Loading [MathJax]/jax/output/SVG/jax.js

Modelling and numerical study of the polyatomic bitemperature Euler system

  • Received: 01 April 2021 Revised: 01 January 2022 Published: 02 April 2022
  • Primary: 82C40; Secondary: 76X05, 65M08, 35L60

  • This paper is devoted to the study of the bitemperature Euler system in a polyatomic setting. Physically, this model describes a mixture of one species of ions and one species of electrons in the quasi-neutral regime. We firstly derive the model starting from a kinetic polyatomic model and performing next a fluid limit. This kinetic model is shown to satisfy fundamental properties. Some exact solutions are presented. Finally, a numerical scheme is derived and proved to coincide with an approximation designed in [3] and extended to second order and two space dimensions in [6]. Some numerical tests are presented.

    Citation: Denise Aregba-Driollet, Stéphane Brull. Modelling and numerical study of the polyatomic bitemperature Euler system[J]. Networks and Heterogeneous Media, 2022, 17(4): 593-611. doi: 10.3934/nhm.2022018

    Related Papers:

    [1] Nora Aïssiouene, Marie-Odile Bristeau, Edwige Godlewski, Jacques Sainte-Marie . A combined finite volume - finite element scheme for a dispersive shallow water system. Networks and Heterogeneous Media, 2016, 11(1): 1-27. doi: 10.3934/nhm.2016.11.1
    [2] Young-Pil Choi, Seok-Bae Yun . A BGK kinetic model with local velocity alignment forces. Networks and Heterogeneous Media, 2020, 15(3): 389-404. doi: 10.3934/nhm.2020024
    [3] José Antonio Carrillo, Yingping Peng, Aneta Wróblewska-Kamińska . Relative entropy method for the relaxation limit of hydrodynamic models. Networks and Heterogeneous Media, 2020, 15(3): 369-387. doi: 10.3934/nhm.2020023
    [4] Steinar Evje, Kenneth H. Karlsen . Hyperbolic-elliptic models for well-reservoir flow. Networks and Heterogeneous Media, 2006, 1(4): 639-673. doi: 10.3934/nhm.2006.1.639
    [5] John D. Towers . The Lax-Friedrichs scheme for interaction between the inviscid Burgers equation and multiple particles. Networks and Heterogeneous Media, 2020, 15(1): 143-169. doi: 10.3934/nhm.2020007
    [6] Piotr Gwiazda, Karolina Kropielnicka, Anna Marciniak-Czochra . The Escalator Boxcar Train method for a system of age-structured equations. Networks and Heterogeneous Media, 2016, 11(1): 123-143. doi: 10.3934/nhm.2016.11.123
    [7] Maya Briani, Emiliano Cristiani . An easy-to-use algorithm for simulating traffic flow on networks: Theoretical study. Networks and Heterogeneous Media, 2014, 9(3): 519-552. doi: 10.3934/nhm.2014.9.519
    [8] Raimund Bürger, Kenneth H. Karlsen, John D. Towers . On some difference schemes and entropy conditions for a class of multi-species kinematic flow models with discontinuous flux. Networks and Heterogeneous Media, 2010, 5(3): 461-485. doi: 10.3934/nhm.2010.5.461
    [9] Huanhuan Li, Meiling Ding, Xianbing Luo, Shuwen Xiang . Convergence analysis of finite element approximations for a nonlinear second order hyperbolic optimal control problems. Networks and Heterogeneous Media, 2024, 19(2): 842-866. doi: 10.3934/nhm.2024038
    [10] Caterina Balzotti, Simone Göttlich . A two-dimensional multi-class traffic flow model. Networks and Heterogeneous Media, 2021, 16(1): 69-90. doi: 10.3934/nhm.2020034
  • This paper is devoted to the study of the bitemperature Euler system in a polyatomic setting. Physically, this model describes a mixture of one species of ions and one species of electrons in the quasi-neutral regime. We firstly derive the model starting from a kinetic polyatomic model and performing next a fluid limit. This kinetic model is shown to satisfy fundamental properties. Some exact solutions are presented. Finally, a numerical scheme is derived and proved to coincide with an approximation designed in [3] and extended to second order and two space dimensions in [6]. Some numerical tests are presented.



    This work is devoted to the modelling and the numerical approximation of the nonconservative polyatomic bitemperature Euler system in the context of plasma physics. Physically, this model describes the interaction of one species of ions and one species of electrons, under the quasi-neutrality assumption.

    The aim of this paper is more precisely to provide a construction and an approximation of the polyatomic Euler bitemperature system. This model is constituted of conservative equations for the mass and the momentum and two nonconservative equations for ionic and electronic energies. It is a variant of the system considered in [23]. The nonconservativity is due to source-terms but also to the presence of products of the velocity by pressure gradients. In Inertial Confinement Fusion applications, high temperature solutions involve shocks, for which those nonconservative products have to be determined. This can be done by defining paths, as proposed by Dal Maso, Le Floc'h and Murat ([24]). However, it is shown in [1] that the numerical adaptation of this theory given in [36] is delicate.

    In order to compute physically consistent shocks, one can use an underlying kinetic system. This approach was usefully adopted in [3] for the monoatomic case. The model is constituted by a kinetic system coupled with Ampère and Poisson equations. Moreover, this construction leads to a kinetic numerical scheme starting from a semi-discretization in space and time of the kinetic model. A DVM approach of this model has also been considered in [20] together with an asymptotic preserving discretisation toward the bitemperature Euler system. However relaxation schemes and discrete BGK schemes have been developed in the polytropic situation for a general γ law including the polyatomic case in [3], [6]. It seems important to consider situations where each species has its own γ constant. Note that for smooth solutions, global in time existence has been proved under the condition that these constants are distinct, (see [5]).

