Skip to main content
Log in

An extended hydrodynamics model for inertial confinement fusion hohlraums

  • Regular Article – Plasma Physics
  • Published:
The European Physical Journal D Aims and scope Submit manuscript

Abstract

In some inertial confinement fusion hohlraum designs, the inside plasma is not sufficiently collisional to be satisfactorily described by the Euler equations implemented in hydrodynamic simulation codes, particularly in converging regions of the expanding plasma flow. To better treat that situation, this paper presents an extended hydrodynamics model including higher moments of the particle velocity distribution function, together with physically justified closure assumptions and relaxation terms. A preliminary one-dimensional numerical implementation of the model is shown to give satisfactory results in a test case involving a high-velocity collision of two plasma flows. Paths to extend that model to three dimensions as needed for an actual hohlraum geometry are briefly discussed.

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

Similar content being viewed by others

Data Availability Statement

This manuscript has data included as electronic supplementary material.

References

  1. J.D. Lindl, Inertial Confinement Fusion - The Quest for Ignition and Energy Gain Using Indirect Drive (Springer-Verlag, New York, 1998)

    Google Scholar 

  2. J.D. Lindl, P. Amendt, R.L. Berger, S.G. Glendinning, S.H. Glenzer, S.W. Haan, R.L. Kauffman, O.L. Landen, L.J. Suter, The physics basis for ignition using indirect-drive targets on the National Ignition Facility. Phys. Plasmas 11, 339 (2004). https://doi.org/10.1063/1.1578638

    Article  ADS  Google Scholar 

  3. S. Atzeni, J. Meyer-ter-Vehn, The Physics of Inertial Fusion (Clarendon Press, Oxford, 2004)

    Book  Google Scholar 

  4. L.F. Berzak Hopkins, S. Le Pape, L. Divol, N.B. Meezan, A.J. Mackinnon, D.D. Ho, O.S. Jones, S. Khan, J.L. Milovich, J.S. Ross, P. Amendt, D. Casey, P.M. Celliers, A. Pak, J.L. Peterson, J. Ralph, J.R. Rygg, Near-vacuum hohlraums for driving fusion implosions with high density carbon ablators. Phys. Plasmas 22, 056318 (2015). https://doi.org/10.1063/1.4921151

    Article  ADS  Google Scholar 

  5. S.H. Glenzer, W.E. Alley, K.G. Estabrook, J.S. De Groot, M.G. Haines, J.H. Hammer, J.-P. Jadaud, B.J. MacGowan, J.D. Moody, W. Rozmus, L.J. Suter, T.L. Weiland, E.A. Williams, Thomson scattering from laser plasmas. Phys. Plasmas 6, 2117 (1999). https://doi.org/10.1063/1.873499

    Article  ADS  Google Scholar 

  6. E. Dattolo, L. Suter, M.-C. Monteil, J.-P. Jadaud, N. Dague, S. Glenzer, R. Turner, D. Juraszek, B. Lasinski, C. Decker, O. Landen, B. MacGowan, Status of our understanding and modeling of x-ray coupling efficiency in laser heated hohlraums. Phys. Plasmas 8, 260 (2001). https://doi.org/10.1063/1.1324659

    Article  ADS  Google Scholar 

  7. C.A. Back, J. Davis, J. Grun, L.J. Suter, O.L. Landen, W.W. Hsing, M.C. Miller, Multi-keV x-ray conversion efficiency in laser-produced plasmas. Phys. Plasmas 10, 2047 (2003). https://doi.org/10.1063/1.1566750

    Article  ADS  Google Scholar 

  8. D.P. Higginson, D. Bailey, N. Meezan, D. Strozzi, S. Wilks, G. Zimmerman, Impact of Multi-species & CBET in Near Vacuum Hohlraum Simulations, in preparation (2021)

  9. C. Chenais-Popovics, P. Renaudin, O. Rancu, F. Gilleron, J.-C. Gauthier, O. Larroche, O. Peyrusse, M. Dirksmöller, P. Sondhauss, T. Missalla, I. Uschmann, E. Förster, O. Renner, E. Krousky, Kinetic to thermal energy transfer and interpenetration in the collision of laser-produced plasmas. Phys. Plasmas 4, 190 (1997). https://doi.org/10.1063/1.872132

    Article  ADS  Google Scholar 

  10. A.S. Wan, T.W. Barbee Jr., R. Cauble, P. Celliers, L.B. Da Silva, J.C. Moreno, P.W. Rambo, G.F. Stone, J.E. Trebes, F. Weber, Electron density measurement of a colliding plasma using soft-x-ray laser interferometry. Phys. Rev. E 55, 6293 (1997). https://doi.org/10.1103/PhysRevE.55.6293

    Article  ADS  Google Scholar 

  11. D.R. Farley, K.G. Estabrook, S.G. Glendinning, S.H. Glenzer, B.A. Remington, K. Shigemori, J.M. Stone, R.J. Wallace, G.B. Zimmerman, J.A. Harte, Radiative jet experiments of astrophysical interest using intense lasers. Phys. Rev. Lett. 83, 1982 (1999). https://doi.org/10.1103/PhysRevLett.83.1982

    Article  ADS  Google Scholar 

  12. O. Renner, M. Šmíd, T. Burian, L. Juha, J. Krása, E. Krouský, I. Matulková, J. Skála, A. Velyhan, R. Liska, J. Velechovský, T. Pisarczyk, T. Chodukowski, O. Larroche, J. Ullschmied, Environmental conditions in near-wall plasmas generated by impact of energetic particle fluxes. High Energy Density Phys. 9, 568 (2013). https://doi.org/10.1016/j.hedp.2013.05.012

    Article  ADS  Google Scholar 

  13. C. Fallon, P. Hayden, N. Walsh, E.T. Kennedy, J.T. Costello, Target geometrical effects on the stagnation layer formed by colliding a pair of laser produced copper plasmas. Phys. Plasmas 22, 093506 (2015). https://doi.org/10.1063/1.4930204

    Article  ADS  Google Scholar 

  14. H.G. Rinderknecht, H.-S. Park, J.S. Ross, P.A. Amendt, D.P. Higginson, S.C. Wilks, D. Haberberger, J. Katz, D.H. Froula, N.M. Hoffman, G. Kagan, B.D. Keenan, E.L. Vold, Highly resolved measurements of a developing strong collisional plasma shock. Phys. Rev. Lett. 120, 095001 (2018). https://doi.org/10.1103/PhysRevLett.120.095001

    Article  ADS  Google Scholar 

  15. S. Le Pape, L. Divol, G. Huser, J. Katz, A. Kemp, J.S. Ross, R. Wallace, S. Wilks, Plasma collision in a gas atmosphere. Phys. Rev. Lett. 124, 025003 (2020). https://doi.org/10.1103/PhysRevLett.124.025003

  16. P.W. Rambo, Kinetic MC-PIC Simulations of Axially Stagnating Plasma, paper 8Q25, 36th APS-DPP meeting, Minneapolis, MN, November 7-11, (1994)

  17. M.J. Rosenberg, H.G. Rinderknecht, N.M. Hoffman, P.A. Amendt, S. Atzeni, A.B. Zylstra, C.K. Li, F.H. Séguin, H. Sio, M Gatu Johnson, J.A. Frenje, R.D. Petrasso, V Yu. Glebov, C. Stoeckl, W. Seka, F.J. Marshall, J.A. Delettrez, T.C. Sangster, R. Betti, V.N. Goncharov, D.D. Meyerhofer, S. Skupsky, C. Bellei, J. Pino, S.C. Wilks, G. Kagan, K. Molvig, A. Nikroo, Exploration of the transition from the hydrodynamiclike to the strongly kinetic regime in shock-driven implosions. Phys. Rev. Lett. 112, 185001 (2014). https://doi.org/10.1103/PhysRevLett.112.185001

    Article  ADS  Google Scholar 

  18. H. Sio, O. Larroche, S. Atzeni, N.V. Kabadi, J.A. Frenje, M. Gatu Johnson, C. Stoeckl, C.K. Li, C.J. Forrest, V. Glebov, P.J. Adrian, A. Bose, A. Birkel, S.P. Regan, F.H. Seguin, R.D. Petrasso, Probing ion species separation and ion thermal decoupling in shock-driven implosions using multiple nuclear reaction histories. Phys. Plasmas 26, 072703 (2019). https://doi.org/10.1063/1.5097605

  19. O. Larroche, H.G. Rinderknecht, M.J. Rosenberg, N.M. Hoffman, S. Atzeni, R.D. Petrasso, P.A. Amendt, F.H. Séguin, Ion-kinetic simulations of D-\(^3\)He gas-filled ICF target implosions with moderate to large Knudsen number. Phys. Plasmas 23, 012701 (2016). https://doi.org/10.1063/1.4939025

    Article  ADS  Google Scholar 

  20. O. Larroche, H.G. Rinderknecht, M.J. Rosenberg, Nuclear yield reduction in inertial confinement fusion exploding-pusher targets explained by fuel-pusher mixing through hybrid kinetic-fluid modeling. Phys. Rev. E 98, 031201 (2018). https://doi.org/10.1103/PhysRevE.98.031201

    Article  ADS  Google Scholar 

  21. O. Larroche, Kinetic simulation of a plasma collision experiment. Phys. Fluids B 5, 2816 (1993). https://doi.org/10.1063/1.860670

    Article  ADS  Google Scholar 

  22. W.T. Taitano, L. Chacón, A.N. Simakov, S.E. Anderson, A conservative phase-space moving-grid strategy for a 1D–2V Vlasov–Fokker–Planck solver. Comput. Phys. Commun. 258, 107547 (2021). https://doi.org/10.1016/j.cpc.2020.107547

