Skip to main content
Log in

A Conservative, Entropic Multispecies BGK Model

  • Published:
Journal of Statistical Physics Aims and scope Submit manuscript

Abstract

We derive a conservative multispecies BGK model that follows the spirit of the original, single species BGK model by making the specific choice to conserve species masses, total momentum, and total kinetic energy and to satisfy Boltzmann’s \(\mathcal {H}\)-Theorem. The derivation emphasizes the connection to the Boltzmann operator which allows for direct inclusion of information from higher-fidelity collision physics models. We also develop a complete hydrodynamic closure via the Chapman-Enskog expansion, including a general procedure to generate symmetric diffusion coefficients based on this model. We numerically investigate velocity and temperature relaxation in dense plasmas and compare the model with previous multispecies BGK models and discuss the trade-offs that are made in defining and using them. In particular, we demonstrate that the BGK model in the NRL plasma formulary does not conserve momentum or energy in general.

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

Similar content being viewed by others

Notes

  1. Throughout the paper, we let w and \(\hat{\mathbf {w}}\) be the magnitude and direction of a generic vector \(\mathbf {w}\).

  2. There is some ambiguity in this definition, as there are many choices that can be made for \(b_\mathrm{{max}}\) and \(b_\mathrm{{min}}\) depending on the type of plasma being modeled [46].

  3. The proof of the multispecies \(\mathcal {H}\)-Theorem in [16] and many other references relies on the fact that the functional relationship \(\phi (\mathbf {c}) + \phi (\mathbf {c}_*) = \phi (\mathbf {c}') + \phi (\mathbf {c}_*')\) implies that \(\phi \) is a linear combination of 1, \(\mathbf {c}\), and \(|\mathbf {c}|^2\). The proof of this fact, under fairly general conditions, can be found in [15, Chapter 2.6].

  4. We drop the notation for functional dependence of the collision operators on the distributions from here forward for convenience.

  5. Another reasonable choice for the cross-species Maxwellian parameters \(\mathbf {v}_{ij}\) and \(T_{ij}\) would be to simply set them as \(\mathbf {v}_j\) and \(T_j\), which are properties of the distribution function of species j that are colliding with particles of species i. Indeed, this is what is suggested in the NRL plasma formulary’s BGK model. However, this choice leads to a system that does not conserve momentum or energy, as will be seen later in this paper. Instead, one can think of \(\mathcal {M}_{ij}\) and \(\mathcal {M}_{ji}\) as the effective Maxwellians that a binary mixture of species i and j are relaxing towards.

  6. The collision rate (31) can be computed explicitly for the case of Maxwell molecules, which is not the focus of this paper, which is looking at more general cross sections.

  7. The fact that the leading order term in the expansion is Maxwellian is a direct consequence of the \(\mathcal {H}\)-Theorem; see [16] for more details.

  8. Note that \({\varepsilon } \mathbf {v}_i^{(1)} = \mathbf V _i\).

  9. Note that a term proportional to \(\nabla \log T\) does not appear in the formula for \(\mathbf V _i\). The BGK model does not predict thermal diffusion, also known as the Soret effect, which is seen in the Chapman-Enskog solution to the Boltzmann model.

  10. We do not include the BGK model Garzó et al. [21] in this comparison, as it chooses a different ansatz for the target Maxwellian and this makes direct comparison more difficult. Similarly the Andries et al. BGK model [1] does not use a separate BGK operator for each species pair interaction and thus does not fit in this framework. See the introduction for further discussion of these models.

  11. Note that [35] does not derive the general form for \(\alpha _{ij}\), and goes through the calculation for three different cross sections separately, at least one of which (hard spheres) contains mistakes.

References

  1. Andries, P., Aoki, K., Perthame, B.: A consistent BGK-type model for gas mixtures. J. Stat. Phys. 516, 993–1018 (2002)

    Article  MathSciNet  MATH  Google Scholar 

  2. Balescu, R.: Irreversible processes in ionized gases. Phys. Fluids 3, 52–63 (1960)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  3. Benedict, L.X., Glosli, J.N., Richards, D.F., Streitz, F.H., Hau-Reige, S.P., London, R.A., Graziani, F.R., Murillo, M.S., Benage, J.F.: Molecular dynamics simulations of electron-ion temperature equilibration in an SF6 Plasma. Phys. Rev. Lett. 102, 205004 (2009)

    Article  ADS  Google Scholar 

  4. Bhatnagar, P.L., Gross, E.P., Krook, K.: Uniformly stable numerical schemes for the Boltzmann equation preserving the compressible Navier-Stokes asymptotics. Phys. Rev. 94, 511–524 (1954)