    We construct in this paper an extension of the model developed in [3] to a polyatomic setting with a continuous energy variable. This model is based on the use of several attractors like in [15] and is shown to satisfy an H theorem. We refer to [10], [13], [14], [19], [11], [34] for other BGK models for polyatomic gas mixtures.

    In the present paper, the derivation is based again on an hydrodynamic limit performed from an underlying polyatomic kinetic model. In the present case, the unknowns of the kinetic equations are the distribution functions f(t,x,v,I) depending on time t, space x, velocity v and of one-dimensional internal energy parameter I>0. This energy parameter collects vibrational and rotational energies. Kinetic models with continuous energy variables have been introduced in [16] where the motivation was to develop a Monte-Carlo method. In [17], the authors derived a mesoscopic model of Boltzmann type associated to the previous microscopic model. This collision operator satisfies fundamental properties (H theorem, ). The generalization to mixtures has been performed in [26]. In [9], a compressible Navier-Stokes model has been derived for mixture of monoatomic and polyatomic gases. For some applications of such models we refer to [35], [38], [29]. In [35] the authors analyse the shock wave structure of some polyatomic gases. So by using an ESBGK model ([2], [21]) they show the presence of a double layer structure that is specific to the polyatomic setting. We mention that in [27], [30], the authors develop polyatomic models with a discrete internal energy variable.

    As observed in [37], [3], it is possible to find an underlying kinetic model where the force term is related to the nonconservative terms. One advantage of the kinetic model, is its conservative form. In the present paper, the kinetic model describing the interspecies interaction is a two component polyatomic BGK model based on one continuous internal energy variable coupled with Ampère and Poisson equations. Hence starting from a standard semi-discretization of this model, the hydrodynamic limit leads to a numerical scheme for the bitemperature model. It must be noted that the obtained scheme is the same as the one obtained in section 3.2 of [3]. In this article however the scheme was obtained by a very different method involving models developed in [7] which are not based on a physically realistic kinetic approximation of the equations. The novelty here is that our polyatomic BGK model gives a physical justification to it in the general case where γe and γi may be distinct. In [4] diffusive terms have been constructed from a Chapman-Enskog expansion up to order 1 leading to a Navier-Stokes type model. This last model generalizes the model that is studied in [22]. In [31], the authors perform a Chapman-Enskog expansion by introducing a small parameter representing the ratio between electronic and ionic molecular masses. They obtain an hyperbolic system with a parabolic regularisation on the electrons. The structure of shock wave solutions for this last model is considered in [39].

    The plan of the paper is the following. Section 2 deals with the presentation of the different models that are used in this paper. In particular, we detail the eigenstructure of the bitemperature system. In section 3, the fluid model is obtained starting from the polyatomic model. In section 4 exact solutions for the model are computed and in section 5, the numerical scheme is developed. The last part is devoted to some numerical tests in 1D and 2D configurations.

    The bitemperature Euler system reads

    {tρ+x(ρu)=0,t(ρu)+x(ρu2+pe+pi)=0,tEe+x(u(Ee+pe))u(cixpecexpi)=νei(TiTe),tEi+x(u(Ei+pi))+u(cixpecexpi)=νei(TiTe), (1)

    where ρ>0 represents the total density of the plasma, u is the average velocity of the plasma, Ee and Ei are the total electronic and ionic energies. Te and Ti represent the electronic and ionic temperatures. The coefficient νei is nonnegative. The physically realistic formulas which are given in [33] make the source term very stiff.

    One has

    ρ=ρe+ρi (2)

    where ρe=neme, ρi=nimi are electronic and ionic densities, electronic and ionic concentrations ne and ni being assumed to be linked by Z=ne/ni1. Z is considered as constant. This situation corresponds to the quasi-neutral regime. me and mi represent the electronic and ionic masses. The mass fractions

    cβ=ρβρ,β=e,i (3)

    are then constant and ce and ci are given by

    ce=Zmemi+Zme,ci=1ce. (4)

    The total energies are linked to the internal electronic and ionic energies by

    Eβ=ρβεβ+12ρβu2,β{e,i}.

    The electronic and ionic pressures and temperatures are related by pe=nekBTe and pi=nikBTi. The electronic and ionic internal energies are then given by

    εe=kB(γe1)meTe,εi=kB(γi1)miTi,

    where γe,γi [1,3], and kB is the Boltzmann constant.

    In the following we denote U=(ρ,ρu,Ee,Ei), Uβ=(ρβ,ρβu,Eβ).

    We complete here the calculations presented in [5]. The system is rewritten by using the variables V=(ρ,u,εe,εi):

    {tρ+uxρ+ρxu=0,tu+uxu+ρ1ρ(pe+pi)xρ+ρ1εepexεe+ρ1εipixεi=0,tεe+uxεe+ρ1epexu=ρ1eνei(TiTe),tεi+uxεi+ρ1ipixu=ρ1iνei(TeTi). (5)

    The matrix of the system (5) writes

    A(V)=uI4+(0ρ00ρ1ρ(pe+pi)0ρ1εepeρ1εipi0ρ1epe000ρ1ipi00). (6)

    We get 4 eigenvalues

    λ1=ua,λ2=λ3=u,λ4=u+a,

    where

    a=β=e,iγβ(γβ1)cβεβ. (7)

    The value of a given by (7) corresponds to the global sound velocity which yields for the classical Euler system γp/ρ. The eigenvectors associated to the eigenvalue u are equal to

    r2=(00(γi1)ci(γe1)ce),r3=(ρ0εeεi).

