scholar 

  77. Yu.B. Ivanov, Alternative Scenarios of Relativistic Heavy-Ion Collisions: III. Transverse Momentum Spectra. Phys. Rev. C 89, 024903 (2014)

    Article  ADS  Google Scholar 

  78. W. Florkowski, W. Broniowski, Hydro-inspired parametrization of freeze-out in relativistic heavy-ions. Acta Phys. Pol. B 35, 2895 (2004)

    ADS  Google Scholar 

  79. W. Florkowski, Particle spectra and hydro-inspired model. Nucl. Phys. A 774, 179 (2006)

    Article  ADS  Google Scholar 

  80. E. Schnedermann, J. Sollfrank, U. Heinz, Thermal phenomenology of hadrons from 200-A/GeV S+S collisions. Phys. Rev. C 48, 2462 (1993)

    Article  ADS  Google Scholar 

  81. S.P. Rode, P.P. Bhaduri, A. Jaiswal, A. Roy, Kinetic freeze out conditions in nuclear collisions with 2–150 A GeV beam energy within a non boost-invariant blast wave model. Phys. Rev. C 98, 024907 (2018)

    Article  ADS  Google Scholar 

  82. G.S. Denicol, Ch. Gale, S. Jeon, A. Monnai, B. Schenke, Ch. Shen, Net baryon diffusion in fluid dynamic simulations of relativistic heavy-ion collisions, Phys. Rev. C 98, 034916 (2018) [arXiv: 1804.10557]

  83. S.R. de Groot, W.A. van Leeuwen, Ch.G. van Weert, Relativistic Kinetic Theory-Principles and Applications (North-Holland, Amsterdam, 1980)

    Google Scholar 

  84. K. Heun, Neue Methoden zur approximativen Integration der Differentialgleichungen einer unabhängigen Veränderlichen. Z. Math. Phys. 45, 23 (1900)

    MATH  Google Scholar 

Download references

Acknowledgements

We thank E. Bratkovskaya and W. Cassing for providing the HSD code and consultations. We appreciate very much extensive discussions with Iu. Karpenko and Yu.B. Ivanov and constructive remarks by G. Sandukovskaya. The work is supported by Slovak Grant VEGA-1/0348/18 and by THOR the COST Action CA15213. A.S.K and E.E.K. acknowledge the support by the Plenipotentiary of the Slovak Government at JINR, Dubna. The work of A. Khvorostukhin was supported by the RFBR Grant no. 18-02-40137 and the NARD project, no. 20.80009.5007.07.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to E. E. Kolomeitsev.

Additional information

Communicated by Giorgio Torrieri.

Appendices

Appendix A: Numerical realization

In this Appendix we discuss the numerical scheme used to integrate the hydrodynamic equations (1b), (9), and (6). The 10-dimensional vector \(\mathbf {S}\) in the right-hand side of Eq. (10) can be written as a combinations of two 5-dimensional vectors

$$\begin{aligned} \mathbf {S}=\big (\mathbf {S}_{\mathrm{cons}},\mathbf {S}_\pi \big ) \end{aligned}$$
(A1)

corresponding to the conservation equations (1b) and (9),

$$\begin{aligned} \mathbf {S}_{\mathrm{cons}}\!=\!\! \left[ \begin{array}{l} 0 \\ -\partial _t\pi ^{00} - {{\,\mathrm{div}\,}}(\mathbf {v} P) - (\partial _x\pi ^{0x} + \partial _y\pi ^{0y} + \partial _z\pi ^{0z}) \\ -\partial _t\pi ^{0x} - \partial _xP - (\partial _x\pi ^{xx}+\partial _y\pi ^{xy}+\partial _z\pi ^{xz}) \\ -\partial _t\pi ^{0y} - \partial _yP - (\partial _x\pi ^{yx} +\partial _y\pi ^{yy} + \partial _z\pi ^{yz}) \\ -\partial _t\pi ^{0z} - \partial _zP - (\partial _x\pi ^{zx} + \partial _y\pi ^{zy}+\partial _z\pi ^{zz}) \end{array} \right] ,\nonumber \\ \end{aligned}$$
(A2)

and to Israel–Stewart relaxation equations (6) for viscous fields

$$\begin{aligned} \mathbf {S}_\pi&= \big [Q^{xy}, Q^{xz}, Q^{yz}, Q^{yy}, Q^{zz}\big ]^{\mathrm{T}} \nonumber \\ Q^{\mu \nu }&=\pi ^{\mu \nu }\left( \mathrm{div}\mathbf {v} -\frac{1}{\gamma \tau _\pi }\right) +\frac{\eta }{\gamma \tau _\pi }\,W^{\mu \nu }. \end{aligned}$$
(A3)