    Article  MathSciNet  Google Scholar 

  23. P.W. Rambo, R.J. Procassini, A comparison of kinetic and multifluid simulations of laser-produced colliding plasmas. Phys. Plasmas 2, 3130 (1995). https://doi.org/10.1063/1.871145

    Article  ADS  Google Scholar 

  24. T.D. Arber, K. Bennett, C.S. Brady, A. Lawrence-Douglas, M.G. Ramsay, N.J. Sircombe, P. Gillies, R.G. Evans, H. Schmitz, A.R. Bell, C.P. Ridgers, Contemporary particle-in-cell approach to laser-plasma modelling. Plasma Phys. Control. Fusion 57, 113001 (2015). https://doi.org/10.1088/0741-3335/57/11/113001

    Article  ADS  Google Scholar 

  25. R. Bird, N. Tan, S. V. Luedtke, S. L. Harrell, M. Taufer, B. Albright, VPIC 2.0: Next generation particle-in-cell simulations, IEEE Transactions on Parallel and Distributed Systems, in press (2021) https://doi.org/10.1109/TPDS.2021.3084795

  26. O. Larroche, Kinetic simulations of fuel ion transport in ICF target implosions. Eur. Phys. J. D 27, 131 (2003). https://doi.org/10.1140/epjd/e2003-00251-1

    Article  ADS  Google Scholar 

  27. C. Bellei, H. Rinderknecht, A. Zylstra, M. Rosenberg, H. Sio, C.K. Li, R. Petrasso, S.C. Wilks, P.A. Amendt, Species separation and kinetic effects in collisional plasma shocks. Phys. Plasmas 21, 056310 (2014). https://doi.org/10.1063/1.4876614

    Article  ADS  Google Scholar 

  28. B.E. Peigney, O. Larroche, V. Tikhonchuk, Ion kinetic effects on the ignition and burn of inertial confinement fusion targets: a multi-scale approach. Phys. Plasmas 21, 122709 (2014). https://doi.org/10.1063/1.4904212

    Article  ADS  Google Scholar 

  29. I. Sagert, W. Bauer, D. Colbry, J. Howell, R. Pickett, A. Staber, T. Strother, Hydrodynamic shock wave studies within a kinetic Monte Carlo approach. J. Comput. Phys 266, 191 (2014). https://doi.org/10.1016/j.jcp.2014.02.019

    Article  ADS  MathSciNet  MATH  Google Scholar 

  30. R.L. Berger, J.R. Albritton, C.J. Randall, E.A. Williams, W.L. Kruer, A.B. Langdon, C.J. Hanna, Stopping and thermalization of interpenetrating plasma streams. Phys. Fluids B 3, 3 (1991). https://doi.org/10.1063/1.859954

    Article  ADS  Google Scholar 

  31. P.W. Rambo, J. Denavit, Interpenetration and ion separation in colliding plasmas. Phys. Plasmas 1, 4050 (1994). https://doi.org/10.1063/1.870875

    Article  ADS  Google Scholar 

  32. D. Ghosh, T.D. Chapman, R.L. Berger, A. Dimits, J.W. Banks, A multispecies, multifluid model for laser-induced counterstreaming plasma simulations. Comput. Fluids 186, 38 (2019). https://doi.org/10.1016/j.compfluid.2019.04.012

    Article  MathSciNet  MATH  Google Scholar 

  33. M. Marciante, C. Enaux, The hydrodynamics of lerna, submitted to J. Comput. Phys. (2021) hal-03335437

  34. G.B. Zimmerman, W.L. Kruer, Numerical simulation of laser-initiated fusion. Comments Plasma Phys. Control. Fusion 2, 51 (1975)

    Google Scholar 

  35. C.H. Chang, A.K. Stagg, A compatible Lagrangian hydrodynamic scheme for multicomponent flows with mixing. J. Comput. Phys. 231, 4279 (2012). https://doi.org/10.1016/j.jcp.2012.02.005

    Article  ADS  MathSciNet  MATH  Google Scholar 

  36. H. Grad, On the kinetic theory of rarefied gases. Comm. Pure Appl. Math. 2, 331 (1949). https://doi.org/10.1002/cpa.3160020403

    Article  MathSciNet  MATH  Google Scholar 

  37. C.D. Levermore, Moment closure hierarchies for kinetic theories. J. Stat. Phys. 83, 1021 (1996). https://doi.org/10.1007/BF02179552

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. J.D. Au, M. Torrilhon, W. Weiss, The shock tube study in extended thermodynamics. Phys. Fluids 13, 2423 (2001). https://doi.org/10.1063/1.1381018

    Article  ADS  MATH  Google Scholar 

  39. F. Forgues, J.G. McDonald, Higher-order moment models for laminar multiphase flows with accurate particle-stream crossing. Int. J. Multiphase Flow 114, 28 (2019). https://doi.org/10.1016/j.ijmultiphaseflow.2019.01.003

    Article  MathSciNet  Google Scholar 

  40. C. Baranger, A. Burbeau-Augoula, P. Seytor, P. Hoch, O. Larroche, J. Métral, B. Rebourcet, Numerical modeling of a self-colliding plasma. Int. J. Numer. Meth. Fluids 65, 1451 (2011). https://doi.org/10.1002/fld.2375

    Article  MathSciNet  MATH  Google Scholar 

  41. M.N. Rosenbluth, W.M. MacDonald, D.L. Judd, Fokker–Planck equation for an inverse-square force. Phys. Rev. 107, 1 (1957). https://doi.org/10.1103/PhysRev.107.1

    Article  ADS  MathSciNet  MATH  Google Scholar 

  42. M. Torrilhon, Characteristic waves and dissipation in the 13-moment case. Continuum. Mech. Thermodyn. 12, 289 (2000). https://doi.org/10.1007/s001610050138

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. R.B. Larson, A method for computing the evolution of star clusters. Mon. Not. R. Astron. Soc. 147, 323 (1970). https://doi.org/10.1093/mnras/147.4.323

    Article  ADS  Google Scholar 

  44. S. Cuperman, I. Weiss, M. Dryer, Higher order fluid equations for multicomponent nonequilibrium stellar (plasma) atmospheres and star clusters. II. Effects of nonzero relative flow velocities and skewing of velocity distribution functions. Astrophys. J. 251, 297 (1981). https://doi.org/10.1086/159465

    Article  ADS  Google Scholar 

  45. M.M. Echim, J. Lemaire, Ø. Lie-Svendsen, A review on solar wind modeling: kinetic and fluid aspects. Surv. Geophys. 32, 1 (2011). https://doi.org/10.1007/s10712-010-9106-y

    Article  ADS  Google Scholar 

  46. J. Ng, A. Hakim, L. Wang, A. Bhattacharjee, An improved ten-moment closure for reconnection and instabilities. Phys. Plasmas 27, 082106 (2020). https://doi.org/10.1063/5.0012067

    Article  ADS  Google Scholar 

  47. N. Böhmer, M. Torrilhon, Entropic quadrature for moment approximations of the Boltzmann-BGK equation. J. Comput. Phys. 401, 108992 (2020). https://doi.org/10.1016/j.jcp.2019.108992

    Article  MathSciNet  MATH  Google Scholar 

  48. M. Torrilhon, Hyperbolic moment equations in kinetic gas theory based on multi-variate Pearson-IV-distributions. Commun. Comput. Phys. 7, 639 (2010). https://doi.org/10.4208/cicp.2009.09.049

    Article  MathSciNet  MATH  Google Scholar 

  49. J. Hamilton, C.E. Seyler, Formulation of 8-moment plasma transport with application to the Nernst effect. Phys. Plasmas 28, 022306 (2021). https://doi.org/10.1063/5.0030117

    Article  ADS  Google Scholar 

  50. S.T. Miller, U. Shumlak, A multi-species 13-moment model for moderately collisional plasmas. Phys. Plasmas 23, 082303 (2016). https://doi.org/10.1063/1.4960041

    Article  ADS  Google Scholar 

  51. P.L. Bhatnagar, E.P. Gross, M. Krook, A model for collision processes in gases. I. Small amplitude processes in charged and neutral one-component systems. Phys. Rev. 94, 511 (1954). https://doi.org/10.1103/PhysRev.94.511

    Article  ADS  MATH  Google Scholar 

  52. F.J. McCormack, Kinetic equations for polyatomic gases: the 17-moment approximation. Phys. Fluids 11, 2533 (1968). https://doi.org/10.1063/1.1691855

    Article  ADS  Google Scholar 

  53. D. Jou, V. Micenmacher, Extended thermodynamics of viscous phenomena in real gases. J. Phys. A: Math. Gen. 20, 6519 (1987). https://doi.org/10.1088/0305-4470/20/18/048

    Article  ADS  Google Scholar 

  54. L.J. Stanek, M.S. Murillo, Analytic models for interdiffusion in dense plasma mixtures. Phys. Plasmas 28, 072302 (2021). https://doi.org/10.1063/5.0047961

    Article  ADS  Google Scholar 

  55. H. Struchtrup, Stable transport equations for rarefied gases at high orders in the Knudsen number. Phys. Fluids 16, 3921 (2004). https://doi.org/10.1063/1.1782751

    Article  ADS  MathSciNet  MATH  Google Scholar 

  56. H. Struchtrup, M. Torrilhon, Regularization of Grad’s 13 moment equations: Derivation and linear analysis. Phys. Fluids 15, 2668 (2003). https://doi.org/10.1063/1.1597472

  57. M. Yu. Timokhin, Ye. A. Bondar, A. A. Kokhanchik, M. S. Ivanov, I. E. Ivanov, I. A. Kryukov, Study of the shock wave structure by regularized Grad’s set of equations. Phys. Fluids 27, 037101 (2015) https://doi.org/10.1063/1.4913673

  58. G.V. Candler, S. Nijhawan, D. Bose, I.D. Boyd, A multiple translational temperature gas dynamics model. Phys. Fluids 6, 3776 (1994). https://doi.org/10.1063/1.868367