    This system is then hyperbolic diagonalisable. The fields 2 and 3 are linearly degenerate. Consider now the fields 1 and 4. The eigenvectors are

    r1(V)=(ρa(γe1)εe(γi1)εi),r4(V)=(ρa(γe1)εe(γi1)εi)

    and

    λ1(V)r1(V)=λ4(V)r4(V)=12aβ=e,i(γβ(γβ1)(γβ+1)cβεβ)>0. (8)

    Hence the fields 1 and 4 are genuinely nonlinear.

    In this section, we generalize the BGK model considered in [3] to a polyatomic setting with a continuous energy variable. We firstly precise the notations and next introduce the BGK model that is proved to satisfy the right conservation properties and an H theorem.

    Kinetic models for a mixture of two polyatomic gases are described by two distribution functions fβ of species β depending on time tR+, space xR3, velocity vR3 and on internal energy IR+.

    Hydrodynamic quantities of species β are defined for αβ0 by

    nβ=R3×R+fβIαβdvdI,uβ=1nβR3×R+vfβIαβdvdI,Eβ=R3×R+(mβv22+I)fβIαβdvdI,

    and

    Eβ=12ρβu2β+(52+αβ)nβkBTβ,β=e,i.

    Velocities and temperatures of the mixture u and T are defined by

    u=1ρ(ρeue+ρiui), (9)
    T=(52+αe)nekBTe+(52+αi)nikBTi+12ρeu2e+12ρiu2i12ρu2(52+αe)nekB+(52+αi)nikB. (10)

    The parameters αe and αi are related to γe and γi by the formula ([26])

    γe=152+αe+1,γi=152+αi+1.

    For example in the diatomic case, we have αe=αi=0. The values αe=αi=1 correspond to γe=γi=5/3 which is a monoatomic mixture.

    Define the entropy of the mixture by

    H(fe,fi)=Hs(fe)+Hs(fi),withHs(fβ)=R3×R+(fβln(fβ)fβ)IαβdvdI (11)

    and the entropy flux by

    Φ(fe,fi)=Φs(fe)+Φs(fi),withΦs(fβ)=R3×R+v(fβln(fβ)fβ)IαβdvdI. (12)

    In this section we consider the following kinetic model for β{e;i},

    tfβ(t,x,v,I)+vxfβ(t,x,v,I)+qβmβEvfβ(t,x,v,I)=1τβ(Mβfβ(t,x,v,I))+1τβδ(¯Mβ(fe,fi)fβ(t,x,v,I)), (13)

    with

    Mβ=nβ(2πkBmβTβ)321Qβ(Tβ)exp((vuβ)22kBTβmβIkBTβ), (14)
    ¯Mβ(fe,fi)=nβ(2πkBmβT#)321Qβ(T#)exp((vu#)22kBT#mβIkBT#), (15)

    where

    Qβ(T)=+0Iαβexp(IkBT)dI,αβ0.

    As suggested in [32] and developed in [3] the definition of u# and T# as

    u#=1τeiρeue+1τieρiui1τeiρe+1τieρi (16)

    and

    T#=1τei(52+αe)nekBTe+1τie(52+αi)nikBTi1τei(52+αe)nekB+1τie(52+αi)nikB+121τeiρeu2e+1τieρiu2i(1τeiρe+1τieρi)(u#)21τei(52+αe)nekB+1τie(52+αi)nikB (17)

    gives the conservations of impulsion and total energy for the model (13, 14, 15). These definitions allow in particular to consider a model such that the molecular mass ratios of the species are different. This situation corresponds to the case τeiτie.

    The model (13, 14, 15, 16, 17) is coupled to Ampère and Poisson equations through the electric field E as

    tE=jε0, (18)
    xE=¯ρε0. (19)

    j represents the plama current, ¯ρ the total charge and ε0 is the permittivity. j and ¯ρ are defined by

    ¯ρ=R3×R+(qefeIαe+qifiIαi)dvdI=neqe+niqi, (20)
    j=R3×R+v(qefeIαe+qifiIαi)dvdI=neqeue+niqiui. (21)

    When one of the two components is monoatomic the model has to be slightly modified. In [8], the authors study for moments systems the link beween polyatomic and monoatomic models. The connexion between monoatomic and polyatomic models can be made by passing to the limit in some parameters. When the two components are monoatomic we refer to [3].

    In the case of one monoatomic component and one polyatomic component the model has to be clearly written. Consider for example that electrons are monoatomic whereas ions remain polyatomic. In that case, the model is described with the distributions fe(t,x,v) and fi(t,x,v,I). In that case, the model reads

    tfe(t,x,v)+vxfe(t,x,v)+qemeEvfe(t,x,v)=1τe(Mefe(t,x,v))+1τei(¯Me(fe,fi)fe(t,x,v)), (22)
    tfi(t,x,v,I)+vxfi(t,x,v,I)+qimiEvfi(t,x,v,I)=1τi(Mifi(t,x,v,I))+1τie(¯Mi(fe,fi)fi(t,x,v,I)), (23)

    with

    Me=ne(2πkBmeTe)32exp((vue)22kBTeme), (24)
    ¯Me(fe,fi)=ne(2πkBmαT#)32exp((vu#)22kBT#me). (25)

    The hydrodynamic quantities of the monoatomic species are computed as

    ne=R3fedv,ue=1neR3vfedv,Ee=R3mev22fedv.

    The total impulsion ρu and the total energy E for the system reads

    ρu=R3mevfedv+R3×R+mivfiIαidvdI,E=R3mev22fedv+R3×R+(miv22+I)fiIαidvdI.