In Ref. [43] it was noted that the algorithm could become more stable if the relaxation equations are solved by a simple centered second-order differences scheme for spatial gradients on the left-hand side of Eqs. (6). We have tested such a separation for the full \(3+1\)D calculations and find out that it leads to uncontrolled solutions. The same phenomenon was observed also for calculation done in the Milne coordinates in Ref. [22], where the authors used also the full SHASTA method for both conservation and relaxation equations. We think that such behaviour is caused by weak steadiness of the Euler method which leads to uncontrolled inaccuracy of a numerical solution of relaxation equations.

Thus we apply the SHASTA method to all ten equations included in Eq. (10). To reach the quadratic precision in time we use Heun’s method [84] which allows storing fewer intermediate points than the mid-point rule.

1.1 1. 3\(+\)1D implementation of the SHASTA algorithm

For completeness, we provide the complete set of formulas for the 3+1 implementation of the SHASTA algorithm extending expressions provided in Ref [1].

For lattice realization of quantities U(xyzt) we will use notations \(U^{[n]}_{ijk}\), where index n stands for temporal steps and ijk for spatial lattice cells in xy, and z directions respectively.

At the first stage of the SHASTA algorithm for the subsequent \((n +1)\)th time step, one calculates the so-called transport-diffused solution

$$\begin{aligned} {\widetilde{U}}_{ijk}^{[n+1]}&= {\widetilde{U}}_{ijk}^x + {\widetilde{U}}_{ijk}^y + {\widetilde{U}}_{ijk}^z - 2U_{ijk}^{[n]} + \Delta t\, {S}_{ijk}, \end{aligned}$$
(A4)

where \(U_{ijk}^{[n]}\) is the full solution at the previous time step and auxiliary quantities \({\tilde{U}}_{ijk}^{x,y,z}\) are defined as

$$\begin{aligned} {\widetilde{U}}_{ijk}^x&=\frac{1}{2}\left( \big [ Q_{ijk}^{x+} \big ]^2 \big (U^{[n]}_{i+1,jk} - U^{[n]}_{ijk}\big ) \right. \nonumber \\&\left. \quad - \big [Q_{ijk}^{x-}\big ]^2 \big (U^{[n]}_{ijk} - U^{[n]}_{i-1,jk}\big ) \right) +(Q_{ijk}^{x+}+Q_{ijk}^{x-}) \, U^{[n]}_{ijk}, \end{aligned}$$
(A5)
$$\begin{aligned} {\widetilde{U}}_{ijk}^y&=\frac{1}{2}\left( \big [Q_{ijk}^{y+}\big ]^2 \big ( U^{[n]}_{i,j+1,k}- U^{[n]}_{ijk} \big ) \right. \nonumber \\&\left. \quad - \big [Q_{ijk}^{y-}\big ]^2 \big ( U^{[n]}_{ijk}- U^{[n]}_{i,j-1,k} \big ) \right) +(Q_{ijk}^{y+}+Q_{ijk}^{y-}) \, U^{[n]}_{ijk}, \end{aligned}$$
(A6)
$$\begin{aligned} {\widetilde{U}}_{ijk}^z&=\frac{1}{2}\left( \big [Q_{ijk}^{z+}\big ]^2 \big (U^{[n]}_{ij,k+1} - U^{[n]}_{ijk}\big ) \right. \nonumber \\&\left. \quad - \big [Q_{ijk}^{z-}\big ]^2\big (U^{[n]}_{ijk} - U^{[n]}_{ij,k-1}\big )\right) + (Q_{ijk}^{z+} + Q_{ijk}^{z-})\, U^{[n]}_{ijk} \end{aligned}$$
(A7)