    Article  ADS  MATH  Google Scholar 

  5. Bird, G.A.: Molecular Gas Dynamics and the Direct Simulation of Rarefied Gas Flwos. Clarendon Press, Oxford (1994)

    Google Scholar 

  6. Bobylev, A.: Exact solutions of the Boltzmann equation. Dokl. Akad. Nauk. SSR 225, 1296–1299 (1975)

    ADS  MathSciNet  MATH  Google Scholar 

  7. Bobylev, A., Nanbu, K.: Theory of collision algorithms for gases and plasmas based on the Boltzmann equation and the Landau-Fokker-Planck equation. Phys. Rev. E. 61, 4576–4586 (2000)

    Article  ADS  Google Scholar 

  8. Bobylev, A., Potapenko, I.: Monte Carlo methods and their analysis for Coulomb collisions in multicomponent plasmas. J. Comput. Phys. 246, 123–144 (2013)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  9. Bobylev, A., Rjasanow, S.: Difference scheme for the Boltzmann equation based on the fast fourier transform. Eur. J. Mech. B/Fluids 2, 293–306 (1997)

    MathSciNet  MATH  Google Scholar 

  10. Bobylev, A., Rjasanow, S.: Fast deterministic method of solving the Boltzmann equation for hard spheres. Eur. J. Mech. B/Fluids 18, 869–887 (1999)

    Article  MathSciNet  MATH  Google Scholar 

  11. L. Boltzmann, Lectures on Gas Theory (S. G. Brush, translator), University of California Press, Berkeley, 1964

  12. Brull, S., Pavan, V., Schneider, J.: Derivation of a BGK model for mixtures. Eur. J. Mech.-B/Fluids 33, 74–86 (2012)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  13. Buet, C.: A discrete-velocity scheme for the Boltzmann operator of rarefied gas dynamics. Trans. Theory Stat. Phys. 25, 33–60 (1996)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  14. Caflisch, R.E.: Monte Carlo and quasi-Monte Carlo methods. Acta Numer. 7, 1–49 (1998)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  15. Cercignani, C.: The Boltzmann Equation and Its Applications. Springer, Berlin (1988)

    Book  MATH  Google Scholar 

  16. Chapman, S., Cowling, C.: The Mathematical Theory of Non-Uniform Gases. Cambridge University Press, Cambridge (1970)

    MATH  Google Scholar 

  17. Filbet, F., Mouhot, C., Pareschi, L.: Solving the Boltzmann equation in \(N \log N\). SIAM J. Sci. Comput. 28, 1029–1053 (2006)

    Article  MathSciNet  MATH  Google Scholar 

  18. Gamba, I.M., Haack, J.: A conservative spectral method for the Boltzmann equation with anisotropic scattering and the grazing collisions limit. J. Comput. Phys. 270, 40–57 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  19. Gamba, I.M., Haack, J.R., Hauck, C.D., Hu, J.: A fast spectral method for the Boltzmann collision operator with general collision kernels. arXiv preprint arXiv:1610.00397 (2016)

  20. Gamba, I.M., Tharkabhushaman, S.H.: Spectral-Lagrangian based methods applied to computation of non-equilibrium statistical states. J. Comput. Phys. 228, 2012–2036 (2009)

    Article  ADS  MathSciNet  Google Scholar 

  21. Garzo, V., Santos, A., Brey, J.: A kinetic model for a multicomponent gas. Phys. Fluids A 1, 380–383 (1989)

    Article  ADS  MATH  Google Scholar 

  22. Gericke, D., Murillo, M., Schlanges, M.: Dense plasma temperature equilibration in the binary collision approximation. Phys. Rev. E 65, 036418 (2002)

    Article  ADS  Google Scholar 

  23. Glosli, J.N., Graziani, F.R., More, R.M., Murillo, M.S., Streitz, F.H., Surh, M.P., Benedict, L.X., Hau-Riege, S., Langdon, A.B., London, R.A.: Molecular dynamics simulations of temperature equilibration in dense hydrogen. Phys. Rev. E 78, 025401(R) (2008)

    Article  ADS  Google Scholar 

  24. Greene, J.: Improved Bhatnagar-Gross-Krook model of electron-ion collisions. Phys. Fluids 16, 2022–2023 (1973)

    Article  ADS  Google Scholar 

  25. Hamel, B.: Kinetic model for binary gas mixtures. Phys. Fluids 8, 418–425 (1965)

    Article  ADS  Google Scholar 

  26. Holway, H.: New statistical models for kinetic theory: methods of construction. Phys. Fluids 9, 1658 (1966)

    Article  ADS  Google Scholar 

  27. Huba, J.: Plasma Formulary. Naval Research Laboratory, Washington D. C (2013)

    Google Scholar 

  28. Landau, L.D.: Kinetic equation for the case of Coulomb interaction. Zh. eks. teor. phys. 7, 203 (1937)

    MATH  Google Scholar 

  29. Larroche, O.: Kinetic simulation of a plasma collision experiment. Phys. Fluids B 5, 2816–2840 (1993)

    Article  ADS  Google Scholar 

  30. Larroche, O.: Kinetic simulations of fuel ion transport in ICF target implosions. Eur. Phys. J. D 27, 131–146 (2003)

    Article  ADS  Google Scholar 

  31. Lenard, A.: On Bogoliubov’s kinetic equation for a spatially homogeneous plasma. Ann. Phys. 10, 390–400 (1960)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  32. Liboff, R.: The Theory of Kinetic Equations. Wiley, New York (1969)