    In that case u# is given by (16) and T# becomes

    T#=1τei32nekBTe+1τie(52+αβ)nikBTi1τei32nekB+1τie(52+αi)nikB+121τeiρeu2e+1τieρiu2i(1τeiρe+1τieρi)(u#)21τei32nekB+1τie(52+αi)nikB. (26)

    As previously, u# and T# are defined in such a way that the conservation of impulsion and total energy are satisfied. Mi and ¯Mi(fe,fi) are given by (14, 15).

    In the present case the definitions of j and ¯ρ given in (20, 21) become

    ¯ρ=R3qefedv+R3×R+qifiIαidvdI,j=R3vqefedv+R3×R+vqifiIαidvdI

    and the electric field is still given by (18, 19). For the sake of conciseness, we continue the presentation only for a two polyatomic species mixture. However, the following steps can be easily adapted when one of the components is monoatomic.

    Proposition 1. The model (13, 14, 15, 16, 17) conservesthe mass per species, the total impulsion and the total energy.

    The proof is straighforward and based as in [3] on the definition of the fictitious quantities (16, 17).

    Theorem 2.1. The model (13, 14, 15, 16, 17) satisfiesan H theorem. The model satisfies the entropic inequality

    1τeR3×R+(Me(fe)fe)ln(fe)IαedvdI+1τiR3×R+(Mi(fi)fi)ln(fi)IαidvdI+1τeiR3×R+(¯Me(fe,fi)fe)ln(fe)IαedvdI+1τieR3×R+(¯Mi(fe,fi)fi)ln(fi)IαidvdI0.

    The equality holds in the above equation if and only if there exists(ne,ni,u,T) R2+×R3×R+ such that

    Mβ=nβ(2πkBmβT)321Qβ(T)exp((vu)22kBTmβIkBT),β{e,i}.

    An important feature of the polyatomic model (13, 14, 15, 16, 17) is that it satisfies an entropy dissipation property that is compatible with the macroscopic one. The entropy dissipation property has already been obtained in [3] for the system (1), starting directly from the fluid system. In the present case, we are able to show that the entropy of the system (1) is compatible with the Boltzmann entropy, see subsection 3.3.

    Suppose that the system (13, 14, 15, 16, 17) is space homogeneous in two directions. So we assume that the system is even in v2 and v3. Hence, the distribution function fβ of species β depends on time tR+, space xR, velocity v=(v1,v2,v3) and on the energy variable IR+. The model (13, 14, 15, 16, 17) can be rewritten in this case

    {tfβ+v1xfβ+qβmβEv1fβ=1ε(Mβfβ)+1τβδ(¯Mβfβ),βδtE=jε2,xE=¯ρε2, (27)

    where ε is a nonnegative parameter proportional to the Knudsen number. In that case, Maxwellian distributions (14, 15) become

    Mβ=nβ(2πkBmβTβ)321Qβ(Tβ)exp((v1uβ)2+v22+v332kBTβmβIkBTβ), (28)
    ¯Mβ(fe,fi)=nβ(2πkBmβT#)321Qβ(T#)exp((v1u#β)2+v22+v232kBT#mβIkBT#) (29)

    where u# and T# are defined in (16) and (17).

    Proposition 2. The system (27, 28, 29, 16, 17) convergesformally to the nonconservative bitemperature Euler system where E is given by generalized Ohm's law

    1ρexpe1ρixpi=(neqeρeniqiρi)E=ρρeρineqeE=ρρeρiniqiE (30)

    and

    νei=kB(52+αe)(52+αi)neniτie(52+αe)ne+τei(52+αi)ni. (31)

    Proof. Performing a Chapman-Enskog expansion, it comes that each component of the species is expanded as

    fβ=Mβ+εgβ,β{e,i} (32)

    so that

    R3×R+gβIαβdvdI=0,R3×R+v1gβIαβdvdI=0, (33)
    R3×R+(12mβv2+I)gβIαβdvdI=0. (34)

    Moreover, Gauss equation in (27) implies that niqi=neqe+O(ε2). So, ne=Zni+O(ε2). Ampère equation then gives ue=ui+O(ε2)=u. Plugging (32) into the first equation of (27) leads to

    tMβ+v1xMβ+qβmβEv1Mβ=gβ+1τβδ(¯MβMβ)+O(ε) (35)

    for {β;δ}{e;i},βδ.

    Mass conservation equation is obtained by integrating w.r.t v and I. The equation of the conservation of the impulsion is then recovered by taking the second moment of (35). Moreover, proceeding as in [3], we get Ohm's law (30). Next the energy equation on the electrons writes

    R3×R+(tMe+v1xMe)(12mev2+I)IαedvdI+R3×R+qemβEv1Me(12mev2+I)IαedvdI=1τeiR3×R+(¯MeMe)(12mev2+I)IαedvdI.

    Moreover a direct computation gives

    R3×R+(¯MeMe)(12mev2+I)IαedvdI=(52+αe)nekB(T#Te).

    So, according to the relation (10) defining T, we get

    R3×R+(¯MeMe)(12mev2+I)IαedvdI=(52+αe)ne(52+αi)niτie(52+αe)ne+τei(52+αi)nikB(TiTe)

    and νei is given by (31).

    As proved in [3], the system (1) owns a dissipative entropy-entropy flux pair

    η=ηe+ηi,Q=uη (36)

    where

    ηβ=ρβmβ(γβ1)[ln((γβ1)ρβεβργββ)+C],β=e,i. (37)

    Here, C is a nonegative constant. With the same method as in [3] we can prove

    Theorem 3.1. Let (η,Q) be defined by (36)-(37). η is a strictly convex dissipative entropy for system (1) and Q is the related entropy flux. More precisely, any smooth solution of the system satisfies the following equality:

    tη(U)+xQ(U)=νeikBTiTe(TiTe)2. (38)

    If U is a weak solution of system (1) obtained as the hydrodynamic limit of the kinetic model, then it satisfies the following inequality

    tη(U)+xQ(U)νeikBTiTe(TiTe)2. (39)

    As in [3], the inequality (39) is obtained by using the proof of the H theorem (Theorem 2.1).