    Article  ADS  MATH  Google Scholar 

  59. Kun Xu, E. Josyula, Multiple translational temperature model and its shock structure solution. Phys. Rev. E 71, 056308 (2005). https://doi.org/10.1103/PhysRevE.71.056308

  60. L.H. Holway Jr., New statistical models for kinetic theory: methods of construction. Phys. Fluids 9, 1658 (1966). https://doi.org/10.1063/1.1761920

    Article  ADS  Google Scholar 

  61. S. I. Braginskii, Transport Processes in a Plasma, Reviews of Plasma Physics - Volume 1, ed. by M. A. Leontovich (Consultants Bureau, New York, 1965), pp. 205–311

  62. A. Decoster, Fluid Equations and Transport Coefficients of Plasmas, Modeling of Collisions, ed. by P. A. Raviart, Gauthier-Villars, Paris (1998), pp. 1–137

  63. H.M. Mott-Smith, The solution of the Boltzmann equation for a shock wave. Phys. Rev. 82, 885 (1951). https://doi.org/10.1103/PhysRev.82.885

    Article  ADS  MathSciNet  MATH  Google Scholar 

  64. K. Abe, G. Sakaguchi, Linear and nonlinear evolution of double-humped ion distributions in strong unmagnetized shock structures. Phys. Fluids 28, 3581 (1985). https://doi.org/10.1063/1.865313

    Article  ADS  MathSciNet  MATH  Google Scholar 

  65. M. Casanova, O. Larroche, J.-P. Matte, Kinetic simulation of a collisional shock wave in a plasma. Phys. Rev. Lett. 67, 2143 (1991). https://doi.org/10.1103/PhysRevLett.67.2143

    Article  ADS  Google Scholar 

  66. F. Vidal, J.-P. Matte, M. Casanova, O. Larroche, Modeling and effects of nonlocal electron heat flow in planar shock waves. Phys. Plasmas 2, 1412 (1995). https://doi.org/10.1063/1.871357

    Article  ADS  Google Scholar 

  67. B. Perthame, Second-order Boltzmann schemes for compressible Euler equations in one and two space dimensions. SIAM J. Numer. Anal. 29, 1 (1992). https://doi.org/10.1137/0729001

    Article  MathSciNet  MATH  Google Scholar 

  68. J.L. Estivalezes, P. Villedieu, High-order positivity-preserving kinetic schemes for the compressible Euler equations. SIAM J. Numer. Anal. 33, 2050 (1996). https://doi.org/10.1137/S0036142994271009

    Article  MathSciNet  MATH  Google Scholar 

  69. V.I. Kogan, The rate of equalization of the temperatures of charged particles in a plasma, in Plasma physics and the problem of controlled thermonuclear reactions, vol. 1, ed. by M.A. Leontovich (Pergamon Press, Oxford, 1961), pp. 153–161

    Google Scholar 

  70. H. Schamel, H. Hamnén, D.F. Düchs, T.E. Stringer, M.R. O’Brien, Nonlinear analysis of Coulomb relaxation of anisotropic distributions. Phys. Fluids B 1, 76 (1989). https://doi.org/10.1063/1.859108

  71. D.V. Sivukhin, Coulomb collisions in a fully ionized plasma, in Reviews of Plasma Physics - Volume 4. ed. by M.A. Leontovich (Consultants Bureau, New York, 1966), p. 93

    Google Scholar 

  72. J.F. Luciani, P. Mora, R. Pellat, Quasistatic heat front and delocalized heat flux. Phys. Fluids 28, 835 (1985). https://doi.org/10.1063/1.865052

    Article  ADS  MATH  Google Scholar 

  73. A. Kurganov, Central Schemes: A Powerful Black-Box Solver for Nonlinear Hyperbolic PDEs, Handbook of Numerical Methods for Hyperbolic Problems: Basic and Fundamental Issues, Edited by Rémi Abgrall, Chi-Wang Shu, Handbook of Numerical Analysis Vol. 17, Chap. 20, pp. 525-548, North Holland publishing, Elsevier (2016) https://doi.org/10.1016/bs.hna.2016.09.008

  74. A. Kurganov, Chi-Tien Lin, On the reduction of numerical dissipation in central-upwind schemes. Commun. Comput. Phys. 2, 141 (2007)

  75. J.P. Boris, D.L. Book, Flux-corrected transport. I. SHASTA, a fluid transport algorithm that works. J. Comput. Phys. 11, 38 (1973). https://doi.org/10.1016/0021-9991(73)90147-2

    Article  ADS  MATH  Google Scholar 

  76. Flux-Corrected Transport. Principles, Algorithms and Applications, ed. by D. Kuzmin, R. Löhner and S. Turek, Second edition, Springer Verlag, Berlin, Heidelberg (2012) https://doi.org/10.1007/978-94-007-4038-9

  77. W.B. VanderHeyden, B.A. Kashiwa, Compatible fluxes for van Leer advection. J. Comput. Phys. 146, 1 (1998). https://doi.org/10.1006/jcph.1998.6070

    Article  ADS  MathSciNet  MATH  Google Scholar 

  78. R. Liska, M. Shashkov, P. Váchal, B. Wendroff, Synchronized flux corrected remapping for ALE methods. Comput. Fluids 46, 312 (2011). https://doi.org/10.1016/j.compfluid.2010.11.013

    Article  MathSciNet  MATH  Google Scholar 

  79. O. Larroche, An efficient explicit numerical scheme for diffusion-type equations with a highly inhomogeneous and highly anisotropic diffusion tensor. J. Comput. Phys. 223, 436 (2007). https://doi.org/10.1016/j.jcp.2006.09.016

  80. B.E. Peigney, O. Larroche, V. Tikhonchuk, Fokker-Planck kinetic modeling of suprathermal \(\alpha \) particles in a fusion plasma. J. Comput. Phys. 278, 416 (2014). https://doi.org/10.1016/j.jcp.2014.08.033

    Article  ADS  MathSciNet  MATH  Google Scholar 

  81. V. Vikas, Z.J. Wang, A. Passalacqua, R.O. Fox, Realizable high-order finite-volume schemes for quadrature-based moment methods. J. Comput. Phys. 230, 5328 (2011). https://doi.org/10.1016/j.jcp.2011.03.038

    Article  ADS  MathSciNet  MATH  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to O. Larroche.

Supplementary Information

Below is the link to the electronic supplementary material.

Supplementary material 1 (jpg 324 KB)

Appendices

Appendix A: Fokker–Planck relaxation times for Coulomb collisions

The expressions of the collision times used in the main part of the paper are derived here from the Fokker–Planck equation governing Coulomb collision processes, in the reference form of Rosenbluth et al. [41]. The term governing the evolution of the velocity distribution function \(f_a(\mathbf {c})\) for particles of species a due to collisions on particles of species b is:

$$\begin{aligned} \left( \frac{\partial f_a}{\partial t}\right) _{a\rightarrow b} = -\frac{\partial J_{(ab)i}}{\partial c_i} \end{aligned}$$
(A.1)

where the current in velocity space \({\mathbf {J}}_{(ab)}\) is the sum of a convection term and a diffusion term:

$$\begin{aligned} J_{(ab)i} = 4\pi \varGamma _{ab}\left( -\frac{m_a}{m_b}\frac{\partial {\mathcal {S}}_b}{\partial c_i}f_a + \frac{\partial ^2{\mathcal {T}}_b}{\partial c_i\partial c_j}\frac{\partial f_a}{\partial c_j}\right) \end{aligned}$$
(A.2)

with the Rosenbluth potentials \({\mathcal {S}}_b\) and \({\mathcal {T}}_b\) defined by:

$$\begin{aligned} \varDelta _c{\mathcal {S}}_b = f_b \quad ; \quad \varDelta _c{\mathcal {T}}_b = {\mathcal {S}}_b \end{aligned}$$
(A.3)

and with

$$\begin{aligned} \varGamma _{ab} = \frac{4\pi e^4Z_a^2Z_b^2}{m_a^2}\mathrm {Log}\varLambda _{ab} \end{aligned}$$

where \(Z_se\) and \(m_s\) are the charge and mass of particles of species \(s=a,b\) and \(\mathrm {Log}\varLambda _{ab}\) is the Coulomb logarithm defined, e.g., in Ref. [71]. We use a definition of the potentials slightly different from the original \(h_b\) and \(g_b\) from reference [41]; the correspondence is:

$$\begin{aligned} {\mathcal {S}}_b = -\frac{m_b}{m_a+m_b}\frac{h_b}{4\pi } \quad ; \quad {\mathcal {T}}_b = -\frac{g_b}{8\pi } \end{aligned}$$

A general integral expression of the Rosenbluth potentials, which is a solution of the Poisson equations (A.3), is

$$\begin{aligned} S_b(\mathbf {c})&= \frac{-1}{4\pi }\int \frac{f_b(\mathbf {c}^\prime )}{|\mathbf {c}-\mathbf {c}^\prime |}\text{ d}^3c^\prime \end{aligned}$$
(A.4)
$$\begin{aligned} T_b(\mathbf {c})&= \frac{-1}{8\pi }\int f_b(\mathbf {c}^\prime )|\mathbf {c}-\mathbf {c}^\prime |\text{ d}^3c^\prime \end{aligned}$$
(A.5)

More useful expressions can be obtained in specific cases.