with

$$\begin{aligned} Q_{ijk}^{x\pm }&=\frac{1/2\mp \lambda \, (v_x)^{[n]}_{ijk}}{1\pm \lambda \,\left[ (v_x)^{[n]}_{i\pm 1,jk} - (v_x)^{[n]}_{ijk} \right] },\, \nonumber \\ Q_{ijk}^{y\pm }&=\frac{1/2\mp \lambda \, (v_y)^{[n]}_{ijk}}{ 1\pm \lambda \, \left[ (v_y)^{[n]}_{i,j\pm 1,k} - (v_y)^{[n]}_{ijk}\right] },\, \nonumber \\ Q_{ijk}^{z\pm }&=\frac{1/2\mp \lambda \, (v_z)^{[n]}_{ijk}}{1\pm \lambda \,\left[ (v_z)^{[n]}_{ij,k\pm 1} - (v_z)^{[n]}_{ijk} \right] }. \end{aligned}$$
(A8)

The velocity components are taken here at the nth time step. Here, parameter \(\lambda =\Delta t/\Delta x=\Delta t/\Delta y=\Delta t/\Delta z\) is the Courant number which is the same for all special directions. In the SHASTA it is restricted to values \(\lambda \le 1/2\).

Further, using the transport-diffused solution one calculates an antidiffusion flux that takes into account an anomalous diffusion

$$\begin{aligned} A^{x,y,z}_{ijk}&=\frac{1}{8} A_{\mathrm{ad}}^{x,y,z}\,{\tilde{\Delta }}_{ijk}^{x,y,z},\quad {\tilde{\Delta }}_{ijk}^{x}={\tilde{U}}_{i+1,jk}^x-{\tilde{U}}_{ijk}^x,\quad \nonumber \\ {\tilde{\Delta }}_{ijk}^{y}&={\tilde{U}}_{i,j+1,k}^y-{\tilde{U}}_{ijk}^y,\quad {\tilde{\Delta }}_{ijk}^{z}={\tilde{U}}_{ij,k+1}^z-{\tilde{U}}_{ijk}^z, \end{aligned}$$
(A9)

where \(A_{\mathrm{ad}}^{x,y,z}\) are the antidiffusive mask coefficients. For simplicity, one takes them to be equal for all special directions and set \(A_{\mathrm{ad}}=1\) as the default value. Next, we calculate the limited antidiffusion fluxes

$$\begin{aligned} \widetilde{A}_{ijk}^x&= \sigma _{ijk}^x \max \Big [0,\min \Big (\sigma _{ijk}^x{\widetilde{\Delta }}_{i+1,jk}^{x},\big |A^x_{ijk}\big |, \sigma _{ijk}^x{\widetilde{\Delta }}_{i-1,jk}^{x}\Big )\Big ],\nonumber \\ \widetilde{A}_{ijk}^y&= \sigma _{ijk}^y \max \Big [0,\min \Big (\sigma _{ijk}^y{\widetilde{\Delta }}_{i,j+1,k}^{y},\big |A^y_{ijk}\big |, \sigma _{ijk}^y{\widetilde{\Delta }}_{i,j-1,k}^{y}\Big )\Big ],\nonumber \\&\qquad \qquad \sigma _{ijk}^{x,y,z}=\mathrm{sgn} A^{x,y,z}_{ijk}. \nonumber \\ \widetilde{A}_{ijk}^z&=\sigma _{ijk}^z \max \Big [0,\min \Big (\sigma _{ijk}^z{\widetilde{\Delta }}_{ij,k+1}^{z},\big |A^z_{ijk}\big |, \sigma _{ijk}^z{\widetilde{\Delta }}_{ij,k-1}^z\Big )\Big ]. \end{aligned}$$
(A10)