    We compute contact discontinuities and rarefaction waves for system (1) in the homogeneous case νei=0. As u is a double eigenvalue we have contact discontinuities of two kinds. Moreover the computation of the rarefaction waves is different from the one for the classical Euler system.

    A contact discontinuity is a weak solution U=(ρ,ρu,Ee,Ei) of (1) such that u is a constant and

    ρ(x,t)=|ρLifx<ut,ρRifx>utEβ(x,t)=|Eβ,Lifx<ut,Eβ,Rifx>utβ=e,i.

    Here, ρL, ρR, Eβ,L, Eβ,R are constant. In that case the homogeneous system related to (1) can be written under the conservative form:

    {tρ+x(ρu)=0,t(ρu)+x(ρu2+pe+pi)=0,tEe+x(u(Ee+ce(pe+pi)))=0,tEi+x(u(Ei+ci(pe+pi)))=0. (40)

    Rankine-Hugoniot jump conditions are

    [u]=0,[pe+pi]=0.

    One can realize those conditions by taking [ρ]=0 or not. In the first case, nontrivial contact discontinuities are obtained with nonequal left and right values of the ionic and electronic pressures. This case is specific to the bitemperature model. The case [ρ]0 appears also in contact discontinuities for the 3×3 monotemperature Euler system.

    For contact discontinuities there is no entropy dissipation:

    tη(U)+xQ(U)=0.

    A rarefaction wave is a selfsimilar continuous, piecewise C1 solution of (1) with νei=0. As we look for a smooth solution, we may use the variable V=(ρ,u,εe,εi). The rarefaction waves are given by solutions V(x,t)=W(xt) of the homogeneous system related to (5) that is, denoting y=x/t:

    {yI+A(W(y))}W(y)=0 (41)

    where A is given by (6). Rarefaction waves are closely related to the integral curves of the eigenvectors of A, as soon as the fields are genuinely nonlinear, see [18] for example. Let us consider the eigenvalue λ+=u+a with a given by (7), with the eigenvector r4 satisfying (8). We solve the ODS

    Φ(ξ)=r4(Φ(ξ)),Φ(0)=WL.

    We set WR=Φ(ξ0) with ξ0>0 and Ψ(ξ)=λ+(Φ(ξ)). Ψ is an increasing monotone function. We set

    W(y)=|WLsiyλ+(WL),Φ(Ψ1(y))ifλ+(WL)yλ+(WR),WRsiyλ+(WR).

    We have WC0(R) and (41). Hence we have to solve for ξ>0:

    {ρ(ξ)=ρ(ξ),u(ξ)=a(ξ),εβ(ξ)=(γβ1)εβ(ξ),β=e,i.(ρ(0),u(0),εe(0),εi(0))=WL.

    We find

    {ρ(ξ)=ρLeξ,u(ξ)=βγβ(γβ1)cβεβ,Le(γβ1)ξ,εβ(ξ)=εβ,Le(γβ1)ξ,β=e,i. (42)

    As

    Tβ=(γβ1)mβkBεβandpβ=(γβ1)ρβεβ,

    we have also:

    Tβ(ξ)=Tβ,Le(γβ1)ξetpβ(ξ)=pβ,Leγβξ,β=e,i.

    If γiγe we cannot parametrize the rarefaction curves by the pressure as one does for the monotemperature Euler system. Hence from (42), the rarefaction curves can be parametrized as follows

    ξ=ln(ρρL),εβ=εβ,L(ρρL)γβ1.

    So

    Tβ=Tβ,L(ρρL)γβ1.

    We retrieve the fact that ρR>ρL. We have also pβ>pβ,L and the specific entropy by species is constant. It is defined by

    Sβ=pβργββ.

    For all ξ:

    Sβ(ξ)=Sβ(0),β=e,i

    hence

    Sβ,L=Sβ,R,β=e,i.

    In this section we derive a numerical scheme starting from a semi-discretization of the kinetic model. In [3], this approach has been developed for a monoatomic gas mixture. We follow the lines of [3] and obtain the scheme for the polyatomic case. For the sake of completeness we give the details.

    First we recall that for Pβ defined by

    Pβ(fβ)=R3×R+(mβmβv1(mβv22+I))fβIαβdvdI (43)

    one has

    Pβ(fnβ)=Unβ,j,Pβ(v1Mβ(Unβ,j))=Fβ(Unβ,j), (44)

    where Fβ is the flux of 3×3 Euler equations with γβ pressure law.

    The spatial discretisation is defined by the step Δx and the cells Cj=]xj12,xj+12[. We consider that Δx is constant whereas the time step Δt: t0=0, tn+1=tn+Δt can be variable.

    We use a finite volume approach: for any unknown V(x,t), we look for the approximation Vnj of the average of V at time tn on the cell Cj. Suppose that Un is known (n0).