    MATH  Google Scholar 

  33. Mieussens, L., Struchtrup, H.: Numerical comparison of Bhatnagar-Gross-Krook models with proper Prandtl number. Phys. Fluids 16, 2797–2813 (2004)

    Article  ADS  MATH  Google Scholar 

  34. Morris, A., Varghese, P., Goldstein, D.: Monte Carlo solution of the Boltzmann equation via a discrete velocity model. J. Comput. Phys. 230, 1265–1280 (2011)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  35. Morse, T.: Energy and momentum exchange between nonequipartition gases. Phys. Fluids 6, 1420–1427 (1963)

    Article  ADS  MathSciNet  Google Scholar 

  36. Morse, T.: Kinetic models equations for a gas mixture. Phys. Fluids 7, 2012–2013 (1964)

    Article  ADS  MathSciNet  Google Scholar 

  37. Munafo, A., Haack, J., Gamba, I., Magin, T.: A spectral-Lagrangian solver for a multi-energy level gas. J. Comput. Phys. 264, 152–176 (2014)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  38. Nanbu, K.: Theory of cumulative small-angle collisions in plasmas. Phys. Rev. E 55, 4642–4652 (1997)

    Article  ADS  Google Scholar 

  39. Paresci, L., Russo, G.: Numerical solution of the Boltzmann equation I: spectrally accurate approximation of the collision operator. SIAM J. Numer. Anal. 37, 1217–1245 (2000)

    Article  MathSciNet  MATH  Google Scholar 

  40. Paresci, L., Russo, G., Toscani, G.: Fast spectral methods for the Fokker-Planck-Landau collision operator. J. Comput. Phys. 165, 216–236 (2000)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  41. Ricci, P., Lapenta, G.: Properties of the Lenard-Balescu collision operator: a numerical study. Phys. Plasmas 9, 430–439 (2002)

    Article  ADS  Google Scholar 

  42. Rogier, F., Schneider, J.: A direct method for solving the Boltzmann equation. Trans. Theory Stat. Phys. 23, 313–338 (1994)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  43. Rosenbluth, M., MacDonald, W., Judd, D.: Fokker-Planck equation for an inverse-square force. Phys. Rev. 107, 1–6 (1957)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  44. Scullard, C., Belt, A., Fennell, S., Jankovic, M., Ng, N., Serna, S., Graziani, F.: Numerical solution of the quantum Lenard-Balescu equation for a one-component plasma. Phys. Plasmas 23(9), 092119 (2016)

    Article  ADS  Google Scholar 

  45. Shakhov, E.M.: On the generalization of the Krook kinetic equation. Izv. Russ. Acad. Sci. Fluid Dyn. 5, 142–145 (1968)

    Google Scholar 

  46. Stanton, L.G., Murillo, M.S.: Ionic transport in high energy density matter. Phys. Rev. E. 93, 043203 (2016)

    Article  ADS  Google Scholar 

  47. Strain, R.M.: On the linearized Balescu-Lenard equation. Commun. Partial Differ. Equ. 32, 1551–1586 (2007)

    Article  MathSciNet  MATH  Google Scholar 

  48. Struchtrup, H.: The BGK-model with velocity-dependent collision frequency. Cont. Mech. Thermodyn. 9, 23–32 (1997)

    Article  MathSciNet  MATH  Google Scholar 

  49. Takizuka, T., Abe, H.: A binary collision model for plasma simulation with a particle code. J. Comput. Phys. 25, 205–219 (1977)

    Article  ADS  MATH  Google Scholar 

  50. Taitano, W.T., Chacón, L., Simakov, A.N., Molvig, K.: A mass, momentum, and energy conserving, fully implicit, scalable algorithm for the multi-dimensional, multi-species Rosenbluth-Fokker-Planck equation. J. Comput. Phys. 297, 357–380 (2015)

    Article  ADS  MathSciNet  MATH  Google Scholar 

  51. Villani, C.: A review of mathematical topics in collisional kinetic theory. Handb. Math. Fluid Dyn. 1, 71–305 (2002)

    Article  MathSciNet  MATH  Google Scholar 

Download references

Acknowledgements

Los Alamos Report LA-UR-16-28001. This manuscript has been authored, in part, by UT-Battelle, LLC, under Contract No. DE-AC0500OR22725 with the U.S. Department of Energy. The United States Government retains and the publisher, by accepting the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up, irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow others to do so, for the United States Government purposes. The Department of Energy will provide public access to these results of federally sponsored research in accordance with the DOE Public Access Plan (http://energy.gov/downloads/doe-public-access-plan). The research of CH was sponsored by the Office of Advanced Scientific Computing Research and performed at the Oak Ridge National Laboratory, which is managed by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725. The research of JH and MSM was performed in part under the auspices of the U.S. Department of Energy by Los Alamos National Laboratory under Contract DE-AC52-06NA25396.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jeffrey R. Haack.