The total incoming and outgoing antidiffusive fluxes in the cell are calculated as

$$\begin{aligned} A^{\mathrm{in}}_{ijk}&= \max \big (0,{\widetilde{A}}^x_{i-1,jk}\big ) -\min \big (0,{\widetilde{A}}^x_{ijk}\big ) +\max \big (0,{\widetilde{A}}^y_{i,j-1,k}\big )\nonumber \\&\quad -\min \big (0,{\widetilde{A}}^y_{ijk}\big ) +\max \big (0,{\widetilde{A}}^z_{ij,k-1}\big ) -\min \big (0,{\widetilde{A}}^z_{ijk}\big ), \end{aligned}$$
(A11)
$$\begin{aligned} A^{\mathrm{out}}_{ijk}&= \max \big (0,{\widetilde{A}}^x_{ijk}\big ) -\min \big (0,{\widetilde{A}}^x_{i-1,jk}\big ) +\max \big (0,{\widetilde{A}}^y_{ijk}\big )\nonumber \\&\quad -\min \big (0,{\widetilde{A}}^y_{i,j-1,k}\big ) +\max \big (0,{\widetilde{A}}^z_{ijk}\big ) -\min \big (0,{\widetilde{A}}^z_{ij,k-1}\big ). \end{aligned}$$
(A12)

The maximal and minimal values of the transport-diffused solution \(U^{[n+1]}_{ijk}\) after the antidiffusion stage are between

$$\begin{aligned}&{\widetilde{U}}^{\min }_{ijk}= \min \left( {\widetilde{U}}_{ij,k-1}^{[n+1]}, {\widetilde{U}}_{i,j-1,k}^{[n+1]}, {\widetilde{U}}_{i-1,jk}^{[n+1]}, {\widetilde{U}}_{ijk}^{[n+1]},\right. \nonumber \\&\left. \quad {\widetilde{U}}_{ij,k+1}^{[n+1]}, {\widetilde{U}}_{i,j+1,k}^{[n+1]}, {\widetilde{U}}_{i+1,jk}^{[n+1]}\right) , \end{aligned}$$
(A13)
$$\begin{aligned}&{\widetilde{U}}^{\max }_{ijk}= \max \left( {\widetilde{U}}_{ij,k-1}^{[n+1]}, {\widetilde{U}}_{i,j-1,k}^{[n+1]}, {\widetilde{U}}_{i-1,jk}^{[n+1]}, {\widetilde{U}}_{ijk}^{[n+1]},\right. \nonumber \\&\left. \quad {\widetilde{U}}_{ij,k+1}^{[n+1]}, {\widetilde{U}}_{i,j+1,k}^{[n+1]}, {\widetilde{U}}_{i+1,jk}^{[n+1]}\right) . \end{aligned}$$
(A14)

This information is then used to determine the fractions of the incoming and outgoing fluxes,

$$\begin{aligned} F^{\mathrm{in}}_{ijk} = \frac{{\widetilde{U}}_{ijk}^{\max } - {\widetilde{U}}_{ijk}^{[n+1]}}{A^{\mathrm{in}}_{ijk}}, \qquad F^{\mathrm{out}}_{ijk} = \frac{{\widetilde{U}}_{ijk}^{[n+1]} - {\widetilde{U}}_{ijk}^{\min }}{A^{\mathrm{out}}_{ijk}}. \end{aligned}$$
(A15)

The final anti-diffusion fluxes are calculated as

$$\begin{aligned} {\hat{A}}^x_{ijk}&={\widetilde{A}}^x_{ijk} \Big [\min (1, F^{\mathrm{in}}_{i+1,jk},F^{\mathrm{out}}_{ijk})\Theta ({\widetilde{A}}^x_{ijk})\nonumber \\&\quad +\min (1, F^{\mathrm{in}}_{ijk},F^{\mathrm{out}}_{i+1,jk})\Theta (-{\widetilde{A}}^x_{ijk})\Big ], \end{aligned}$$
(A16)
$$\begin{aligned} {\hat{A}}^y_{ijk}&={\widetilde{A}}^y_{ijk} \Big [\min (1, F^{\mathrm{in}}_{i,j+1,k},F^{\mathrm{out}}_{ijk})\Theta ({\widetilde{A}}^y_{ijk})\nonumber \\&\quad +\min (1, F^{\mathrm{in}}_{ijk},F^{\mathrm{out}}_{i,j+1,k})\Theta (-{\widetilde{A}}^y_{ijk})\Big ], \end{aligned}$$
(A17)
$$\begin{aligned} {\hat{A}}^z_{ijk}&={\widetilde{A}}^y_{ijk} \Big [\min (1, F^{\mathrm{in}}_{ij,k+1},F^{\mathrm{out}}_{ijk})\Theta ({\widetilde{A}}^z_{ijk})\nonumber \\&\quad +\min (1, F^{\mathrm{in}}_{ijk},F^{\mathrm{out}}_{ij,k+1})\Theta (-{\widetilde{A}}^z_{ijk})\Big ]. \end{aligned}$$
(A18)

Finally, the full solution for \(n+1\) time step is given by

$$\begin{aligned} U_{ijk}^{[n+1]}&={\widetilde{U}}_{ijk}^{[n+1]} +\big ({\hat{A}}^x_{i-1,jk}-{\hat{A}}^x_{ijk}\big )\nonumber \\&\quad +\big ({\hat{A}}^y_{i,j-1,k}-{\hat{A}}^y_{ijk}\big ) +\big ({\hat{A}}^z_{ij,k-1}-{\hat{A}}^z_{ijk}\big ). \end{aligned}$$
(A19)

In Fig. 19 we compare the numerical results of the 3\(+\)1D SHASTA code with the one-pass method in the time evolution, as given by Eq. (A4) and of the SHASTA code improved by Heun’s method with the exact results of the Bjorken model [42] with the viscosity \(\eta =10\) MeV/fm\(^2\). The exact solutions for energy density, \(\epsilon \), longitudinal velocity, \(v_z\) and one component of the viscous tensor, \(\pi ^{yy}\) are shown in Fig. 19 by dotted lines for two times elapsed after initialization, \(\Delta t=4\,\mathrm{fm}/c\) and \(8\,\mathrm{fm}/c\). The solid lines show the results for the one-step 3D SHASTA. We see typical increasing oscillation expanding from the boundaries (with the square-root divergent boundary conditions) inwards the small z regions. This is typical behaviour for algorithms using a single-time step approach. Applying Heun’s method, we obtain much smoother behaviour of the solution shown by dashed lines. Although such scheme works well for model tasks, we observe that for actual 3\(+\)1D calculations in the cartesian coordinates the algorithm becomes sometimes unstable (in contrast with the 2\(+\)1D calculation reported in [43]). Also the described entirely-3D approach can lead to problems with anti-diffusion as was noted in [22, 43]. To avoid the instability of the code and other possible troubles, we applied the 3D splitting method for solving the 3D problem.

Fig. 19
figure 19

Comparison of the exact results for energy density, \(\epsilon \), longitudinal velocity, \(v_z\) and one component of the viscous tensor, \(\pi ^{yy}\) of the Bjorken model [42] with the numerical results of the 3\(+\)1D SHASTA code with the one-pass method in the time evolution, Eq. (A4), and of the SHASTA code improved by Heun’s method. Calculations are done for the viscosity \(\eta =10\) MeV/fm\(^2\). The exact solutions are shown two times elapsed after initialization, \(\Delta t=4\,\mathrm{fm}/c\) and \(8\,\mathrm{fm}/c\)

1.2 Implementation of 3D splitting SHASTA for viscous hydrodynamics

We split the 3D task (10) into three sequential 1D propagations

$$\begin{aligned}&\partial _t\mathbf {U}+\partial _x(v_x\mathbf {U})=\mathbf {S}^{\mathrm{1D}}_x \nonumber \\&\partial _t\mathbf {U}+\partial _y(v_y\mathbf {U})=\mathbf {S}^{\mathrm{1D}}_y \nonumber \\&\partial _t\mathbf {U}+\partial _z(v_z\mathbf {U})=\mathbf {S}^{\mathrm{1D}}_z \end{aligned}$$
(A20)