    First step. For β=e,i we set

    ρnβ,j=cβρnj (45)

    and Unβ,j=(ρnβ,j,ρnβ,junj,Eβ,j). fne(v,I), fni(v,I) are computed by projection on the maxwellian states:

    j,fnβ,j(v,I)=Mβ(Unβ,j,v,I),β=e,i. (46)

    Second step. We solve the transport equations tfβ+v1xfβ=0 by using a numerical flux hβ,j+12(v,I)=hβ(fβ,j(v,I),fβ,j+1(v,I),v,I). Here we choose the HLL type flux

    hβ(f,g,v,I)=v1(λ2λ2λ1f(v,I)λ1λ2λ1g(v,I))+λ1λ2λ2λ1(g(v,I)f(v,I)), (47)

    where λ10λ2 are distinct real constants to be fixed. We set for all (v,I)

    fn+12β,j(v,I)=fnβ,j(v,I)ΔtΔx(hnβ,j+12(v,I)hnβ,j12(v,I)).

    Remark 1. This scheme is stable for λ1v1λ2 and appropriate CFL condition. As we are going to integrate this formula w.r.t. v1R, we have no chance to prove stability at this level. The resulting macroscopic scheme owns however stability properties that can be proven by a different way, as we shall show at the end of this section.

    We apply Pe on fe, Pi on fi, and obtain Un+12e,j and Un+12i,j:

    Un+12β,j=Pβ(fn+12β,j)=(ρn+12β,jρn+12β,jun+12β,jEn+12β,j)

    Denoting for β=e,i

    Fβ,j+12=Fβ(Uβ,j,Uβ,j+1),Fβ(Uβ,Vβ)=Pβ(hβ(Mβ(Uβ),Mβ(Vβ),)), (48)

    we have by using (44), (46) and (47)

    Fβ(Uβ,Vβ)=λ2λ2λ1Fβ(Uβ)λ1λ2λ1Fβ(Vβ)+λ1λ2λ2λ1(VβUβ),

    and

    Un+12β,j=Unβ,jΔtΔx(Fnβ,j+12Fnβ,j12).

    Hence Un+12β,j is computed by the HLL scheme (λ10λ2).

    Remark 2. We could have chosen the upwind scheme to approximate the transport equations, but it is more difficult to integrate the formulas w.r.t. v1 in this case, because the numerical flux depends on the sign of v1.

    It is easy to prove the following result.

    Proposition 3. For β=e,i

    ρn+12β,j=cβρn+12j (49)

    where ρ is defined in (2).

    Third step. We take into account the force terms and the source terms. For all jZ, α,β{e,i} and βα, we define

    fn+34β,j(v,I)=fn+12β,jΔtqβmβEn+1jvfn+34β,j+Δtτβδ(¯Mβ(fn+34e,j,fn+34i,j)fn+34β,j) (50)

    and

    Un+1β,j=Pβ(fn+34β,j). (51)

    One obtains the following equations for β=e,i:

    {ρn+1β,j=cβρn+12jρn+1β,jun+1β,j=ρnβ,junβ,jΔtΔx(Fnβ,j+12,2Fnβ,j12,2)ΔtqβmβEn+1jρn+1β,jEn+1β,j=Enβ,jΔtΔx(Fnβ,j+12,3Fnβ,j12,3)En+1jun+1β,jΔtqβmβρn+1β,j+Δtνei(Tn+1β,jTn+1β,j),ββ. (52)

    Subsequently, it is necessary to ensure that the quasineutrality constraints are satisfied, which correspond to Maxwell-Gauss and Maxwell-Ampère equations in the limit ε0:

    qemeρn+1e,j+qimiρn+1i,j=0,qemeρn+1e,jun+1e,j+qimiρn+1i,jun+1i,j=0.

    By proposition 3 the first condition is satisfied and ρn+1j=ρn+12j. The second condition is equivalent to un+1i,j=un+1e,j=un+1j. As a consequence if Un+1j=(ρn+1j,ρn+1jun+1j,En+1e,j,En+1i,j) then Un+1e,j and Un+1i,j satisfy (45), so our notation is consistent. By applying these properties to (52) for β=e,i, one gets:

    ceρn+1jun+1j=ceρnjunjΔtΔx(Fne,j+12,2Fne,j12,2)ΔtqemeEn+1jρn+1e,j,ciρn+1jun+1j=ciρnjunjΔtΔx(Fni,j+12,2Fi,nj12,2)ΔtqimiEn+1jρn+1i,j.

    Hence, by multiplying the first equation by ci and the second equation by ce, and then by substracting one to the other, one obtains, analoguously to the continuous case, the discrete generalized Ohm law:

    En+1jqimiρn+1i,j=En+1jqemeρn+1e,j=1Δx(δnj+12δnj12),

    where

    δnj+12=ciFne,j+12,2+ceFni,j+12,2. (53)

    Finally, defining Fj+12 by

    Fj+12=(Fe,j+12,1+Fi,j+12,1Fe,j+12,2+Fi,j+12,2Fe,j+12,3Fi,j+12,3), (54)

    we get a scheme that is consistent with the Euler system (1):

    Proposition 4. For any n0 if Un={Unj}jZ is the approximate solution of the system (1) at time tn, we set

    Unβ,j=(cβρnj,cβρnjunj,Eβ),β=e,i. (55)

    The kinetic flux hβ is defined by (47). The numerical fluxes Fβ,j+12, δj+12 and Fj+12 are then defined by (48), (53), (54). The approximate solution at time tn+1 is defined by the implicit scheme

    {ρn+1j=ρnjΔtΔx(Fnj+12,1Fnj12,1),ρn+1jun+1j=ρnjunjΔtΔx(Fnj+12,2Fnj12,2),En+1e,j=Ene,jΔtΔx(Fne,j+12,3Fne,j12,3)un+1jΔtΔx(δnj+12δnj12)+Δtνei(Tn+1i,jTn+1e,j),En+1i,j=Eni,jΔtΔx(Fni,j+12,3Fni,j12,3)+un+1jΔtΔx(δnj+12δnj12)Δtνei(Tn+1i,jTn+1e,j). (56)

    Remark 3. The scheme is implicit but easy to implement because the first two equations of (56) give ρn+1j and un+1j explicitly and the two last equations can be expressed as a linear 2×2 system for the unknown Te, Ti.