Appendices

Positivity of the Temperature \(T_{ij}\)

In this section, we show the temperature \(T_{ij}\) defined in (36) is non-negative.

Proposition 2

Suppose that \(T_i\) and \(T_j\) are non-negative and that \(\rho _i \nu _{ij} + {\rho _j \nu _{ji}} > 0\). Then the temperature \(T_{ij}\) defined in (36) is also non-negative.

Proof

A quick inspection of (36) shows that it sufficient to show that

$$\begin{aligned} \rho _i \nu _{ij} \bigg (v_i^2 - v_{ij}^2\bigg ) + \rho _j \nu _{ji} \bigg (v_i^2 - v_{ij}^2\bigg ) \end{aligned}$$
(112)

is non-negative. We write the formula for \(\mathbf {v}_{ij}\) in (34) as

$$\begin{aligned} \mathbf {v}_{ij} = \lambda _{ij} \mathbf {v}_i + \lambda _{ji} \mathbf {v}_j, \end{aligned}$$
(113)

where the coefficients

$$\begin{aligned} \lambda _{ij} = \frac{\rho _i \nu _{ij}}{\rho _i \nu _{ij} + {\rho _j \nu _{ji}}} \quad \text {and} \quad \lambda _{ij} = \frac{\rho _j \nu _{ji}}{\rho _i \nu _{ij} + {\rho _j \nu _{ji}}} \end{aligned}$$
(114)

satisfy \(\lambda _{ij} + \lambda _{ji} = 1\).

We divide (112) by \(\rho _i \nu _{ij} + {\rho _j \nu _{ji}}\) and, using the fact that

$$\begin{aligned} v_{ij}^2 = \lambda _{ij}^2 v_i^2 + \lambda _{ji}^2 v_j^2 + 2\lambda _{ij}\lambda _{ji} \mathbf {v}_i \cdot \mathbf {v}_j \end{aligned}$$
(115)

compute

$$\begin{aligned} \lambda _{ij} (v_i^2 - v_{ij}^2) + \lambda _{ji} (v_j^2 - v_{ij}^2)&= \lambda _{ij} v_i^2 + \lambda _{ji} v_j^2 - v_{ij}^2, \nonumber \\&= \lambda _{ij}(1 - \lambda _{ij})v_i^2 + \lambda _{ji}(1 - \lambda _{ji}) v_j^2 - 2\lambda _{ij}\lambda _{ji} \mathbf {v}_i \cdot \mathbf {v}_j, \nonumber \\&= \lambda _{ij}\lambda _{ji} v_i^2 + \lambda _{ji}\lambda _{ij} v_j^2 - 2\lambda _{ij}\lambda _{ji} \mathbf {v}_i \cdot \mathbf {v}_j, \nonumber \\&= \lambda _{ij}\lambda _{ji}(\mathbf {v}_i - \mathbf {v}_j)^2, \end{aligned}$$
(116)

which is non-negative. Hence the expression in (112) is non-negative, and this completes the proof. \(\square \)

Momentum and Temperature Transfer Rates for the Boltzmann Equation

In this section, we reproduce the binary mixture results of Morse [35] in more detail in order to clarify the approximations made in deriving the collision rates in Sect. 4. Morse provides formulas for the momentum and energy relaxation rates for some specific cross sections. Here we provide a more general formula for these quantities that takes a general momentum transfer cross section as an input.

To derive the relaxation rates, we consider a spatially homogeneous gas mixture in which each species has a Maxwellian distribution with number density \(n_i\), bulk velocity \(\mathbf {v}_i\), and temperature \(T_i\) that are slightly out of equilibrium with each other. Taking moments of the Boltzmann equation, we find the formulas for momentum and energy relaxation for species i due to collisions with species j:

$$\begin{aligned} \frac{\partial (\rho _i \mathbf {v}_i)}{\partial t}&= \int m_i \mathbf {c}(\mathcal {M}_i(\mathbf {c}') \mathcal {M}_j(\mathbf {c}_*') - \mathcal {M}_i(\mathbf {c}) \mathcal {M}_j(\mathbf {c}_*)) g \sigma _{ij}\, d\varvec{\Omega }d\mathbf {c}_*d\mathbf {c}, \nonumber \\&= \int m_i (\mathbf {c}' - \mathbf {c}) \mathcal {M}_i(\mathbf {c}) \mathcal {M}_j(\mathbf {c}_*)g \sigma _{ij} \, d\varvec{\Omega }d\mathbf {c}_*d\mathbf {c}, \end{aligned}$$
(117)
$$\begin{aligned} \frac{\partial (n_i E_i)}{\partial t}&= \int \frac{m_i c ^2}{2} (\mathcal {M}_i(\mathbf {c}') \mathcal {M}_j(\mathbf {c}_*') - \mathcal {M}_i(\mathbf {c}) \mathcal {M}_j(\mathbf {c}_*)) g \sigma _{ij}\, d\varvec{\Omega }d\mathbf {c}_*d\mathbf {c}, \nonumber \\&= \int \frac{m_i}{2} ((c')^2 - c^2) \mathcal {M}_i(\mathbf {c}) \mathcal {M}_j(\mathbf {c}_*)g \sigma _{ij}\, d\varvec{\Omega }d\mathbf {c}_*d\mathbf {c}. \end{aligned}$$
(118)

Here \(n_i E_i = \int \frac{m_i}{2} c^2 f_i d\mathbf {c}\) is the total kinetic energy of species i.

To compute the above integrals, we make a change of variables from \(\mathbf {c}\) and \(\mathbf {c}_*\) to the relative velocity \(\mathbf {g}\) and the center of mass velocity \(\mathbf {G}\):

$$\begin{aligned} \mathbf {g}= \mathbf {c}- \mathbf {c}_*\quad \text {and} \quad \mathbf {G}= \frac{m_i \mathbf {c}+ m_j \mathbf {c}_*}{m_i + m_j}. \end{aligned}$$
(119)

In terms of \(\mathbf {g}\) and \(\mathbf {G}\), (117) and (118) take the form

$$\begin{aligned} \frac{\partial (\rho _i \mathbf {v}_i)}{\partial t}&= -\mu _{ij} \int \mathbf {g}\mathcal {M}_i(\mathbf {c}) \mathcal {M}_j(\mathbf {c}_*) g \sigma _{ij}^{(1)} (g)\, d\mathbf {g}d\mathbf {G}, \end{aligned}$$
(120)
$$\begin{aligned} \frac{\partial (n_i E_i)}{\partial t}&= -\mu _{ij} \int \mathcal {M}_i(\mathbf {c}) \mathcal {M}_j(\mathbf {c}_*) (\mathbf {G}\cdot \mathbf {g})\, g \sigma _{ij}^{(1)}(g) d\mathbf {g}d\mathbf {G}, \end{aligned}$$
(121)

where \(\sigma _{ij}^{(1)}\) is the momentum transfer cross section defined by

$$\begin{aligned} \sigma _{ij}^{(1)} = 2\pi \int (1 - \cos \theta ) \sigma (g,\cos \theta ) \sin \theta d\theta . \end{aligned}$$
(122)

This formula for \(\sigma _{ij}\) arises when choosing \(\mathbf {g}\) as the polar direction for \(\varvec{\Omega }\) in the angular integrals in (117) and (118); that is

$$\begin{aligned} \int (\mathbf {g}- g \mathbf{\Omega } ) \sigma _{ij}(g,\varvec{\Omega }) d\varvec{\Omega }&= 2\pi \mathbf {g}\int (1- \cos \theta ) \sin \theta \sigma _{ij}(g,\cos \theta ) d\theta . \end{aligned}$$
(123)

Next we rewrite the Maxwellians that appear in (120) and (121) in terms of the variables \(\mathbf {G},\mathbf {g}\) and the velocity difference \(\Delta \mathbf {v}= \mathbf {v}_j - \mathbf {v}_i\):

$$\begin{aligned} \mathcal {M}_i(\mathbf {c}) \mathcal {M}_j(\mathbf {c}_*)&= \frac{n_in_j (m_im_j)^{3/2}}{(2\pi T_i)^{3/2}(2\pi T_j)^{3/2}}&e^{-\frac{m_j T_i + m_i T_j}{2 T_i T_j}\left[ \mathbf {G}+ \frac{T_j(\mu _{ij}\mathbf {g}- m_i \mathbf {v}_i) - T_i(\mu _{ij}\mathbf {g}+ m_j \mathbf {v}_j)}{m_iT_j + m_jT_i}\right] ^2} \nonumber \\&\ \quad \times \; e^{-\frac{m_im_j}{2(m_iT_j + m_jT_i)}(\mathbf {g}+ \Delta \mathbf {v})^2}. \end{aligned}$$
(124)