Here, we replace the 3D propagation with three 1D propagations and the source term (A1) is split into three terms corresponding to the propagation along the separate axes. The source terms (A2) and (A3) contain time derivatives and since we replace the 3D propagation by three 1D propagations, we have to include factors 1/3 before the corresponding terms in \(\mathbf {S}^{\mathrm{1D}}_{x,y,z}\). As the result, the final expression for the source term responsible for the propagation along axis \(k=x,y,z\) is

$$\begin{aligned} \mathbf {S}^{1D}_k&=\Big (0, -\frac{1}{3}\,\partial _t\pi ^{00} - \partial _k (Pv_k)- \partial _k\pi ^{0k}, \nonumber \\&\quad -\frac{1}{3}\,\partial _t\pi ^{0x} - \partial _x P - \partial _k\pi ^{kx}, \nonumber \\&\quad -\frac{1}{3}\,\partial _t\pi ^{0y} - \partial _y P - \partial _k\pi ^{ky} \nonumber \\&\quad -\frac{1}{3}\,\partial _t\pi ^{0z} - \partial _z P - \partial _k\pi ^{kz}, \nonumber \\&\qquad Q^{xy}_k, Q^{xz}_k, Q^{yz}_k, Q^{yy}_k, Q^{zz}_k \Big ), \end{aligned}$$
(A21)

where

$$\begin{aligned} Q^{ij}_k&= \pi ^{ij}\left( \partial _kv_k-\frac{1}{3\gamma \tau _\pi }\right) +\frac{\eta }{\gamma \tau _\pi }\,W^{ij}_k , \\ W^{ij}_k&= \frac{2}{3}\,u^iu^j\,\theta _k-u^i{\mathcal {D}}_k u^j-u^j{\mathcal {D}}_k u^i \\&\quad -\delta _{ik}\partial _k u^j-\delta _{jk}\partial _k u^i,\quad i\ne j,\\ W^{ii}_k&= 2\Big \{\frac{1}{3}\left[ 1+(u^i)^2\right] \,\theta _k-u^i{\mathcal {D}}_k u^i-\delta _{ik}\partial _k u^k \Big \}, \end{aligned}$$

and

$$\begin{aligned} \theta _k&\equiv \frac{1}{3}\,\partial _t\gamma +\partial _ku^k, \end{aligned}$$
(A22)
$$\begin{aligned} {\mathcal {D}}_k u^i&=\frac{1}{3}\,\gamma \partial _t u^i+u^k\partial _k u^i. \end{aligned}$$
(A23)

As we see, after summation over index k

$$\begin{aligned} \sum _k\theta _k=\theta ,\,\, \sum _k{\mathcal {D}}_k u^i={\mathcal {D}}u^i,\,\, \sum _k W^{ij}_k=W^{ij}. \end{aligned}$$
(A24)

we recover \(\theta \) and \(W^{\mu \nu }\) defined in Eq. (6), and \({\mathcal {D}}u^i= u_\mu \partial ^\mu u^i\).

Every 1D equation in Eq. (A20) is solved using the standard one-dimensional SHASTA method [24,25,26].

If in some fluid cells the relaxation time becomes smaller than the time step, \(\gamma \tau _\pi <\Delta t\), then we must obtain the formal solution (11),

$$\begin{aligned} \pi ^{ij}(t_{n+1})&=\big [\pi ^{ij}(t_{n})-\eta W^{ij}\big ]e^{-\Delta t/(\gamma \tau _\pi )}+\eta W^{ij} \end{aligned}$$
(A25)

after completion of the propagation along all three axes. To obtain the correct result in the 3D splitting approach, we start with the full solution at the nth time step, \(\pi ^{ij}(t_{n})\), and perform the following sequence of steps in the spatial directions staring,e.g., with the x directions

$$\begin{aligned}{}[\pi ^{ij}(t_{n+1})]_x&= e^{-\Delta t/(\gamma \tau _\pi )} \pi ^{ij}(t_{n})+\big (1-e^{-\Delta t/(\gamma \tau _\pi )}\big )\eta W^{ij}_x, \nonumber \\ [\pi ^{ij}(t_{n+1})]_y&= [\pi ^{ij}(t_{n+1})]_{x}+\big (1-e^{-\Delta t/(\gamma \tau _\pi )}\big )\eta W^{ij}_y, \nonumber \\ [\pi ^{ij}(t_{n+1})]_z&= [\pi ^{ij}(t_{n+1})]_{y}+\big (1-e^{-\Delta t/(\gamma \tau _\pi )}\big )\eta W^{ij}_z, \end{aligned}$$
(A26)

so that after the third step we recover the expected expression.

When one uses 3D splitting method, it is necessary to change the order of 1D propagations to decrease numerical errors. A similar kind of inaccuracy would appear if one uses the same relations among \(\pi ^{ij}\) matrix elements, e.g., \(\pi ^{xx}=\pi ^{xx}(\pi ^{yy},\pi ^{zz})\) permanently. Therefore, we change the independent spatial diagonal components, \(\pi ^{ii}\) and \(\pi ^{jj}\), at every time step, see Eq. (B10).

The results of the application of the 3D splitting scheme for the viscous 2nd-order Bjorken expansion are shown in Fig. 19 by dash-dotted lines. The calculations are performed for the same spatial step, \(\Delta x=0.2\) fm, as used for the 3D SHASTA and Heun’s-improved 3D SHASTA calculations. We see that the numerical results are very close to the theoretical predictions and fluctuations are weaker than for the Heun’s-method improved algorithm. These fluctuations decrease even further if one takes a shorter step. The results of calculations with \(\Delta x=0.1\) fm are shown by dot-dot-dashed lines in Fig. 19. For the most shown quantities, the 3D splitting results are smooth and almost coincide with the exact solutions, only the energy density \(\epsilon \) for the later time, \(\Delta t=8\,\mathrm{fm}/c\), deviates from the exact solution. This deviation vanishes if we go to a smaller step, e.g., \(\Delta x=0.05\) fm, as shown by short dashes. In the actual calculations we have verified on several examples that our results do not change when we reduce the spatial steps from 0.2 fm to 0.1 fm.

Appendix B: Reconstruction of local quantities and exact initialization

The hydrodynamics code evolves the components of the energy-stress tensor and baryon current. The equation of state is formulated in the local system where the energy density and the particle number should be defined. To make a Lorentz transformation from the laboratory frame in the local rest frame one also needs to define a 4-velocity of the fluid element. If we know the components of the ideal stress tensor, \(T_{\mathrm{id}}^{\mu \nu }=T^{\mu \nu }-\pi ^{\mu \nu }\) and the baryon current \(J^\mu =n\, u^\mu \), other quantities can be recovered as follows:

$$\begin{aligned} n&= J^0/\gamma ,\quad \epsilon =T_{\mathrm{id}}^{00}- M\,v, \nonumber \\ M^2&= T^{0x}_{\mathrm{id}}\,T^{0x}_{\mathrm{id}} + T^{0y}_{\mathrm{id}}\,T^{0y}_{\mathrm{id}} + T^{0z}_{\mathrm{id}}\,T^{0z}_{\mathrm{id}}. \end{aligned}$$
(B1)

The modulus of the fluid velocity can be found as a root of the equation

$$\begin{aligned} v=\frac{M}{T^{00}_{\mathrm{id}}-P\big (T^{00}_{\mathrm{id}}-M\, v,J^0/\gamma \big )} \end{aligned}$$
(B2)

and, therefore, depends on the chosen equation of state \(P=P(\epsilon ,n)\). The direction of the fluid velocity is determined as

$$\begin{aligned} v^i=\frac{v}{M}\,T^{0i}_{\mathrm{id}}. \end{aligned}$$
(B3)

Relations (B1), (B2), and (B3) can be used for the ’ideal’ initialization of the hydrodynamic phase when \(\pi ^{\mu \nu }=0\) and \(T^{0\nu }=T^{0\nu }_{\mathrm{id}}\). For the ’exact’ initialization we have to solve the eigenvalue problem \(u_\mu T^{\mu \nu }=\varepsilon u^\nu \), which leads to the quartic algebraic equation

$$\begin{aligned} \varepsilon ^4 + a_1\varepsilon ^3 + a_2\varepsilon ^2 + a_3\varepsilon + a_4=0, \end{aligned}$$
(B4)

where coefficients \(a_{1,2,3,4}\) are functions of the energy–momentum tensor invariants,

$$\begin{aligned} a_1&=-T^\nu _\nu =-\text {Tr}\,T^\mu _\nu ,\quad a_2=\frac{1}{2}\left( a_1^2-T^\mu _\nu T_\mu ^\nu \right) , \nonumber \\ a_3&=a_1a_2-\frac{a_1^3+T^\mu _\nu T^\nu _\lambda T^\lambda _\mu }{3}, \nonumber \\ a_4&=-\det T^{\mu \nu }=-\varepsilon ^{\kappa \lambda \mu \nu }T^0_\kappa T^1_\lambda T^2_\mu T^3_\nu . \end{aligned}$$
(B5)

The corresponding velocity is calculated as

$$\begin{aligned} v_z&=\frac{T^{03}(\varepsilon +T^{11})(\varepsilon +T^{22}) -T^{02}T^{23}(\varepsilon +T^{11})-T^{01}T^{13}(\varepsilon +T^{22})+T^{01}T^{12}T^{23}+T^{02}T^{12}T^{13}-T^{03}(T^{12})^2}{(\varepsilon +T^{33})(\varepsilon +T^{22})(\varepsilon +T^{11})-(T^{12})^2(\varepsilon +T^{33}) -(T^{13})^2(\varepsilon +T^{22})-(T^{23})^2(\varepsilon +T^{11})+2T^{12}T^{13}T^{23}}, \nonumber \\ v_y&=\frac{T^{02}(\varepsilon +T^{11})-T^{01}T^{12} +\left[ T^{12}T^{13}-T^{23}(\varepsilon +T^{11})\right] v_z}{(\varepsilon +T^{22})(\varepsilon +T^{11})-(T^{12})^2}, \qquad \nonumber \\ v_x&=\frac{T^{01}-T^{12}v_y-T^{13}v_z}{\varepsilon +T^{11}}. \end{aligned}$$
(B6)

Given the four velocity, \(u^\mu =\gamma (1,\mathbf {v})\), we calculate baryon density as

$$\begin{aligned} n&=J^\nu u_\nu , \end{aligned}$$
(B7)

Then knowing the equation of state \(P=P(\varepsilon ,n)\), we can define the ideal part of the energy–momentum tensor \(T_{\mathrm{id}}^{\mu \nu }\) in Eq. (3). The viscous parts of the full \(T^{\mu \nu }\) tensor (2) follow as

$$\begin{aligned} \Pi&=\frac{1}{3}\,\left( \varepsilon -T^\nu _\nu \right) -P, \nonumber \\ \pi ^{\mu \nu }&=T^{\mu \nu }-(\varepsilon +P+\Pi )u^\mu u^\nu +(P+\Pi )g^{\mu \nu } , \end{aligned}$$
(B8)

and the baryon diffusion current as

$$\begin{aligned} V^\mu&=J^\mu -n u^\mu . \end{aligned}$$
(B9)

In the code we use quantities \(\pi ^{xy}\), \(\pi ^{xz}\), \(\pi ^{yz}\), \(\pi ^{ii}\), and \(\pi ^{jj}\) as independent variables. Other components can be recovered with the help of the following expressions

$$\begin{aligned} \pi ^{0i}&= \pi ^{ik} v_k+\pi ^{ii} v_i+\pi ^{ij} v_j, \nonumber \\ \pi ^{0j}&= \pi ^{jk} v_k+\pi ^{ij} v_i+\pi ^{jj} v_j, \nonumber \\ \pi ^{kk}&= \frac{1}{1-v_k^2} \Big [\pi ^{0i} v_i+\pi ^{0j}v_j+(\pi ^{ik} v_i +\pi ^{jk} v_j )v_k \nonumber \\&\quad -(\pi ^{ii} +\pi ^{jj}) \Big ],\nonumber \\ \pi ^{0k}&= \pi ^{kk}v_k+\pi ^{ik}v_i+\pi ^{jk}v_j, \nonumber \\ \pi ^{00}&=\pi ^{ii}+\pi ^{jj}+\pi ^{kk}. \end{aligned}$$
(B10)

We emphasize that these expressions do not develop anomalously large values for the case of small fluid velocities.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Khvorostukhin, A.S., Kolomeitsev, E.E. & Toneev, V.D. Hybrid model with viscous relativistic hydrodynamics: a role of constraints on the shear-stress tensor. Eur. Phys. J. A 57, 294 (2021). https://doi.org/10.1140/epja/s10050-021-00599-1

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epja/s10050-021-00599-1

Keywords

Navigation