    This numerical scheme is the same as the one obtained in section 3.2 of [3] and for which a second order two-dimensional extension is presented in [6], (λ1λ20). However, in these two articles, the scheme was obtained by a very different method involving models developed in [7]. Those models are formally like discrete BGK equations but are not based on a physically realistic kinetic representation of the equations. The novelty here is that our polyatomic BGK model gives a physical justification to it, including the case γeγi.

    We recall that stability and entropy properties where proved in these references. In particular, a discrete entropy inequality holds under appropriate choices of λ1 and λ2 and a CFL condition using the sound velocity of each species. These theoritical conditions give rise however to too much numerical diffusion. We replace them by using the global sound velocity defined in (7) (see [6]):

    λ1min(ua)max(u+a)λ2,max(|λ1|,|λ2|)ΔtΔx1. (57)

    As pointed out in section 5, the presented scheme has already been tested in [3] and [6]. The aim of this section is to investigate more precisely the polyatomic case with two tests where γe=53 and γi=75. We make use of the second order extension in space and time with affine reconstruction and Heun scheme developed in [6].

    We solve the bitemperature Euler system with the following Riemann data, whose orders of magnitude are those encountered in physical situations:

    ρ=1,ρ+=ρ,u=100000,u+=u,Te,=2.3×107,Te,+=Te,,Ti,=2.3×106,Ti,+=Ti,.

    We set Z=1 and a final computation time t=4.0901×107sec. The numerical simulations are performed on the interval [0,1] with 2000 cells. The values of λ1 and λ2 are computed at every time-step with the condition (57). We set νei=0 so that the solution consists of two rarefaction waves propagating to left and right, the contact discontinuity being trivial. In order to determine the analytical solution, denoting (ρ,u,εe,εi) the intermediate state, one has to find ξ>0 such that:

    {ρ=ρ±eξu=u+ξ0a(s)ds=u+ξ0a(s)dsεβ=e(γβ1)ξεβ,±,β=e,i,

    with

    a(s)=(βγβ(γβ1)cβεβ,±e(γβ1)s)1/2.

    Hence we find numerically ξ>0 such that

    u+=ξ0a(s)ds.

    The numerical results are depicted in Figures 1 and 2. We compare the exact and computed results for density, velocity and temperatures. As already observed in [3] for γe=γi=5/3, a peak of ionic temperature happens at x=1/2. This peak is similar as the one observed classically for the monotemperature 3×3 Euler system.

    Figure 1. 

    Double rarefaction. Left: density. Right: velocity

    .
    Figure 2. 

    Double rarefaction. Left: electronic temperature. Right : ionic temperature

    .

    In this test case, we consider an implosion problem, introduced in [28] in the monoatomic case γe=γi=53. We compared our approach with the conservative one of this paper in [6]. Here we set γe=53 and γi=75 and keep the other parameters unchanged, that is: the physical domain is the square [1,1]×[1,1]. Initial data for this Riemann problem is as follows: ρ=1 kg.m3, u=0 m.s1 and temperatures are given by:

    Te(x1,x2,0)=2,3×106K,Ti(x1,x2,0)=1.7406×106Kif(x1)2+(x2)2<14,Te(x1,x2,0)=2,3×107K,Ti(x1,x2,0)=1.7406×107Kotherwise.

    The relaxation frequency νei is chosen realistically, according to the formulae given by the NRL formulary [33]. Due to stiffness of the source term one has Te=Ti very quickly.

    Thanks to symmetry properties of the problem, it is only necessary to solve it on the quarter domain [0,1]×[0,1], equipped with suitable boundary conditions.

    On Figure 3, are given the isovalues of the total density and of the electronic temperature at time t=4.0901×107 sec. We can see that the qualitative behaviour is the same as in the monoatomic case, see Figure 7 of [6]. Here also one can observe a shock propagating towards the center of the domain, a contact discontinuity, and a rarefaction wave propagating to the exterior. However the values of densities and temperatures are different.

    Figure 3. 

    Total density (left) and electronic temperature (right) at time t=4.0901×107s for an implosion test case with νei given by the NRL formulary with a grid of 500 by 500 points

    .

    When the shock front reaches the center, a peak of density occurs. This peak occurs at time t=8.798×107 sec. in the monoatomic case, while it occurs here at time t=9.2×107 sec., see Figure 4. The temperatures are also maximal at the center when this peak occurs, as shown in Figure 5.

    Figure 4. 

    Implosion test case with νei given by the NRL formulary with a grid of 500 by 500 points. Density along the first bisector at 4 different times: the peak occurs for t=9.2×107 sec

    .
    Figure 5. 

    Implosion test case with νei given by the NRL formulary with a grid of 500 by 500 points. Left: isovalues of the density when the peak occurs. Right: isovalues of the electronic and ionic temperatures when the peak occurs

    .