The case of a Maxwellian distribution for target particles. If \(f_b\) is the Maxwellian:

$$\begin{aligned} f_b(\mathbf {c}) = n_b\left( \frac{m_b}{2\pi k_BT_b}\right) ^{3/2}e^{-u^2} \end{aligned}$$

with

$$\begin{aligned} \mathbf {u} = \left( \frac{m_b}{2k_BT_b}\right) ^{1/2}(\mathbf {c}-\mathbf {v}_b) \end{aligned}$$

where \(n_b\), \(\mathbf {v}_b\) and \(T_b\) are the density, bulk velocity and temperature of species b, then the Rosenbluth potentials can be explicitly computed:

$$\begin{aligned} {\mathcal {S}}_b(\mathbf {c})&= -\frac{n_b}{4\pi }\left( \frac{m_b}{2k_BT_b}\right) ^{1/2}\frac{\text{ erf }(u)}{u} \\ {\mathcal {T}}_b(\mathbf {c})&= -\frac{n_b}{8\pi }\left( \frac{2k_BT_b}{m_b}\right) ^{1/2}\left( \frac{e^{-u^2}}{\sqrt{\pi }} + \left( u+\frac{1}{2u}\right) \text{ erf }(u)\right) \end{aligned}$$

For electrons, to lowest order in powers of the electron/ion mass ratio, we get:

$$\begin{aligned} \frac{\partial {\mathcal {S}}_e}{\partial c_i}&= \frac{n_e}{3\pi ^{3/2}}\left( \frac{m_e}{2k_BT_e}\right) ^{3/2}(c_i-v_{e,i}) \\ \frac{\partial ^2{\mathcal {T}}_e}{\partial c_i\partial c_j}&= -\frac{n_e}{6\pi ^{3/2}}\left( \frac{m_e}{2k_BT_e}\right) ^{1/2}\delta _{ij} \end{aligned}$$

so that the electron collision term for ions of species a finally reads:

$$\begin{aligned} \left( \frac{\partial f_a}{\partial t}\right) _{ae}&= \frac{4\sqrt{2\pi }e^4Z_a^2m_e^{1/2}n_e}{3m_a(k_BT_e)^{3/2}}\mathrm {Log}\varLambda _{ae} \\&\quad \times \frac{\partial }{\partial c_i}\left( (c_i-v_{e,i})f_a + \frac{k_BT_e}{m_a}\frac{\partial f_a}{\partial c_i}\right) \end{aligned}$$

For collisions on ions, we get:

$$\begin{aligned} \frac{\partial {\mathcal {S}}_b}{\partial c_i}&= \frac{n_b}{3\pi ^{3/2}}\left( \frac{m_b}{2k_BT_b}\right) ^{3/2}R(u)(c_i-v_{b,i})\\ \frac{\partial ^2{\mathcal {T}}_b}{\partial c_i\partial c_j}&= -\frac{n_b}{6\pi ^{3/2}}\left( \frac{m_b}{2k_BT_b}\right) ^{1/2} \nonumber \\&\quad \times \left[ \left( \delta _{ij}-\frac{u_iu_j}{u^2}\right) L(u) + \frac{u_iu_j}{u^2}R(u)\right] \nonumber \end{aligned}$$
(A.6)

where the following functions have been defined:

$$\begin{aligned} R(u)&= \frac{3}{2u^2}\left( \frac{\sqrt{\pi }\mathrm {erf}(u)}{2u}-e^{-u^2}\right) \\&\mathrel {\mathop {\sim }\limits _{0}} 1-\frac{3u^2}{5} \dots \\&\mathrel {\mathop {\sim }\limits _{\infty }} \frac{3\sqrt{\pi }}{4u^3} \\ L(u)&= \frac{3}{4u^2}\left( e^{-u^2}+\left( 2u-\frac{1}{u}\right) \frac{\sqrt{\pi }}{2}\mathrm {erf}(u)\right) \\&\mathrel {\mathop {\sim }\limits _{0}} 1-\frac{u^2}{5} \dots \\&\mathrel {\mathop {\sim }\limits _{\infty }} \frac{3\sqrt{\pi }}{4u} \end{aligned}$$

To get synthetic formulas for the orders of magnitude of collision times, we will use a more practical approximation of R(u), which reproduces those two limits:

$$\begin{aligned} R(u) \rightarrow \left( 1+\left( \frac{2}{9\pi }\right) ^{1/3}2u^2\right) ^{-3/2} \end{aligned}$$

Definition of a slowing-down time From the expression (A.2) of the current in velocity space, the slowing-down rate of distribution a by target particles b is obtained:

$$\begin{aligned} \frac{{\mathrm{d}}v_{a,i}}{{\mathrm{d}}t}&= \frac{1}{n_a}\int c_i \left( \frac{\partial f_a}{\partial t}\right) _{a\rightarrow b}d^3c \nonumber \\&= -\frac{4\pi \varGamma _{ab}}{n_a}\left( 1+\frac{m_a}{m_b}\right) \int \frac{\partial {\mathcal {S}}_b}{\partial c_i}f_a({\mathbf {c}})d^3c \end{aligned}$$
(A.7)

Among the two terms in the latter expression, the second one comes directly from the “slowing-down” term in the Fokker–Planck equation, and the first one comes from the variation of the diffusion tensor inside the region where \(f_a\) takes on non-negligible values. Let us notice that so far, no approximation was made, and that expression is exact; in particular it conserves momentum in \(a\leftrightarrow b\) collisions, since it is in the form \(1/n_am_a\) \(\times \) a factor which is symmetric in the exchange \(a\leftrightarrow b\) \(\times \) the integral which is antisymmetric in the exchange \(a\leftrightarrow b\), as can be seen after three integrations by parts, taking into account that \(\varDelta _cS_a = f_a\). We thus obtain

$$\begin{aligned} n_am_a\frac{{\mathrm{d}}v_{a,i}}{{\mathrm{d}}t} + n_bm_b\frac{{\mathrm{d}}v_{b,i}}{{\mathrm{d}}t} = 0 \end{aligned}$$

When the distribution \(f_a\) is very localized (very cold), in the integral (A.7) we can factor out \(\partial {\mathcal {S}}_b/\partial c_i\). But then, the symmetry which leads to the explicit momentum conservation is broken, because a further assumption was made about \(f_a\) with respect to \(f_b\). If we further assume that \(f_b\) is Maxwellian, inserting Eq. (A.6), we obtain:

$$\begin{aligned} \frac{{\mathrm{d}}v_{a,i}}{{\mathrm{d}}t}&= -4\pi \varGamma _{ab}\left( 1+\frac{m_a}{m_b}\right) \frac{n_b}{3}\left( \frac{m_b}{2\pi k_BT_b}\right) ^{3/2} \nonumber \\&\times R(u)(v_{a,i}-v_{b,i}) \end{aligned}$$
(A.8)

where we recall that:

$$\begin{aligned} u = \left( \frac{m_b}{2k_BT_b}\right) ^{1/2}|\mathbf {v}_a-\mathbf {v}_b| \end{aligned}$$

The exact expression of R(u) was derived above in the case of a Maxwellian \(f_b\). If in addition \(f_b\) is also localized (with a thermal velocity \(k_BT_b/m_b\) much smaller than the relative velocity \(\mathbf {v}_a-\mathbf {v}_b\)), then the lost symmetry is recovered, since we know (see above) that in that case

$$\begin{aligned} R(u) \rightarrow \frac{3\sqrt{\pi }}{4u^3} \end{aligned}$$

so that the slowing-down rate reads:

$$\begin{aligned} \frac{{\mathrm{d}}v_{a,i}}{{\mathrm{d}}t} = -n_b\varGamma _{ab}\left( 1+\frac{m_a}{m_b}\right) \frac{v_{a,i}-v_{b,i}}{|\mathbf {v}_a-\mathbf {v}_b|^3} \end{aligned}$$

We can check that the above expression is in the form \(1/n_am_a\) \(\times \) a term which is antisymmetric in the exchange \(a\leftrightarrow b\). To recover the symmetry needed for momentum conservation outside of the cold distribution limit we can fix the “faulty” part in expression (A.8):

$$\begin{aligned} \frac{4\pi }{3} \left( \frac{m_b}{2\pi k_BT_b}\right) ^{3/2}R(u) \end{aligned}$$

by replacing throughout the quadratic thermal velocity of target particles with an expression which is symmetric in the exchange \(a\leftrightarrow b\), for example the mean:

$$\begin{aligned} \frac{k_BT_b}{m_b} \rightarrow \left( \frac{k_BT_b}{m_b}\right) ^* = \frac{1}{n_a+n_b}\left( n_a\frac{k_BT_a}{m_a}+n_b\frac{k_BT_b}{m_b}\right) \end{aligned}$$

which is an easily accessible quantity in practice since it is the ratio of total pressure to total density, or the sum:

$$\begin{aligned} \frac{k_BT_b}{m_b} \rightarrow \left( \frac{k_BT_b}{m_b}\right) ^* = \frac{k_BT_a}{m_a}+\frac{k_BT_b}{m_b} \end{aligned}$$
(A.9)

which is more satisfactory from a physical point of view because it is supposed to be the squared average relative velocity in the collision of a particle a on a particle b. Using the synthetic expression for R(u) given above, the slowing-down rate is finally put in a form with the requested symmetry:

$$\begin{aligned}&\frac{{\mathrm{d}}v_{a,i}}{{\mathrm{d}}t} = -\frac{n_bm_b}{n_am_a+n_bm_b}\frac{v_{a,i}-v_{b,i}}{\tau _R} \\&\frac{{\mathrm{d}}}{{\mathrm{d}}t}(v_{a,i}-v_{b,i}) = -\frac{v_{a,i}-v_{b,i}}{\tau _R} \end{aligned}$$

where the slowing-down time \(\tau _R\) of ions by target particles reads, in the case of ions (labelled by subscript b):

$$\begin{aligned} \tau _{Rab}&= \frac{m_a^2m_b^2}{4\pi e^4Z_a^2Z_b^2 (m_a+m_b)(n_am_a+n_bm_b)\mathrm {Log}\varLambda _{ab}} \nonumber \\&\quad \times \left( |\mathbf {v}_a-\mathbf {v}_b|^2 + \left( \frac{9\pi }{2}\right) ^{1/3}\left( \frac{k_BT_b}{m_b}\right) ^* \right) ^{3/2} \end{aligned}$$
(A.10)

It may be questionable to use, in the limit of a vanishing relative velocity, an expression of the slowing-down rate which is strictly valid for a single particle a colliding on target particles b. In the limit \(\mathbf {v}_a-\mathbf {v}_b\rightarrow 0\) the slowing-down rate can be computed exactly if the two distributions are assumed to remain Maxwellian (although this is questionable when \(T_a\ne T_b\)).

Thus if \(f_a\) is the Maxwellian for particles of mass \(m_a\) with parameters \(n_a\), \(\mathbf {v}_a\) and \(T_a\), we can write, to first order in \(|\mathbf {v}_a-\mathbf {v}_b|\):

$$\begin{aligned} f_a({\mathbf {c}})&\sim n_a\left( \frac{m_a}{2\pi k_BT_a}\right) ^{3/2}e^{-m_a|\mathbf {c}-\mathbf {v}_b|^2/2k_BT_a} \\&\quad \times \left( 1 + \frac{m_a}{k_BT_a}(v_{a,i}-v_{b,i})(c_i-v_{b,i})\right) \end{aligned}$$

Using expression (A.6) we can then explicitly compute the slowing-down rate:

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(v_{a,i}-v_{b,i}) \sim -\frac{v_{a,i}-v_{b,i}}{\tau _{RM}} \end{aligned}$$

with the slowing-down time

$$\begin{aligned} \tau _{RM} = \frac{3m_a^2m_b^2\left( \frac{k_BT_a}{m_a}+\frac{k_BT_b}{m_b}\right) ^{3/2}}{4\sqrt{2\pi } e^4Z_a^2Z_b^2 (m_a+m_b)(n_am_a+n_bm_b) \mathrm {Log}\varLambda _{ab}} \end{aligned}$$

The vanishing-velocity limit of the approximate expression (A.10) is recovered provided the second definition of the average temperature (A.9) (the sum of the quadratic mean velocities) is used.

In the case of electrons, we only write the part corresponding to the time derivative on the ion velocity:

$$\begin{aligned} \frac{{\mathrm{d}}v_{a,i}}{{\mathrm{d}}t} = -\frac{v_{a,i}}{\tau _{Rae}} \end{aligned}$$

since momentum conservation in collisions involves all ion species simultaneously, and the electron bulk velocity relaxes very quickly to a value determined by that set of species. We then get, to lowest order in powers of the electron/ion mass ratio:

$$\begin{aligned} \tau _{Rae} \sim \frac{3m_a(k_BT_e)^{3/2}}{4\sqrt{2\pi }e^4Z_a^2m_e^{1/2}n_e\mathrm {Log}\varLambda _{ae}} \end{aligned}$$

Definition of a diffusion time From the form of the second Fokker–Planck collision term, a velocity diffusion (or thermalization) time \(\tau _D\) of ions by target particles can be defined in the following way:

$$\begin{aligned} \frac{\partial f}{\partial t} = \frac{\partial }{\partial c_i}\left( D_{ij}\frac{\partial f}{\partial c_j}\right) \end{aligned}$$

with

$$\begin{aligned} \text{ Tr }({\mathbf {D}}) = \frac{1}{2}\frac{{\mathrm{d}}\langle c^2\rangle }{{\mathrm{d}}t} = \frac{\langle c^2\rangle }{\tau _D} \end{aligned}$$

where \(\langle c^2\rangle \) is the mean quadratic width of the distribution f, which reads for a Maxwellian:

$$\begin{aligned} \langle c^2\rangle = 3\frac{k_BT}{m} \end{aligned}$$

In the case of electrons this is:

$$\begin{aligned} \tau _{\mathrm{Dae}} = \frac{T_a}{T_e} \tau _{\mathrm{Rae}} \end{aligned}$$

and for ions (labelled by subscript b):

$$\begin{aligned} \tau _{\mathrm{Dab}} = \frac{T_a}{T_b} \tau _{\mathrm{Rab}} \frac{3R(u)}{R(u)+2L(u)} \end{aligned}$$

where

$$\begin{aligned} u = \left( \frac{m_b}{2k_BT_b}\right) ^{1/2}v \end{aligned}$$

This also reads

$$\begin{aligned} \tau _{\mathrm{Dab}} = \frac{3m_ak_BT_a}{4\pi e^4Z_a^2Z_b^2n_b\mathrm {Log}\varLambda _{ab}}\frac{v}{\mathrm {erf}\left( \left( \frac{m_b}{2k_BT_b}\right) ^{1/2}v\right) } \end{aligned}$$

A practical approximate formula for the ion-ion diffusion time, with the correct limits for \(u\rightarrow 0\) and \(u\rightarrow \infty \), can be designed in the same way as for the slowing-down time:

$$\begin{aligned} \tau _{Dab} = \frac{3m_ak_BT_a}{4\pi e^4Z_a^2Z_b^2n_b\mathrm {Log}\varLambda _{ab}}\left( v^2+\frac{\pi }{2}\frac{k_BT_b}{m_b}\right) ^{1/2} \end{aligned}$$
(A.11)

Close to thermal equilibrium, for all particle species a Fokker–Planck term is recovered, which takes on the form:

$$\begin{aligned} \frac{\partial f_a}{\partial t} = \frac{1}{\tau _{\mathrm{Rab}}}\frac{\partial }{\partial c_i}\left( c_if_a + \frac{k_BT_b}{m_a}\frac{\partial f_a}{\partial c_i}\right) \end{aligned}$$

with a characteristic time which is the slowing-down time \(\tau _{Rab}\) of particles a by the distribution of target particles b.

A global relaxation time The collision times estimated in the preceding sections pertain to the evolution of localized parts of the test-particle distribution in velocity space. Those times will acquire a global meaning for the whole distribution if they can be defined so as to keep, at least approximately, the same value over the region where the test distribution function is not negligible. This clearly applies to the relaxation of an ion distribution on electrons, thanks to the large difference in characteristic velocities, or for ion-ion collisions in the case of a plasma interpenetration with a relative velocity larger than the ion thermal velocity. Moreover, in those two cases, the self-collisions of the ion distribution draw it back to the Maxwellian, which strengthens the global character of the interaction with target particles.

In the case of ion-ion collisions with a relative velocity comparable with the thermal velocity, as already mentioned by Kogan at the end of his paper [69], it is more difficult to define a global relaxation coefficient, even though it can be explicitly calculated in the case of two Maxwellians. The result given by Kogan for temperature relaxation with a vanishing relative velocity (once corrected for a missing factor with respect to the Rosenbluth collision term [41]), is:

$$\begin{aligned} \tau _{ab} = \frac{3(m_bk_BT_a+m_ak_BT_b)^{3/2}}{8\sqrt{2\pi m_am_b}e^4Z_a^2Z_b^2n_b\mathrm {Log}\varLambda _{ab}} \end{aligned}$$
(A.12)

which is the characteristic time to use in the temperature relaxation equation:

$$\begin{aligned} \frac{{\mathrm{d}}T_a}{{\mathrm{d}}t} = \frac{T_b-T_a}{\tau _{ab}} \end{aligned}$$

which leads to the symmetric rate

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(T_a-T_b) = \frac{T_b-T_a}{\tau _{TM}} \end{aligned}$$

where

$$\begin{aligned} \tau _{TM}&= \left( \frac{1}{\tau _{ab}}+\frac{1}{\tau _{ba}}\right) ^{-1} \\&= \frac{3m_am_b\left( \frac{k_BT_a}{m_a}+\frac{k_BT_b}{m_b}\right) ^{3/2}}{8\sqrt{2\pi } e^4Z_a^2Z_b^2 (n_a+n_b) \mathrm {Log}\varLambda _{ab}} \end{aligned}$$

This time is very similar to the limit slowing-down time \(\tau _{RM}\), and is actually the same in the case of equal-mass particles \(m_a=m_b\). This seems to make expression (A.10) a decent candidate for a global relaxation time taking into account plasma interpenetration and/or pressure anisotropy.

But actually the faster particles in the distribution will relax more slowly, so that the distribution will be distorted away from the Maxwellian. In particular when the temperatures are very different (\(T_a\gg T_b\)) we know [79] that the test-particle distribution will acquire a colder component in the target particle region, while the rest of the distribution will slow down with an almost vanishing-divergence current in velocity space, which can actually be used to model the slowing-down of the fast \(\alpha \) particles from fusion reactions in ICF [28, 80]. Even when the temperatures are the same, as discussed in the text, in actual interpenetration calculations performed with a kinetic code, it is found (see the animation provided as supplementary material: animation1.gif) that at the end of the relaxation the distribution shifts from the two-beam to a bi-Maxwellian shape, and accordingly the limit relaxation rate near isotropy is closer to the analytic value found in the latter case (see Appendix B). Thus obviously, instead of looking for a single analytic formula valid for all cases, we have to design a heuristic relaxation rate accounting for the actual behaviour of the plasma, supposedly found in kinetic calculations.

Of course we should not expect a relaxation rate, however cleverly designed, to account for the diversity of kinetic effects. It will only be used as a reasonable order of magnitude in the situations expected to occur in hohlraum plasmas, and specifically as an important input in their modeling through extended hydrodynamics.

Appendix B: Relaxation of the anisotropy of a bi-Maxwellian

Kogan [69] has given an analytic expression of the rate of self-collision relaxation of a bi-Maxwellian, i.e. a distribution reading:

$$\begin{aligned} f(c_x,c_\bot )&= N\left( \frac{m}{2\pi k_BT_\parallel }\right) ^{1/2}\frac{m}{2\pi k_BT_\bot } \nonumber \\&\quad \times \mathrm {exp}\left( \frac{-m}{2k_B}\left( \frac{c_x^2}{T_\parallel }+\frac{c_\bot ^2}{T_\bot }\right) \right) \end{aligned}$$
(B.13)

The general result, valid for all values of the degree of anisotropy, is the following (a correction factor 2 was included, bringing Kogan’s expression of the collision term in agreement with that of Rosenbluth et al. [41]):

$$\begin{aligned} \frac{{\mathrm{d}}T_\parallel }{{\mathrm{d}}t}&= \frac{8Z^4e^4N\text{ Log }\varLambda }{5}\left( \frac{\pi }{m (k_BT)^3}\right) ^{1/2} \nonumber \\&\quad \times (T-T_\parallel )F(\frac{T_\parallel -T}{T}) \end{aligned}$$
(B.14)

where the function F(x) reads:

$$\begin{aligned} F(x)&= -\frac{5\sqrt{1+x}}{x^2} \left[ 1 + \frac{1}{\sqrt{6}}\left( \frac{\sqrt{x}}{2\sqrt{1+x}}\right. \right. \nonumber \\&\quad \left. \left. - \frac{\sqrt{1+x}}{\sqrt{x}}\right) \text{ Log }\frac{\sqrt{1+x}+\sqrt{\frac{3}{2}x}}{\sqrt{1+x}-\sqrt{\frac{3}{2}x}}\right] \end{aligned}$$
(B.15)

That expression is valid without restrictions for \(x>0\) (\(T_\parallel >T_\bot \)), and in the reverse case its analytic extension in the complex plane of values of \(\sqrt{x}\) must be used, noticing that for all complex values of z

$$\begin{aligned} \text{ Log }\frac{1+iz}{1-iz} = 2i\text{ Arctg }z \end{aligned}$$

so that for \(x<0\) :

$$\begin{aligned} F(x)&= -\frac{5\sqrt{1+x}}{x^2} \left[ 1 - \frac{1}{\sqrt{6}}\left( \frac{\sqrt{-x}}{\sqrt{1+x}}\right. \right. \\&\quad \left. \left. +2\frac{\sqrt{1+x}}{\sqrt{-x}}\right) \text{ Arctg }\frac{\sqrt{-\frac{3}{2}x}}{\sqrt{1+x}}\right] \end{aligned}$$

F(x) is plotted as the dashed green curve on Fig. 5. For a small anisotropy \(F(x)\mathrel {\mathop {\rightarrow }\limits _{x\rightarrow 0}}1\), leading to the following definition of the ion-ion collision time for a near-Maxwellian distribution:

$$\begin{aligned} \frac{{\mathrm{d}}T_\parallel }{{\mathrm{d}}t} = \frac{T-T_\parallel }{\tau _{\mathrm{Max}}} \end{aligned}$$

with

$$\begin{aligned} \tau _{\mathrm{Max}} = \frac{5m_i^{1/2}(k_BT_i)^{3/2}}{8\sqrt{\pi }Z^4e^4N_i\text{ Log }\varLambda _{ii}} \end{aligned}$$
(B.16)

The particular numerical factor in the above expression of the collision time arises from the expansion of the relaxation rate about isotropy (\(x\sim 0\)), as can be cross-checked through a direct calculation from the Rosenbluth potentials, given in the next paragraph. It is specific to the relaxation of the anisotropy of a bi-Maxwellian distribution, and numerically different from those found in other collisional relaxation processes near isotropy, even though its order of magnitude and functional dependencies on mass, density and temperature are the same.

Direct calculation from the Rosenbluth potentials The evolution of the second-order moment of the distribution due to collisions reads:

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(Nk_BT_\parallel )&= \int mc_x^2\left( \frac{\partial f}{\partial t}\right) _c(\mathbf {c}){\mathrm{d}}^3c \\&= 8\pi m\varGamma \int \left( {\mathcal {S}}\frac{\partial }{\partial c_x}(2c_xf) + \frac{\partial {\mathcal {T}}}{\partial c_x}\frac{\partial f}{\partial c_x}\right) {\mathrm{d}}^3c \end{aligned}$$

where expressions (A.1)–(A.3) were inserted, dropping all species-specific subscripts since a single species is involved. Using the integral expressions of the potentials (A.4) and (A.5), this reads:

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(Nk_BT_\parallel )&= -m\varGamma \int \int \frac{f(\mathbf {c}^\prime )}{|\mathbf {c}-\mathbf {c}^\prime |} \left( 4f(\mathbf {c})\right. \\&\quad \left. +(5c_x-c_{x}^\prime )\frac{\partial f}{\partial c_x}(\mathbf {c})\right) d^3cd^3c^\prime \end{aligned}$$

Using Kogan’s change of variables (with unit Jacobian):

$$\begin{aligned} (\mathbf {c},\mathbf {c}^\prime ) \rightarrow \left( \mathbf {u} = \mathbf {c}-\mathbf {c}^\prime ,\mathbf {t} = \frac{\mathbf {c}+\mathbf {c}^\prime }{2}\right) \end{aligned}$$

and taking into account that f is the bi-Maxwellian (B.13) to write its \(c_x\)-derivative, we obtain:

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(Nk_BT_\parallel )&= -4m\varGamma \int \int \left( 1 - \frac{m}{k_BT_\parallel }\left( t_x+\frac{3}{4}u_x\right) \right. \\&\left. \times \!\left( t_x\!+\!\frac{u_x}{2}\right) \right) f\left( {\mathbf {t}}\!-\!\frac{{\mathbf {u}}}{2}\right) f\left( {\mathbf {t}}\!+\!\frac{{\mathbf {u}}}{2}\right) \frac{{\mathrm{d}}^3u}{|\mathbf {u}|}{\mathrm{d}}^3t \end{aligned}$$

or, inserting the expression of the distribution function:

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(k_BT_\parallel )&= -4m\varGamma N\left( \frac{m}{2\pi k_BT_\bot }\right) ^2\frac{m}{2\pi k_BT_\parallel } \\&\quad \times \int \int \left( 1 \!-\! \frac{m}{k_BT_\parallel }\left( t_x\!+\!\frac{3}{4}u_x\right) \left( t_x\!+\!\frac{u_x}{2}\right) \right) \\&\quad \times \mathrm {exp}\left( \frac{-m}{k_B}\left( \frac{t_x^2}{T_\parallel }+\frac{t_\bot ^2}{T_\bot }\right) \right) \\&\quad \times \mathrm {exp}\left( \frac{-m}{4k_B}\left( \frac{u_x^2}{T_\parallel }+\frac{u_\bot ^2}{T_\bot }\right) \right) \frac{{\mathrm{d}}^3u}{|\mathbf {u}|}d^3t \end{aligned}$$

Integrations over \(\mathbf {t}\) are straightforward, and we are left with the following integral over \(\mathbf {u}\):

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(k_BT_\parallel )&= -2m\varGamma N\frac{m}{4\pi k_BT_\bot } \left( \frac{m}{4\pi k_BT_\parallel }\right) ^{1/2} \\&\quad \times \int \left( 1 - \frac{3mu_x^2}{4k_BT_\parallel }\right) \mathrm {exp}\\&\quad \left( \frac{-m}{4k_B}\left( \frac{u_x^2}{T_\parallel }+\frac{u_\bot ^2}{T_\bot }\right) \right) \frac{{\mathrm{d}}^3u}{|\mathbf {u}|} \end{aligned}$$

We now define:

$$\begin{aligned} \left( \frac{m}{4k_BT_\parallel }\right) ^{1/2}u_x = r\cos \theta \end{aligned}$$

and

$$\begin{aligned} \left( \frac{m}{4k_BT_\bot }\right) ^{1/2}u_\bot = r\sin \theta \end{aligned}$$

which splits the integral into an angular part and a radial part which can be integrated in a straightforward way, finally leading to:

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(k_BT_\parallel )&= \!-\!\frac{m\varGamma N}{\sqrt{\pi }}\int _0^\pi \frac{(1\!-\!3\cos ^2\theta )\sin \theta d\theta }{\left( \frac{k_BT_\parallel }{m}\cos ^2\theta \!+\!\frac{k_BT_\bot }{m}\sin ^2\theta \right) ^{1/2}} \\&= \frac{8\sqrt{\pi }Z^4e^4N\text{ Log }\varLambda }{\sqrt{mk_BT}}\frac{\sqrt{1+x}}{x} \\&\quad \times \left[ 1 - \frac{1}{\sqrt{6}}\left( \sqrt{\frac{1+x}{x}}-\frac{1}{2}\sqrt{\frac{x}{1+x}}\right) \right. \\&\quad \left. \mathrm {Log}\frac{\sqrt{1+x}+\sqrt{\frac{3x}{2}}}{\sqrt{1+x}-\sqrt{\frac{3x}{2}}}\right] \end{aligned}$$

where we inserted \(x=(T_\parallel -T)/T\). It can be checked that the expression (B.14)–(B.15) obtained by Kogan is recovered (including the previously mentioned correction factor). Expanding the above expression about isotropy (\(x\sim 0\)) we find

$$\begin{aligned} \frac{{\mathrm{d}}}{{\mathrm{d}}t}(k_BT_\parallel )&\mathrel {\mathop {\sim }\limits _{x\sim 0}} -\frac{8\sqrt{\pi }Z^4e^4N\text{ Log }\varLambda }{5\sqrt{mk_BT}}x \\&\mathrel {\mathop {\sim }\limits _{x\sim 0}} \frac{8\sqrt{\pi }Z^4e^4N\text{ Log }\varLambda }{5\sqrt{m(k_BT)^3}}(k_BT-k_BT_\parallel ) \end{aligned}$$

Appendix C: Moments of a two-component distribution with azimuthal symmetry

The velocity distribution function is assumed to be in the form \(f(\mathbf {c}) = f^{(1)}(\mathbf {c}) + f^{(2)}(\mathbf {c})\) with

$$\begin{aligned} f^{(n)}(\mathbf {c}) = \frac{\rho _n}{m}f_\parallel ^{(n)}(\mathbf {c})f_\bot ^{(n)}(\mathbf {c}) \end{aligned}$$

and the factors are defined by

$$\begin{aligned} f_\parallel ^{(n)}(\mathbf {c})= & {} \left( \frac{m}{2k_BT_{\parallel n}}\right) ^{1/2} \nonumber \\&\times F_\parallel \left( \left( \frac{m}{2k_BT_{\parallel n}}\right) ^{1/2}\varOmega _i(c_i-v_{ni})\right) \nonumber \\ \end{aligned}$$
(C.17)

and

$$\begin{aligned} f_\bot ^{(n)}(\mathbf {c})= & {} \frac{m}{2k_BT_{\bot n}}F_\bot \left( (c_i-v_{ni})\right. \nonumber \\&\times \left. \left[ \frac{m}{2k_BT_{\bot n}}(\delta _{ij}-\varOmega _i\varOmega _j)\right] (c_j-v_{nj})\right) \nonumber \\ \end{aligned}$$
(C.18)

\(T_{\parallel n}\) and \(T_{\bot n}\) are the parallel and perpendicular temperatures of beam number n, \(\mathbf {\Omega }=(\mathbf {v}_2-\mathbf {v}_1)/|\mathbf {v}_2-\mathbf {v}_1|\), and the functions \(F_\parallel \) and \(F_\bot \) are normalized as follows:

$$\begin{aligned} \int _{-\infty }^\infty F_\parallel (x){\mathrm{d}}x = 1 \quad ; \quad \int _0^\infty F_\bot (x){\mathrm{d}}x = \frac{1}{\pi } \end{aligned}$$

and for higher moments:

$$\begin{aligned}&\int _{-\infty }^\infty xF_\parallel (x){\mathrm{d}}x = 0 \quad ; \quad \int _{-\infty }^\infty x^2F_\parallel (x){\mathrm{d}}x = \frac{1}{2} \\&\int _0^\infty xF_\bot (x){\mathrm{d}}x = \frac{1}{\pi } \end{aligned}$$

In the specific case of a bi-Maxwellian distribution, we have:

$$\begin{aligned} F_\parallel (x) = \frac{1}{\sqrt{\pi }}\mathrm {e}^{-x^2} \quad \mathrm {and} \quad F_\bot (x) = \frac{1}{\pi }\mathrm {e}^{-x} \end{aligned}$$

The first two moments are obviously

$$\begin{aligned} \rho&= m\int \left( f^{(1)}(\mathbf {c}) + f^{(2)}(\mathbf {c})\right) {\mathrm{d}}^3c = \rho _1+\rho _2 \\ \rho \mathbf {v}&= m\int \mathbf {c}\left( f^{(1)}(\mathbf {c}) + f^{(2)}(\mathbf {c})\right) {\mathrm{d}}^3c = \rho _1\mathbf {v}_1+\rho _2\mathbf {v}_2 \end{aligned}$$

The bulk velocity is the barycentre of the distribution:

$$\begin{aligned} \mathbf {v} = \frac{\rho _1\mathbf {v}_1+\rho _2\mathbf {v}_2}{\rho _1+\rho _2} \end{aligned}$$

The moment of order 2 reads, using notations from [81]:

$$\begin{aligned} mM_{ij}^2 = m\int c_ic_jf(\mathbf {c}){\mathrm{d}}^3c = \rho v_iv_j+P_{ij} \end{aligned}$$

where \(P_{ij}\) is the pressure tensor:

$$\begin{aligned} P_{ij}&= m\int (c_i-v_i)(c_j-v_j)\left( f^{(1)}(\mathbf {c})+f^{(2)}(\mathbf {c})\right) {\mathrm{d}}^3c \nonumber \\&= P_{ij}^{(1)}+P_{ij}^{(2)} + \frac{\rho _1\rho _2}{\rho _1+\rho _2}|\mathbf {v}_2-\mathbf {v}_1|^2\varOmega _i\varOmega _j \end{aligned}$$
(C.19)

where

$$\begin{aligned} P_{ij}^{(n)} = m\int (c_i-v_{ni})(c_j-v_{nj})f^{(n)}(\mathbf {c}){\mathrm{d}}^3c \end{aligned}$$

The vector \(\mathbf {\Omega }\) being given, an orthonormal basis \((\mathbf {\Omega },\mathbf {U},\mathbf {V})\) can be defined, in which the vector \(\mathbf {x}=\mathbf {c}-\mathbf {v}_n\) has components \((x_\parallel ,x_U,x_V)\). The pressure tensor of component n then reads

$$\begin{aligned} P_{ij}^{(n)}= & {} \int (x_\parallel ^2\varOmega _i\varOmega _j+x_U^2U_iU_j+x_V^2V_iV_j + \cdots ) \\&\times \frac{\rho _nm^{3/2}}{(2k_BT_{\parallel n})^{1/2}2k_BT_{\bot n}} F_\parallel \left( \left( \frac{m}{2k_BT_{\parallel n}}\right) ^{1/2}x_\parallel \right) \\&\times F_\bot \left( -\left[ \frac{m(x_U^2+x_V^2)}{2k_BT_{\bot n}}\right] \right) {\mathrm{d}}^3x \end{aligned}$$

where the ellipsis stands for crossed terms such as \(x_\parallel x_U\varOmega _iU_j\) whose integral against \(F_\bot \) vanishes. The sum of non-vanishing terms is

$$\begin{aligned} P_{ij}^{(n)}&= \rho _n\frac{k_BT_{\parallel n}}{m}\varOmega _i\varOmega _j + \rho _n\frac{k_BT_{\bot n}}{m}(U_iU_j+V_iV_j) \\&= \rho _n\frac{k_BT_{\parallel n}}{m}\varOmega _i\varOmega _j + \rho _n\frac{k_BT_{\bot n}}{m}(\delta _{ij}-\varOmega _i\varOmega _j) \end{aligned}$$

Substituting that expression into (C.19) we finally get the pressure tensor for a two-component distribution with azimuthal symmetry around the axis \(\mathbf {\Omega }\):

$$\begin{aligned} P_{ij} = P_\parallel \varOmega _i\varOmega _j + P_\bot (\delta _{ij}-\varOmega _i\varOmega _j) \end{aligned}$$

with

$$\begin{aligned} P_\parallel&= \rho _1\frac{k_BT_{\parallel 1}}{m}+\rho _2\frac{k_BT_{\parallel 2}}{m} +\frac{\rho _1\rho _2}{\rho _1+\rho _2}|\mathbf {v}_2-\mathbf {v}_1|^2 \\ P_\bot&= \rho _1\frac{k_BT_{\bot 1}}{m}+\rho _2\frac{k_BT_{\bot 2}}{m} \end{aligned}$$

The moment of order 3 reads, using notations from [81]:

$$\begin{aligned} mM_{ijk}^3&= m\int c_ic_jc_kf(\mathbf {c}){\mathrm{d}}^3c \\&= \rho v_iv_jv_k+v_iP_{jk}+v_jP_{ik}+v_kP_{ij}+Q_{ijk} \end{aligned}$$

where \(Q_{ijk}\) is twice the heat flux tensor:

$$\begin{aligned} Q_{ijk}&= m\int (c_i-v_i)(c_j-v_j)(c_k-v_k) \\&\quad \times \left( f^{(1)}(\mathbf {c})+f^{(2)}(\mathbf {c})\right) {\mathrm{d}}^3c = Q_{ijk}^{(1)}+Q_{ijk}^{(2)} \\&\quad +\! (v_{1i}\!-\!v_i)P_{jk}^{(1)}\!+\!(v_{1j}\!-\!v_j)P_{ik}^{(1)}\!+\!(v_{1k}\!-\!v_k)P_{ij}^{(1)}\\&\quad +(v_{2i}-v_i)P_{jk}^{(2)} +(v_{2j}-v_j)P_{ik}^{(2)}\\&\quad +(v_{2k}-v_k)P_{ij}^{(2)} +\frac{\rho _1\rho _2(\rho _1-\rho _2)}{(\rho _1+\rho _2)^2}\\&\quad |\mathbf {v}_2-\mathbf {v}_1|^3\varOmega _i\varOmega _j\varOmega _k \end{aligned}$$

\(Q_{ijk}^{(n)}\) is the moment of order 3 restricted to component n and computed in the reference frame centered on its bulk velocity \(\mathbf {v}_{n}\):

$$\begin{aligned} Q_{ijk}^{(n)} = m\int (c_i-v_{ni})(c_j-v_{nj})(c_k-v_{nk})f^{(n)}(\mathbf {c}){\mathrm{d}}^3c \end{aligned}$$

Assuming \(\int _{-\infty }^\infty x^3F_\parallel (x){\mathrm{d}}x=0\), \(Q_{ijk}^{(n)}\) can be computed in the same way as \(P_{ij}^{(n)}\) above. As expected, the result vanishes since the integral contains only terms in which at least one of the factors \(x_p\) enters to an odd power. Inserting values already obtained for \(v_{ni}-v_i\) and \(P_{ij}^{(n)}\) and considering the dependence of the result on degrees of freedom parallel and perpendicular to \(\mathbf {\Omega }\), we finally get

$$\begin{aligned} Q_{ijk}= & {} Q_\parallel \varOmega _i\varOmega _j\varOmega _k + Q_\bot [\varOmega _i(\delta _{jk}-\varOmega _j\varOmega _k) \\&+\varOmega _j(\delta _{ik}-\varOmega _i\varOmega _k)+\varOmega _k(\delta _{ij}-\varOmega _i\varOmega _j)] \end{aligned}$$

with

$$\begin{aligned} Q_\parallel&= \frac{\rho _1\rho _2}{\rho _1+\rho _2}|\mathbf {v}_2-\mathbf {v}_1|\left[ 3\left( \frac{k_BT_{\parallel 2}}{m}-\frac{k_BT_{\parallel 1}}{m}\right) \right. \\&\quad \left. +\frac{\rho _1-\rho _2}{\rho _1+\rho _2}|\mathbf {v}_2-\mathbf {v}_1|^2\right] \\ Q_\bot&= \frac{\rho _1\rho _2}{\rho _1+\rho _2}|\mathbf {v}_2-\mathbf {v}_1|\left( \frac{k_BT_{\bot 2}}{m}-\frac{k_BT_{\bot 1}}{m}\right) \end{aligned}$$

The heat flux vector is

$$\begin{aligned} q_i=\frac{1}{2}Q_{ijj} = \frac{1}{2}(Q_\parallel +2Q_\bot )\varOmega _i \end{aligned}$$

The intrinsic moment of order 4 (computed in the reference frame centered on the global bulk velocity \(\mathbf {v}\)) reads

$$\begin{aligned} R_{ijkm}&= m\int (c_i-v_i)(\ldots )\left( f^{(1)}(\mathbf {c})+f^{(2)}(\mathbf {c})\right) {\mathrm{d}}^3c \\&= R_{ijkm}^{(1)} +\left[ (v_{1i}-v_i)(v_{1j}-v_j)P_{km}^{(1)} + \cdots \right] \\&\quad +\rho _1(v_{1i}-v_i)(v_{1j}-v_j)(v_{1k}-v_k)(v_{1m}-v_m) \\&\quad +R_{ijkm}^{(2)} +\left[ (v_{2i}-v_i)(v_{2j}-v_j)P_{km}^{(2)} + \cdots \right] \\&\quad +\rho _2(v_{2i}-v_i)(v_{2j}-v_j)(v_{2k}-v_k)(v_{2m}-v_m) \end{aligned}$$

where the ellipsis stands for permutations making the preceding expression symmetric, and \(R_{ijkm}^{(n)}\) is the moment of order 4 restricted to component n and computed in the reference frame centered on its bulk velocity \(\mathbf {v}_{n}\):

$$\begin{aligned} R_{ijkm}^{(n)} = m\int x_ix_jx_kx_mf^{(n)}(\mathbf {x}){\mathrm{d}}^3x \end{aligned}$$

where the same definition as above \(x_i=c_i-v_{ni}=x_\parallel \varOmega _i+x_UU_i+x_VV_i\) is used. Hence,

$$\begin{aligned} R_{ijkm}^{(n)}&= m\int x_\parallel ^4f^{(n)}(\mathbf {x}){\mathrm{d}}^3x \varOmega _i\varOmega _j\varOmega _k\varOmega _m \\&\quad + m\int x_\parallel ^2x_U^2f^{(n)}(\mathbf {x}){\mathrm{d}}^3x \left[ \varOmega _i\varOmega _j\left( \delta _{km}-\varOmega _k\varOmega _m\right) \right. \\&\quad \left. +\! \cdots \right] \!+\! m\int x_U^2x_V^2f^{(n)}(\mathbf {x})d^3x \left[ U_iU_jV_kV_m \!+\!\!\cdots \right] \\&\quad +\! m\int x_U^4f^{(n)}(\mathbf {x})d^3x \left( U_iU_jU_kU_m \!+\! V_iV_jV_kV_m\right) \end{aligned}$$

In the latter, it was noticed that

$$\begin{aligned} \int x_U^{2p}f^{(n)}(\mathbf {x}){\mathrm{d}}^3x = \int x_V^{2p}f^{(n)}(\mathbf {x}){\mathrm{d}}^3x \end{aligned}$$

since the distribution is assumed isotropic in the transverse velocity plane. After some manipulations we get

$$\begin{aligned} \left[ U_iU_jV_kV_m +\cdots \right]&= \left( \delta _{ij}-\varOmega _i\varOmega _j\right) \left( \delta _{km}-\varOmega _k\varOmega _m\right) \\&\quad + \left( \delta _{ik}-\varOmega _i\varOmega _k\right) \left( \delta _{jm}-\varOmega _j\varOmega _m\right) \\&\quad + \left( \delta _{im}-\varOmega _i\varOmega _m\right) \left( \delta _{jk}-\varOmega _j\varOmega _k\right) \\&\quad -3\left( U_iU_jU_kU_m + V_iV_jV_kV_m\right) \end{aligned}$$

and since, once again due to isotropy in the transverse velocity plane,

$$\begin{aligned} \int x_U^4 f^{(n)}({\mathbf {x}}){\mathrm{d}}^3x = 3\int x_U^2 x_V^2 f^{(n)}({\mathbf {x}}){\mathrm{d}}^3x \end{aligned}$$

the expression for \(R_{ijkm}^{(n)}\) can be simplified somewhat:

$$\begin{aligned} R_{ijkm}^{(n)} =&\,\, m\int x_\parallel ^4f^{(n)}({\mathbf {x}}){\mathrm{d}}^3x\ \varOmega _i\varOmega _j\varOmega _k\varOmega _m \\&\quad + m\int x_\parallel ^2x_U^2f^{(n)}({\mathbf {x}}){\mathrm{d}}^3x\\&\quad \left[ \varOmega _i\varOmega _j\left( \delta _{km}-\varOmega _k\varOmega _m\right) + \cdots \right] \\&\quad + m\int x_U^4f^{(n)}({\mathbf {x}}){\mathrm{d}}^3x \ \\&\quad \times \frac{1}{3}\left[ \left( \delta _{ij}-\varOmega _i\varOmega _j\right) \left( \delta _{km}-\varOmega _k\varOmega _m\right) \right. \\&\quad + \left( \delta _{ik}-\varOmega _i\varOmega _k\right) \left( \delta _{jm}-\varOmega _j\varOmega _m\right) \\&\quad \left. + \left( \delta _{im}-\varOmega _i\varOmega _m\right) \left( \delta _{jk}-\varOmega _j \varOmega _k\right) \right] \end{aligned}$$

Inserting the form chosen for the distribution components, we get

$$\begin{aligned} m\int x_\parallel ^4f^{(n)}({\mathbf {x}}){\mathrm{d}}^3x&= R_{\parallel \parallel }^{(n)} \\ m\int x_\parallel ^2x_U^2f^{(n)}({\mathbf {x}}){\mathrm{d}}^3x&= \rho _n\frac{k_BT_{\parallel n}}{m}\frac{k_BT_{\bot n}}{m} \\ m\int x_U^4f^{(n)}({\mathbf {x}}){\mathrm{d}}^3x&= 3\rho _n\left( \frac{k_BT_{\bot n}}{m}\right) ^2 \end{aligned}$$

where, in the specific cases of a bi-Maxwellian or a “waterbag” (flat-top) distribution, respectively:

$$\begin{aligned} R_{\parallel \parallel }^{(n)} = 3\rho _n\left( \frac{k_BT_{\parallel n}}{m}\right) ^2 \quad \mathrm {or}\quad \frac{9}{5}\rho _n\left( \frac{k_BT_{\parallel n}}{m}\right) ^2 \end{aligned}$$

The final expression for the tensor of order 4 is thus

$$\begin{aligned} R_{ijkl}= & {} R_{\parallel \parallel } \varOmega _i\varOmega _j\varOmega _k\varOmega _l + R_{\parallel \bot }[\varOmega _i\varOmega _j(\delta _{kl}-\varOmega _k\varOmega _l) + \cdots ] \\&+R_{\bot \bot } [(\delta _{ij}-\varOmega _i\varOmega _j)(\delta _{kl}-\varOmega _k\varOmega _l) +\cdots ] \end{aligned}$$

where the first symmetrized bracket contains 6 terms, and the second one 3 terms. The components are

$$\begin{aligned} R_{\parallel \parallel }= & {} R_{\parallel \parallel }^{(1)} + \rho _1\left[ 6\frac{k_BT_{\parallel 1}}{m}\frac{\rho _2^2|\mathbf {v}_2-\mathbf {v}_1|^2}{(\rho _1+\rho _2)^2} + \frac{\rho _2^4|\mathbf {v}_2-\mathbf {v}_1|^4}{(\rho _1+\rho _2)^4}\right] \nonumber \\&+R_{\parallel \parallel }^{(2)} + \rho _2\left[ 6\frac{k_BT_{\parallel 2}}{m}\frac{\rho _1^2|\mathbf {v}_2-\mathbf {v}_1|^2}{(\rho _1+\rho _2)^2} + \frac{\rho _1^4|\mathbf {v}_2-\mathbf {v}_1|^4}{(\rho _1+\rho _2)^4}\right] \nonumber \\\end{aligned}$$
(C.20)
$$\begin{aligned} R_{\parallel \bot }= & {} \rho _1\frac{k_BT_{\bot 1}}{m}\left( \frac{k_BT_{\parallel 1}}{m}+\frac{\rho _2^2|\mathbf {v}_2-\mathbf {v}_1|^2}{(\rho _1+\rho _2)^2}\right) \nonumber \\&+ \rho _2\frac{k_BT_{\bot 2}}{m}\left( \frac{k_BT_{\parallel 2}}{m}+\frac{\rho _1^2|\mathbf {v}_2-\mathbf {v}_1|^2}{(\rho _1+\rho _2)^2}\right) \end{aligned}$$
(C.21)
$$\begin{aligned} R_{\bot \bot }= & {} \rho _1\left( \frac{k_BT_{\bot 1}}{m}\right) ^2+\rho _2\left( \frac{k_BT_{\bot 2}}{m}\right) ^2 \end{aligned}$$
(C.22)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Larroche, O. An extended hydrodynamics model for inertial confinement fusion hohlraums. Eur. Phys. J. D 75, 297 (2021). https://doi.org/10.1140/epjd/s10053-021-00305-2

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1140/epjd/s10053-021-00305-2

Navigation