Performing the \(\mathbf {G}\) integration in (120) and (121), we obtain

$$\begin{aligned} \frac{\partial (\rho _i \mathbf {v}_i)}{\partial t}&= -\frac{n_i n_j (m_i m_j)^{3/2} \mu _{ij}}{(2\pi )^{3/2} (m_i T_j + m_j T_i)^{3/2}} \int \mathbf {g}g \sigma _{ij}^{(1)}(g) e^{-\frac{m_im_j}{2(m_iT_j + m_jT_i)}(\mathbf {g}+ \Delta \mathbf {v})^2} d\mathbf {g},\end{aligned}$$
(125)
$$\begin{aligned} \frac{\partial n_i E_i}{\partial t}&= -\frac{n_i n_j (m_i m_j)^{3/2} \mu _{ij}}{(2\pi )^{3/2} (m_i T_j + m_j T_i)^{5/2}} \nonumber \\&\quad \times \int \left[ \mu _{ij}(T_i - T_j)g^2 + \mathbf {g}\cdot (m_i T_j \mathbf {v}_i + m_j T_i \mathbf {v}_j)\right] g \sigma _{ij}^{(1)}(g) e^{-\frac{m_im_j}{2(m_iT_j + m_jT_i)}(\mathbf {g}+ \Delta \mathbf {v})^2} d\mathbf {g}. \end{aligned}$$
(126)

Next we recast the energy transfer Eq. (126) as temperature transfer via

$$\begin{aligned} \frac{\partial n_i E_i}{\partial t}&= \frac{3}{2} n_i \frac{\partial T_i}{\partial t} + \frac{\partial \rho _i \mathbf {v}_i}{\partial t} \cdot \mathbf {v}_i. \end{aligned}$$
(127)

Thus, using (125), we have

$$\begin{aligned} \frac{3}{2} n_i \frac{\partial T_i}{\partial t}&= -\frac{n_i n_j (m_i m_j)^{3/2} \mu _{ij}}{(2\pi )^{3/2} (m_i T_j + m_j T_i)^{5/2}} \nonumber \\&\quad \times \; \int \left[ \mu _{ij}(T_i - T_j)g^2 + m_j T_i \mathbf {g}\cdot \Delta \mathbf {v}\right] g \sigma _{ij}^{(1)}(g) e^{-\frac{m_im_j}{2(m_iT_j + m_jT_i)}(\mathbf {g}+ \Delta \mathbf {v})^2} d\mathbf {g}. \end{aligned}$$
(128)

To perform the integration in (128), we rewrite \(\mathbf {g}\) using spherical coordinates with \(\Delta \mathbf {v}\) as the polar direction and obtain

$$\begin{aligned} \frac{3}{2} n_i \frac{\partial T_i}{\partial t}&= -\frac{ n_i n_j (m_i m_j)^{3/2} \mu _{ij}}{(2\pi )^{3/2} (m_i T_j + m_j T_i)^{5/2}} \\&\quad \times \; \Bigg [ 4\pi \mu _{ij}(T_i - T_j) \int _0^\infty g^5 \sigma _{ij}^{(1)}(g) e^{-K(g^2 + \Delta v^2)} \frac{\sinh (2K g \Delta v) }{2Kg\Delta v} dg \nonumber \\&- 4\pi m_j T_i (\Delta v)^2 \int _0^\infty g^4 \sigma _{ij}^{(1)}(g) e^{-K(g^2 + \Delta v^2)} \frac{2Kg\Delta v \cosh (2Kg\Delta v) - \sinh (2Kg\Delta v)}{4 K^2 g^2 (\Delta v)^3} dg\Bigg ], \nonumber \end{aligned}$$
(129)

where \(K = \frac{m_i m_j}{2(m_iT_j + m_j T_i)}\). Next, we make the approximations that \(K (\Delta v)^2\) and \((T_j -T_i)\) are small. Taylor expanding the hyperbolic functions and making the change of variable \(K g^2 = w^2\), the result is

$$\begin{aligned} \frac{3}{2} n_i \frac{\partial T_i}{\partial t}&= - \frac{ n_i n_j (m_i m_j)^{3/2} \mu _{ij}}{(2\pi )^{3/2} (m_i T_j + m_j T_i)^{5/2}} \nonumber \\ \times&\Bigg [ \frac{4\pi }{K^3} \mu _{ij}(T_i - T_j) \int _0^\infty w^5 \sigma _{ij}^{(1)}(w) e^{-w^2} e^{-K\Delta v^2} \left( 1 + \frac{2}{3} w^2 K (\Delta v)^2 + O(K^2(\Delta v)^4)\right) dw \nonumber \\&- \frac{4\pi m_j T_i}{K^3} K(\Delta v)^2 \int _0^\infty w^5 \sigma _{ij}^{(1)}(w) e^{-w^2} e^{-K\Delta v^2} \left( \frac{2}{3} + \frac{16}{75}w^2 K (\Delta v)^2 + O(K^2 (\Delta v)^4) \right) dw\Bigg ]. \end{aligned}$$
(130)

Neglecting terms of \(O(K^2 (\Delta v)^4)\) and higher in the above equation, we obtain

$$\begin{aligned} \frac{3}{2} n_i \frac{\partial T_i}{\partial t}&= - \frac{ 4\pi n_i n_j (m_i m_j)^{3/2} \mu _{ij}}{(2\pi )^{3/2} K^3 (m_i T_j + m_j T_i)^{5/2}} \nonumber \\&\quad \times \; e^{-K\Delta v^2} \left( \mu _{ij}I_5^{ij} \left( (T_i - T_j) - \frac{1}{3} m_j (\Delta v)^2\right) + \frac{2}{3} \mu _{ij}(I_7^{ij} - I_5^{ij})(T_i - T_j) K (\Delta v)^2 \right) , \end{aligned}$$
(131)

where

$$\begin{aligned} I_n^{ij} = \int _0^\infty w^n \sigma ^{(1)}_{ij}(w) e^{-w^2} dw \quad \text {and} \quad w^2 = K g^2. \end{aligned}$$
(132)

Neglecting terms of order \((T_j - T_i) K (\Delta v)^2\) in (131) we obtain the general formFootnote 11

$$\begin{aligned} \frac{3}{2} n_i \frac{\partial T_i}{\partial t}&= \frac{ 8 \pi n_i n_j (m_i m_j)^{3/2} \mu _{ij}^2}{3(2\pi )^{3/2} K^3 (m_i T_j + m_j T_i)^{5/2}} I_5^{ij} \left( \frac{3}{2} (T_j - T_i) + \frac{1}{2} m_j (\Delta v)^2\right) , \nonumber \\&= \frac{ 64 \pi n_i n_j (m_i T_j + m_j T_i)^{1/2} \mu _{ij}^2}{3(2\pi )^{3/2} (m_i m_j)^{3/2} } I_5^{ij} \left( \frac{3}{2} (T_j - T_i) + \frac{1}{2} m_j (\Delta v)^2\right) , \nonumber \\&= \alpha _{ij} \left( \frac{3}{2} (T_j - T_i) + \frac{1}{2} m_j (\Delta v)^2\right) . \end{aligned}$$
(133)

Here

$$\begin{aligned} \alpha _{ij} = \frac{64\pi \mu _{ij}^2 n_i n_j (m_i T_j + m_j T_i)^{1/2}}{3 (2\pi m_i m_j)^{3/2}} I_5^{ij} \end{aligned}$$
(134)

is the coefficient for energy transfer and the key quantity used in deriving BGK collision rates.

Using (133), we write the relaxation of temperature differences as

$$\begin{aligned} \frac{\partial (T_j - T_i)}{\partial t}&= -\alpha _{ij} \left[ \frac{n_i + n_j}{n_i n_j} (T_j - T_i) + \frac{1}{3} \frac{\rho _j - \rho _i}{n_i n_j} (\Delta v)^2\right] , \end{aligned}$$
(135)

For momentum transfer, a similar calculation for the integral in (125) gives

$$\begin{aligned} \frac{\partial (\rho _i \mathbf {v}_i)}{\partial t}&= \frac{32\pi n_i n_j \mu _{ij}(m_i T_j + m_j T_i)^{1/2}}{3 (2\pi )^{3/2} (m_i m_j)^{1/2} } I_5^{ij}\ \Delta \mathbf {v}, \end{aligned}$$
(136)

and the relaxation of the velocity difference is

$$\begin{aligned} \frac{\partial (\mathbf {v}_j - \mathbf {v}_i)}{\partial t}&= - \alpha _{ij} \left( \frac{\rho _i + \rho _j}{\rho _i \rho _j}\right) \frac{ (m_i + m_j) }{2} (\mathbf {v}_j - \mathbf {v}_i). \end{aligned}$$
(137)

Two Useful Lemmas

Below are two useful lemmas that are used in the computation of temperature and velocity relations in Sect. 5.1. The proofs are provided for the reader’s convenience.

Lemma 1

Let \(A \in \mathbf {R}^{n \times n}\) be a symmetric matrix with strictly positive components. Then the linear system

$$\begin{aligned} -\Big ( \sum _j A_{ij} \Big ) w_i + \sum _j A_{ij} w_j = 0, \quad i=1,\dots ,n. \end{aligned}$$
(138)

has a one parameter family of solutions

$$\begin{aligned} w_i = \bar{w} = \frac{1}{n} \sum _j w_j, \quad i=1,\dots ,n. \end{aligned}$$
(139)

Furthermore, the matrix corresponding to the linear system in (138) is negative semi-definite.

Proof

Because A is symmetric, multiplying (138) by \(w_i\) and summing over the index i gives

$$\begin{aligned} 0 = \sum _{i,j} A_{ij} \bigg (w_iw_j - w_i^2\bigg ) = \sum _{i,j} A_{ij} \bigg (w_i w_j - w_j^2\bigg ) = -\frac{1}{2} \sum _{i,j} A_{ij} (w_i - w_j)^2 \end{aligned}$$
(140)

Hence because the components of A are all strictly positive, \(w_i\) must be independent of i, and the formula for \(\bar{w}\) follows immediately. \(\square \)

The next Lemma is used to derive the symmetric form of the diffusion coefficient matrix D in (98).

Lemma 2

Let \(A \in \mathbf {R}^{n \times n}\) be a symmetric matrix given by

$$\begin{aligned} A_{ij} = \gamma _{ij} > 0, \quad i \ne j, \quad \qquad A_{ii} = - \sum _{j,j\ne i} A_{ij}, \end{aligned}$$
(141)

and let \(B \in \mathbf {R}^{n \times n}\) be the symmetric matrix

$$\begin{aligned} B = A - \alpha \mathbf {a}\mathbf {a}^T, \end{aligned}$$
(142)

where \(\mathbf {a}\) is a vector such that \(a_i > 0\ \forall i\) and \(\alpha > 0\). Furthermore, let \(\mathbf {b}\) satisfy \(\sum _j b_j = 0\). Then B is a symmetric negative definite matrix and the linear system

$$\begin{aligned} B\mathbf {w}= \mathbf {b}\end{aligned}$$
(143)

has the same solution as the constrained system

$$\begin{aligned} A\mathbf {w}&= \mathbf {b}\nonumber \\ \mathbf {a}^T \mathbf {w}&= 0. \end{aligned}$$
(144)

Furthermore, the solution is independent of the magnitude of \(\alpha \).

Proof

We will first show that B is negative definite. We have

$$\begin{aligned} \sum _i \sum _j B_{ij} w_i w_j = -\frac{1}{2} \sum _i \sum _j A_{ij}(w_i - w_j)^2 - \alpha \left( \sum _i w_i a_i\right) ^2 < 0, \end{aligned}$$
(145)

where we have used the result of Lemma 1, and the assumption that \(a_i > 0\).

We next show that (143) and (144) are equivalent. It is obvious that a solution of (144) is also a solution of (143). Thus we need only to show that (143) implies (144). We suppose that \(\mathbf {w}\) solves (143) and write A in terms of its eigenvector decomposition

$$\begin{aligned} A = \sum _{j=1}^{n} \lambda _j \mathbf {q}_j \mathbf {q}_j^T. \end{aligned}$$
(146)

Because A is symmetric, these eigenvectors are orthogonal. In Lemma 1, we showed that the null space A is spanned by the vector \(\mathbf e = \frac{1}{\sqrt{n}} (1\ 1\ \dots \ 1)^T\), which we identify with the eigenvector \(\mathbf {q}_n\) having eigenvalue \(\lambda _n = 0\). Using (146), we can write the system in (143) as

$$\begin{aligned} \left( \sum _{j=1}^{n-1} \lambda _j \mathbf {q}_j \mathbf {q}_j^T - \alpha \mathbf {a}\mathbf {a}^T\right) \mathbf {w}= \mathbf {b}. \end{aligned}$$
(147)

Left multiplying (147) by \(\mathbf {q}_n^T\), we have

$$\begin{aligned} - \alpha \mathbf {q}_n^T \mathbf {a}\mathbf {a}^T \mathbf {w}= \mathbf {q}_n^T \mathbf {b}= 0, \end{aligned}$$
(148)

due to the orthogonality of the eigenvectors and the assumption on \(\mathbf {b}\) that its components sum to zero. Because the elements of \(\mathbf {a}\) are all strictly positive, \(\alpha \mathbf {q}_n^T \mathbf {a}>0\). It follows then from (148) that \(\mathbf {a}^T \mathbf {w}= 0\). Hence (147) reduces to

$$\begin{aligned} \sum _{j=1}^{n-1} \lambda _j \mathbf {q}_j \mathbf {q}_j^T \mathbf {w}= A\mathbf {w}= \mathbf {b}. \end{aligned}$$
(149)

Therefore \(\mathbf {w}\) solves (144), and the two systems are equivalent.

Next, we show that the solution to (143) is independent of \(\alpha \). To this end, suppose that \(B_1 \mathbf {w}_1 = B_2 \mathbf {w}_2 = \mathbf {b}\), where \(B_i = A - \alpha _i \mathbf {a}\mathbf {a}^T\), \(i=1,2\), and \(\alpha _1,\alpha _2 > 0\). Following the arguments above, we conclude that \(\mathbf {a}^T \mathbf {w}_1 = \mathbf {a}^T\mathbf {w}_2 = 0\). Hence,

$$\begin{aligned} 0 = B_1 \mathbf {w}_1 - B_2 \mathbf {w}_2 = A(\mathbf {w}_1 - \mathbf {w}_2) = B_1 (\mathbf {w}_1 - \mathbf {w}_2). \end{aligned}$$
(150)

Since \(B_1\) is invertible, it follows that \(\mathbf {w}_1 = \mathbf {w}_2\). \(\square \)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Haack, J.R., Hauck, C.D. & Murillo, M.S. A Conservative, Entropic Multispecies BGK Model. J Stat Phys 168, 826–856 (2017). https://doi.org/10.1007/s10955-017-1824-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10955-017-1824-9

Keywords

Navigation