    [1] A comment on the computation of non-conservative products. J. Comput. Phys. (2010) 229: 2759-2763.
    [2] Entropy condition for the ES BGK model of Boltzmann equation for mono and polyatomic gases. Eur. J. Mech. B Fluids (2000) 19: 813-830.
    [3] Modelling and numerical approximation for the nonconservative bitemperature Euler model. ESAIM Math. Model. Numer. Anal. (2018) 52: 1353-1383.
    [4] A viscous approximation of the bitemperature Euler system. Comm. Math. Sci. (2019) 17: 1135-1147.
    [5] Global existence of smooth solutions for a non-conservative bitemperature Euler model. SIAM J. Math. Anal. (2021) 53: 1886-1907.
    [6] A discrete velocity numerical scheme for the two-dimensional bitemperature Euler system. SIAM J. Numer. Anal. (2022) 60: 28-51.
    [7] Discrete kinetic schemes for multidimensional systems of conservation laws. SIAM J. Numer. Anal. (2000) 37: 1973-2004.
    [8] Monoatomic rarefied gas as a singular limit of polyatomic gas in extended thermodynamics. Physics letter A (2013) 337: 2136-2140.
    [9] On the Chapman-Enskog asymptotics of mixture of monoatomic and polyatomic rarefied gases. Kin. Rel. Mod. (2018) 11: 821-858.
    [10] BGK polyatomic model for rarefied flows. J. Sci. Comput. (2019) 78: 1893-1916.
    [11]

    M. Bisi, R. Monaco and A. J. Soares, A BGK model for reactive mixtures of polyatomic gases with continuous internal energy, J. Phys. A, 51 (2018), 125501, 29 pp.

    [12] Dynamical pressure in a polyatomic gas: Interplay between kinetic theory and Extended Thermodynamics. Kinet. Relat. Models (2018) 11: 71-95.
    [13] On a kinetic BGK model for slow chemical reactions. Kinet. Relat. Models (2011) 4: 153-167.
    [14]

    M. Bisi and R. Travaglini, A polyatomic model for mixtures for monoatomic and polyatomic gases, Phys. A, 547 (2020), 124441, 18 pp.

    [15] A general consistent BGK model for gas mixtures. Kinet. Relat. Models (2018) 11: 1377-1393.
    [16] Statistical collision model for Monte-Carlo simulation of polyatomic mixtures. Journ. Comput. Phys. (1975) 18: 405-420.
    [17] Microreversible collisions for polyatomic gases and Boltzmann's theorem. Eur. Journ. Fluid Mech. (1994) 13: 237-254.
    [18] (2000) Hyperbolic Systems of Conservation Laws. Oxford: Oxford Lecture Series in Mathematics and its Applications, 20. Oxford University Press.
    [19] An Ellipsoidal Statistical Model for a monoatomic and polyatomic gas mixture. Comm. Math. Sci. (2021) 19: 2177-2194.
    [20] A kinetic approach of the bi-temperature Euler model. Kinet. Relat. Models (2020) 13: 33-61.
    [21] On the Ellipsoidal Statistical Model for polyatomic gases. Cont. Mech. Thermodyn (2009) 20: 489-508.
    [22] Navier-Stokes equations with several independent pressure laws and explicit predictor-corrector schemes. Numer. Math. (2005) 101: 451-478.
    [23] Numerical methods for weakly ionized gas. Astrophysics and Space Science (1998) 260: 15-27.
    [24] Definition and weak stability of nonconservative products. J. Math. Pures et Appl. (1995) 74: 483-548.
    [25] Sur un modèle de type Borgnakke-Larsen conduisant à des lois d'énergie non-linéaires en température pour les gas parfaits polyatomiques. Ann. Fac. Sci. Toulouse Math. (1997) 6: 257-262.
    [26] A kinetic model allowing to obtain the energy law of polytropic gases in the presence of chemical reactions. Eur. J. Mech. B Fluids (2005) 24: 219-236.
    [27] The kinetic equilibrium regime. Physica A (1998) 260: 49-72.
    [28]

    E. Estibals, H. Guillard and A. Sangam, Derivation and numerical approximation of two-temperature Euler plasma model, J. Comput. Phys., 444 (2021), 110565, 48 pp.

    [29] Poiseuille flow and thermal transpiration of a rarefied polyatomic gas through a circular tube with applications to microflows. Boll. Unione Mat. Ital. (2011) 4: 19-46.
    [30]

    V. Giovangigli, Multicomponent Flow Modeling. Modeling and Simulation in Science, Engineering and Technology, Birkhäuser Boston, Inc., Boston, MA, 1999.

    [31] Kinetic theory of plasmas. Maths models and methods in the Appl. Sci. (2009) 19: 527-599.
    [32] Improved Bhatnagar-Gross-Krook model of electron-ion collisions. Phys.Fluids (1973) 16: 2022-2023.
    [33]

    J. D. Huba, NRL Plasma Formulary, Revised 2013 version, NRL.

    [34] A consistent kinetic model for a two-component mixture of polyatomic molecules. Comm. in Math. Sci. (2019) 17: 149-173.
    [35] Shock wave structure in polyatomic gases: Numerical analysis using a model Boltzmann equation. AIP Conf. Proc. (2016) 1786: 180004.
    [36] Numerical methods for nonconservative hyperbolic systems: A theoretical framework. SIAM J. Numer. Anal. (2006) 44: 300-321.
    [37] A kinetic scheme for the Saint-Venant system with a source term. Calcolo (2001) 38: 201-231.
    [38] Fluid modeling for the Knudsen compressor: Case of polyatomic gases. Kinet. Relat. Models (2010) 3: 353-372.
    [39]

    Q. Wargnier, S. Faure, S. B. Graille, T. Magin and M. Massot, Numerical treatment of the nonconservative product in a multiscale fluid model for plasmas in thermal nonequilibrium: Application to solar physics, SIAM J. Sci. Comput., 42 (2020), B492–B519.

    [40] (1966) Physics of Shock Waves and High-Temperature Hydrodynamic Phenomena.Academic Press.
  • Reader Comments
  • © 2022 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(919) PDF downloads(180) Cited by(0)

Figures and Tables

Figures(5)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog