Analysis and Simulation for a Model of Electron Impact Excitation/Deexcitation and Ionization/Recombination ˚ Bokai Yan:, Russel E. Caisch:, Farzin Barekat: and Jean-Luc Cambier; April 10, 2015 Abstract This paper describes a kinetic model and a corresponding Monte Carlo simulation method for excitation/deexcitation and ionization/recombination by electron impact in a plasma free of external elds. The atoms and ions in the plasma are represented by continuum densities and the electrons by a particle distribution. A Boltzmann-type equation is formulated and a corresponding H-theorem is formally derived. An ecient Monte Carlo method is developed for an idealized analytic model of the excitation and ionization collision cross sections. To accelerate the simulation, the reduced rejection method and binary search method are used to overcome the singular rate in the recombination process. Numerical results are presented to demonstrate the eciency of the method on spatially homogeneous problems. The evolution of the electron distribution function and atomic states is studied, revealing the possibility under certain circumstances of system relaxation towards stationary states that are not the equilibrium states, a potential non-ergodic behavior. Keywords. Monte Carlo Method, Excitation-deexcitation, Ionization-recombination, Boltzmann equation, H theorem, Singular reaction rates. Distribution A: Approved for public release; distribution unlimited. This research was supported by AFOSR and AFRL. The work of Barekat, Caisch and Yan was partially supported by AFOSR grant FA9550-14-1-0283. : Mathematics Department, University of California at Los Angeles, Los Angeles, CA 90095-1555, USA. [email protected], [email protected], [email protected] ; Air Force Research Laboratory, Edwards AFB, CA 93524, USA. [email protected] ˚ 1 Contents 1 Introduction 4 2 Formulation 5 2.1 The Distribution Function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 2.2 Interaction Types and Rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7 2.3 The Principle of Detailed Balance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 2.4 The Boltzmann Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 2.4.1 Ionization and Recombination . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 2.4.2 Excitation and Deexcitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 3 Entropy and the H-Theorem 11 3.1 Entropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 3.2 Relation to H-theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 3.3 H-Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 3.3.1 Ionization and Recombination . . . . . . . . . . . . . . . . . . . . . . . . . . . 17 3.3.2 Excitation and Deexcitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18 3.3.3 Full Kinetic Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 4 Idealized Energetic Cross Sections 20 4.1 Ionization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Recombination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 4.3 Excitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 4.4 Deexcitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 5 Monte Carlo Particle Method 21 22 5.1 The Framework of Algorithm . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 5.2 Sample Electrons from a Singular Distribution . . . . . . . . . . . . . . . . . . . . . . 23 5.2.1 Acceptance/Rejection Method . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 5.2.2 Direct Search Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 5.2.3 Binary Search Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 5.2.4 Reduced Rejection Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 5.3 Large Scale Variation in the Rates . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 5.4 Particle/Continuum Representation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26 5.5 A No Time Counter DSMC method 26 . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 6 Computational Results 29 6.1 Equilibrium initial data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 6.2 Non-equilibrium initial data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 6.3 The stationary states without ionization/recombination . . . . . . . . . . . . . . . . 38 6.4 Comparison of dierent methods for sampling electrons in recombination . . . . . . . 38 6.5 The large scale variation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 7 Conclusions and Future Work 42 Appendix A Proof of Lemma 2.1 45 Appendix B Proof of Lemma 3.6 46 Appendix C Proof of Proposition 3.1 47 Appendix D Proof of Lemma 3.8 and Theorem 3.9 49 Appendix E Algorithm for binary search method 52 E.1 Padding by zeros . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 E.2 Building a table . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 E.3 Sampling a particle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 E.4 Update the table when some vk changes 3 . . . . . . . . . . . . . . . . . . . . . . . . . 53 1 Introduction Simulations of plasma dynamics can be computationally challenging due to the large number of relevant physical processes and the wide range of space/time scales over which they operate. This is particularly true when accounting for inelastic collisional processes such as excitation/deexcitation and ionization/recombination in an atomic plasma, due to the large number of electronic states and reaction pathways. The relative signicance of collisional interactions is determined by examining the strengths of the various terms of the Boltzmann equation for the distribution function Bt f ` v¨∇x f ` a¨∇v f “ Qpf q where Q f: (1.1) eq being the equilibrium, Maxwellian is the collision operator, with a stationary state f distribution. The left-hand-side of (1.1) is the streaming operator in physical (x) and velocity (v) apE, Bq is the acceleration due to the electric (E ) and magnetic (B ) elds. Thus, the number Kn “ Lc {L0 , in which Lc is the mean free path between collisions and L0 is a space, and Knudsen characteristic length scale of the gradient operator, determines the relative strength of collisions versus advection. In the limit of small Knudsen number and weak elds, the energy distribution function f for each of the particle species is close to an Maxwellian equilibrium, and the evolution of the plasma can be described by uid-like equations, in which individual particle interactions do not need to be resolved. On the other hand, for Kn Á 1 and/or large elds, the distribution f can be far from equilibrium, in which case particle representations and a Monte Carlo simulation procedure are more appropriate. A variety of physical and numerical models have been used to solve for the Boltzmann equation (mostly designed for transport studies) using uid approximations (e.g. [1]), spherical harmonics expansions (e.g. [24]), or Monte-Carlo (e.g. [5, 6]). In this work, we do not intend to provide yet another analysis tool for non-equilibrium plasma, but examine in detail the algorithms for a fully kinetic, Monte Carlo simulation of electron-induced inelastic collisions for excitation, ionization, and their reverse processes (deexcitation and recombination) for a dilute atomic plasma. In particular, we develop an ecient procedure for dealing with divergent recombination rates at low energy, provide a proof of the H-theorem for inelastic processes, and examine the properties of the stationary states. For simplicity, we consider the simplest atomic species, Hydrogen, using a Bohr model of the Atomic State Distribution Function (ASDF); this is sucient to examine the eects of multiple scales in the energies exchanged, as well as timescales of reactions, while focusing on some mathematical and algorithmic aspects of the solution. For the same reason, other physical processes such as radiative transitions, atomic (heavy-particle) collisions, charge-exchange and elastic (Coulomb) collisions are also not considered. Note that we are also solving only for the local inelastic collision operator, and other eects due to transport in physical and velocity space, i.e. left-hand-side of (1.1), would be treated separately. This paper consists of two parts. The rst part (Section 2 - 4) reviews the mathematical description of excitation/deexcitation and ionization/recombination, in terms of the dierential cross sections for each process. The modeling of the ionization cross sections by electron impact has been started from the early quantum models in the 1920s (see for example a recent review [7]). In this work, as a platform for development of numerical methods, our model uses the semi-classical cross section (for example, see [8]) for the energetic cross section of ionization and an extension of this cross section for excitation. The cross sections for the reverse processes are then found through detailed balance. A Boltzmann equation is also formulated for the evolution of the electron distribution and for the atom and ion densities. We then formulate and derive a version of Boltzmann's H -Theorem for excitation/deexcitation and ionization/recombination, originating from the classical denition of physical entropy. Although not unexpected, this new result, to the best of our knowledge. 4 H function formulation for atomic plasma is a The second part (Section 5 and 6) of this paper describes the development of a particle simulation method for these kinetic processes. Ionization and recombination processes have been included in the kinetic equations describing particles collisions in, for example, [911]. In those work the reaction rates depend on the macroscopic number density and temperature, since they were computed by assuming the electron distribution to be a Maxwellian. For non-equilibrium conditions, (see for example [12]), more realistic reaction rates depend on the electron velocity (or energy) rather than temperature. The recombination rates are particularly dicult to treat, since they are singular as the electron velocity goes to 0; thus, an electron with lower energy can more easily be recombined, see for example [13]. This singularity presents a new numerical challenge that we overcome in this work, using either the Reduced Rejection method of [14] or a binary search method. We employ here the framework of the Kinetic Monte Carlo (KMC) method ( [15, 16]), or equivalently the Gillespie method [17]. kinetic processes. Simulations are performed by random sampling from all of the relevant We would also like to mention that a deterministic method to solve the same system was developed as well [18]. The remainder of this paper is organized as follows. The formulation of the kinetic processes and their rates and the resulting Botlzmann equations, along with some preliminaries, are presented in Section 2. The analogue of Boltzmann's H -theorem is formulated in Section 3. The models for the cross sections are described in Section 4, and the Monte Carlo simulation method, using a particle representation for the electrons, is presented in Section 5. Computational results from this simulation method are presented in Section 6. Finally, conclusions and future directions are discussed in Section 7, while some technical details are presented in Appendices A-E. 2 Formulation 2.1 The Distribution Function The plasma is assumed to consist of electrons (e), a single type of ion, of unit charge, and neutral atoms in various electronic states. The atoms and ions are modeled as a continuum with uniform density; the electrons are modeled through an energy distribution function, which is simulated using a discrete set of particles. Throughout this paper, the temperature that the Boltzmann constant kB “ 1 Te will have units of energy so (i.e., it can be omitted). The problems solved here will be spatially homogeneous (no spatial gradients and no transport) and isotropic in velocity. Thus, we can describe the system in terms of a time-dependent electron energy distribution function (EEDF) for the free electrons, f pE, tq (E is the energy of the particle). ne and temperature Te can be The normalization is such that macroscopic electron number density obtained from by taking the rst two moments: ż8 ne “ 3 ne Te “ 2 f pEq dE, (2.1a) Ef pEq dE. (2.1b) ż08 0 Te denes the scale of the macroscopic energy of the electrons, without necessarily assuming an f and the velocity distribution function F (under the assumption that F is isotropic, i.e., F “ F pvq for v “ |v|) are Here, equilibrium, Maxwellian distribution. Note that the energy distribution function related by 4π f pEq “ me since E “ me v 2 {2 and ˆ dv “ 4πv 2 dv . 5 2E me ˙1 2 F, (2.2) We use the following notation for the other variables. kmax : the total number of levels included in atoms, nk ptq : the number density of atoms at excitation level, k ˜ na : number density of all neutral atoms “ 1, . . . , kmax , ¸ “ ÿ nk , k n` : number density of ions p” ne q , Lk : the energy of atoms at the L` : the ionization energy of an unexcited atom, gk : the degeneracy of level g` : the degeneracy of an ion, ge : the k -th excitation level, k “ 1, . . . , kmax , k, (spin) degeneracy of the electronp“ 2q. As mentioned earlier, we use the simple Bohr model for Hydrogen, such that: Lk “ p1 ´ k ´2 qIH , where IH gk “ 2k 2 , L` “ IH “ L8 , g` “ 2, (2.3) is the Rydberg unit of energy (13.6eV). By denition, the atomic ground state, from which k “ 1, i.e. L1 ” 0. Without transport of particles, n` ” ne , while mass and energy conservation imply number density of (bound and free) electrons and εtot the all energy levels are measured, corresponds to charge neutrality of the system is assured, i.e. d dt ntot “ d dt εtot “ 0, with ntot the total total energy of the system dened by ntot “ ne `na “ n` `na , ż ÿ εtot “ εe `εa `ε` “ Ef pEq dE ` nk Lk ` n` L` . (2.4a) (2.4b) k Note that momentum conservation is not used here, since it leads to terms of order of the ratio of masses, me {M` Æ 1{2000, which are neglected. At equilibrium, the electron distribution is the (isotropic) Maxwell-Boltzmann distribution 1? f eq pEq “ ne 2pπTe3 q´ 2 E e´E{Te . The equilibrium densities neq k and neq ` (2.5) for atomic level populations and ions are given by the Boltz- mann and Saha distributions respectively [19]. The former describes the ratio of number densities between two levels (e.g. l, u), and the latter is the ratio between electron and ion densities and that 1 of the atomic level from which ionization takes place Boltzmann : Saha : where h neq u Blu pTe q “ eq “ nl neq neq e Sk pTe q “ ` eq nk gu ´∆Elu {Te e , gl “ g` Ze e´Bk {Te , gk (2.6a) (2.6b) is the Planck constant, and ˆ Ze “ ge 2πme Te h2 1 ˙3{2 “ ge λ´3 e , (2.7) We emphasize that only electron-impact collisions are examined in tho work. Therefore Te , the temperature of the Maxwellian distribution of electrons (2.5) is also the governing temperature of Boltzmann and Saha equilibria, when these are reached. Therefore, the complete system at equilibrium can be described by a temperature T , applied to all plasma components. Furthermore, the equilibrium distributions are used here only to express thermodynamic relations and verify the H-theorem; no assumption is made about the actual thermodynamic state of the plasma. For example, partial equilibrium conditions can be initially created with dierent temperatures, such as in some of the test cases described in later sections. 6 is the total electron partition function, product of a term for internal states (spin states, one for the translational states. Here λe is the de Broglie wavelength of an electron (see [19] for example). We have also dened the energy gap ∆Elu “ Lu´Ll Zxa and c` Bk “ L`´Lk . ne “ n` ) to: and the binding energy These expressions are equivalent (under the assumption oc charge neutrality, where ge ” 2) and neq a neq “ gk e´Lk {Te , k Zxa (2.8a) ´L` {2Te , neq ` “c` e (2.8b) are: Zxa “ kÿ max gk e´Lk {Te , (2.9a) k“1 « 2g` na c` “ Zxa Here Zxa ˆ 2πme Te h2 ˙ 3 ff1{2 2 . (2.9b) is the atomic partition function for internal states only (ground and excited electronic states). We end this section with a lemma, which shows that specication of temperature is equivalent to specication of total energy; the proof is given in Appendix A for completeness. Lemma 2.1. Given ntot ą 0 the total number of free and bound electrons, and εtot ą 0 the total energy of the system, there are unique values of temperature Te and densities ne , na and nk , @k satisfying ne ` na “ ntot , ÿ eq 3 ne Te ` nk Lk ` neq ` L` “ εtot , 2 k eq with neq k and n` “ ne obtained using (2.6a)-(2.6b). 2.2 Interaction Types and Rates We consider here a thermal (single-uid) plasma; because of the small electron/atom mass ratio, the relative velocity and relative energy of the collision are well approximated by electron velocity and kinetic energy. We also assume also that the dierential cross sections are independent of the angular variables and only depend on the collision energy. We consider four types of collisions, ordered in pairs of forward and backward processes: u by an electron of energy E0 with E0 ą ∆Elu leaves a post-collision electron with energy E1 “ E0 ´∆Elu . We write this process as exc pE ; E q. pl, E0 q Ñ pu, E1 q, with a cross section σlu 0 1 The reverse process is deexcitation from the upper-state u down to a lower-state l after collision with an electron of energy E1 , leaving a post-collision electron with energy E0 “ E1 `∆Elu . dex We write this process as pu, E1 q Ñ pl, E0 q, with cross section σul pE1 ; E0 q. Ionization from an atomic level k , by a free electron with energy E0 ą Bk , which generates a scattered electron with energy of E1 “ E0 ´Et , a new electron with energy E2 “ Et ´Bk , and an ion. The transferred energy Et satises Bk ď Et ď E0 . We write this process as pk, E0 , Et q Ñ pE1 , E2 q, with σkion pE0 ; E1 , E2 q. The reverse process is three-body recombination, where a free electron (E1 ) and an ion recombine into an atomic state k , after collision with a second free electron (E2 ). This second electron gains energy from the collision, with post-collisional energy E0 “ E2 `E1 `Bk . We rec write this process as pE1 , E2 q Ñ pk, E0 q and σk pE1 , E2 ; E0 q. 1. The excitation of a lower state 2. 3. 4. l to an upper-state 7 Explicit formulae for these cross sections are presented in Section 4. For each process, we can dene a rate r per electron (more precisely, per electron pair for recombination): rion pk, E0 , Et q “ nk vpE0 qσkion pE0 ; E1 , E2 q, (2.10a) rrec pk, E1 , E2 q “ n` vpE1 qvpE2 qσkrec pE1 , E2 ; E0 q, exc pE ; E q, rexc pl, u, E0 q “ nl vpE0 qσlu 0 1 dex dex r pu, l, E q “ n vpE qσ pE ; E q 1 where vpEq “ a 2E{me u 1 1 ul (2.10b) (2.10c) (2.10d) 0 2 is the electron velocity. An integrated ionization rate per electron ion pk, E q “ rtot 0 ż E0 is: ion pk, E q. rion pk, E0 , Et q dEt “ nk vpE0 qσtot 0 (2.11) Bk The total, macroscopic rates are (in units of (volume Rion “ Rrec “ ÿż 8 Rdex “ ´1 ): time) ion rtot pk, E0 qf pE0 q dE0 , k Bk ÿ ż 8ż 8 (2.12a) rrec pk, E1 , E2 qf pE1 qf pE2 q dE1 dE2 , k Rexc “ ¨ 0 ÿÿ 0 ż8 (2.12b) rexc pl, u, E0 qf pE0 q dE0 , (2.12c) l uąl ∆Elu ÿÿż8 rdex pu, l, E1 qf pE1 q dE1 (2.12d) u lău 0 2.3 The Principle of Detailed Balance The principle of detailed balance determines the relations between the rates of dierent processes [19], by asserting that, at thermal equilibrium, each reaction by its inverse q Ñ p. In other words, rpÑq “ rqÑp for all pairs p Ñ q is exactly counterbalanced of states p and q . Hereafter, to simply expressions and unless specied otherwise, we will use the short notations vp “ vpEp q, and fp “ f pEp q, for p “ 0, 1, 2 (2.13) σ ion “ σkion pE0 ; E1 , E2 q, (2.14a) σ rec “ σkrec pE1 , E2 ; E0 q, σ exc “ σ exc pE ; E q, 0 lu (2.14b) (2.14c) 1 dex pE ; E q, σ dex “ σul 1 0 (2.14d) For the ionization and recombination processes, the principle of detailed balance states that: eq ion “ neq f eq f eq v v σ rec . neq ` 1 2 1 2 k f0 v0 σ eq eq and Note that nk , f (2.15) neq ` are temperature dependent, describing the statistical properties of the reactants (average over initial states), while the microscopic quantities vp , σ ion and σ rec are indepen- dent of temperature. Using the Saha equilibrium relation (2.6b), we obtain the Fowler-Nordheim relation between microscopic variables [19]: gk E0 σ ion “ 16πme g` E1 E2 σ rec . h3 2 (2.16) To lighten the formulations, we slightly abuse the notation by using the same symbols σ for cross-sections (in units of m´2 ) and dierential cross sections (m´2 /eV). The appropriate identication and choice of units can be easily obtained by counting the number of energy integration variables in the macroscopic rates. 8 Similarly, for the excitation and deexcitation processes, detailed balance implies: eq exc “ neq f eq v σ dex . neq u 1 1 l f0 v0 σ (2.17) Using the equilibrium Boltzmann relation (2.6a), we then obtain a similar relationship between microscopic variables, the Klein-Rosseland relation [19]: gl E0 σ exc “ gu E1 σ dex . (2.18) The fact that the cross-sections for forward and reverse processes are not independent is to be expected from invariance under time-reversal of the microscopic, quantum interactions. Thus, eqs. (2.16,2.18) are more fundamental than (2.15,2.17) and in turn, can be used to relate the rates of forward and backward processes, independently of thermodynamic equilibrium conditions. 2.4 The Boltzmann Equation We now formulate the Boltzmann-type equations for the electron energy distribution atom density function nk ptq, for inelastic collisions only. f pE, tq and the We present the results here and go through the details for ionization/recombination processes in Section 2.4.1 and excitation/deexcitation processes in Section 2.4.2. Since we are only examining here the collisional processes, we can restrict ourselves to a homogeneous system with no electro-magnetic elds, so that the streaming operators on the left-side of (1.1) vanish. Thus, the reduced Boltzmann equations for ionization / recombination (IR) and excitation / deexcitation (ED) are: Bt f pEq “ QeIR ` QeED , Bt nk “ QkIR ` QkED , Q` IR Bt n` “ ` (2.19) Q` ED , with the ionization/recombination collision operators QeIR “ ÿij ` ˘ pδ1 `δ2 ´δ0 q nk f0 v0 σ ion ´ n` f1 f2 v1 v2 σ rec dE0 dEt , k QkIR ij ` “´ Q` IR “ ÿij ˘ nk f0 v0 σ ion ´ n` f1 f2 v1 v2 σ rec dE0 dEt , (2.20) pnk f0 v0 σ ion ´ n` f1 f2 v1 v2 σ rec q dE0 dEt , k and the excitation/deexcitation collision operators QeED “ ÿÿż ´ ¯ pδ1 ´δ0 q nl f0 v0 σ exc ´ nu f1 v1 σ dex dE0 , u lău QkED “ ÿÿ pδuk ´δlk q ż ´ ¯ nl f0 v0 σ exc ´ nu f1 v1 σ dex dE0 , (2.21) u lău Q` ED Here E1 and E2 “ 0. in the ionization/recombination equations (2.20) are dened by E1 “ E0 ´Et , and E1 E2 “ Et ´Bk , (2.22) in the excitation/deexcitation equations (2.21) is dened by E1 “ E0 ´∆Elu . k In QED , δij “ 1 is the Kronecker delta function (“ 1 if i “ j, 0 (2.23) otherwise) and we used the short notations (2.13), (2.14) and δp “ δpE ´Ep q, 9 p “ 0, 1, 2, (2.24) 2.4.1 Ionization and Recombination Considering ionization and recombination from and to a single level k only, the kinetic equation for electrons includes a source term: ż e f0 v0 σkion pE0 ; E1 , E2 qδpE1 ´ Eq dEt dE0 ż ` nk f0 v0 σkion pE0 ; E1 , E2 qδpE2 ´ Eq dEt dE0 ż ´ nk f0 v0 σkion pE0 ; E1 , E2 qδpE0 ´ Eq dEt dE0 ż ´ n` f1 f2 v1 v2 σkrec pE1 , E2 ; E0 qδpE1 ´ Eq dE1 dE2 ż ´ n` f1 f2 v1 v2 σkrec pE1 , E2 ; E0 qδpE2 ´ Eq dE1 dE2 ż ` n` f1 f2 v1 v2 σkrec pE1 , E2 ; E0 qδpE0 ´ Eq dE1 dE2 . Bt f pEq|k “ QIR |k “ nk The rst three lines describe respectively the gain/gain/loss of electrons during an ionization process, the next three lines describe the loss/loss/gain of electrons during a recombination process, and the integration takes place over the initial states. These expressions implicitly contain the relationships between the electron energies, i.e. dE1 dE2 . (2.22) from energy conservation, and such that dE0 dEt “ Summing over all excitation levels, we obtain: e Bt f pEq “ QIR “ kÿ max ij ` ˘ pδ1 `δ2 ´δ0 q nk f0 v0 σ ion ´ n` f1 f2 v1 v2 σ rec dE0 dEt , (2.25) k“1 with δp dened in (2.24). Similarly, the equation for the atomic state is: Bt nk “ QkIR “ ij n` f1 f2 v1 v2 σkrec pE1 , E2 ; E0 q dEt dE0 ij ´ nk f0 v0 σkion pE0 ; E1 , E2 q dEt dE0 ij ` ˘ nk f0 v0 σ ion ´ n` f1 f2 v1 v2 σ rec dE0 dEt . “´ (2.26) The rst term describes the gain of atomic level density in the recombination process, while the second term describes the loss of neutral density in the k th level due to the ionization. The equation for the ions is ż ` Bt n` “ QIR “ Bt f pEq dE “ ´ ÿ Bt nk k “ ÿij pnk f0 v0 σ ion ´ n` f1 f2 v1 v2 σ rec q dEt dE0 . k 10 (2.27) 2.4.2 Excitation and Deexcitation Consider now an excitation and deexcitation process between two levels (l, u) only. The kinetic equation for the electrons is ż e Bt f pEqulu “ QED ulu “ exc pE0 ; E1 qδpE1 ´ Eq dE1 nl f pE0 qvpE0 qσlu ż exc ´ nl f pE0 qvpE0 qσlu pE0 ; E1 qδpE0 ´ Eq dE0 ż dex pE1 ; E0 qδpE0 ´ Eq dE0 ` nu f pE1 qvpE1 qσul ż dex ´ nu f pE1 qvpE1 qσul pE1 ; E0 qδpE1 ´ Eq dE1 . (2.28) The rst two lines correspond respectively to the gain/loss of electrons in the excitation processes, while the next two lines correspond to the gain/loss of electrons in the deexcitation processes. pl, uq and using short notations again, we obtain the compact form: ´ ¯ ÿÿż pδ1 ´ δ0 q nl f0 v0 σ exc ´ nu f1 v1 σ dex dE0 . “ (2.29) Summing over all pairs of levels Bt f pEq “ QeED u lău Here E0 and E1 k Bt nk “QED “ are related by (2.23). Similarly the equation for the density function of atoms is ÿż lăk ` ÿż ÿż exc nl f pE0 qvpE0 qσkl pE0 ; E1 q dE0 ´ uąk uąk “ ÿż dex nu f pE1 qvpE1 qσku pE1 ; E0 q dE1 ´ exc nk f pE0 qvpE0 qσuk pE0 ; E1 q dE0 dex nk f pE1 qvpE1 qσlk pE1 ; E0 q dE1 (2.30) lăk ¯ ÿÿż ´ nl f0 v0 σ exc ´ nu f1 v1 σ dex pδuk ´ δlk q dE0 . u lău The rst two tems correspond to the gain and loss of atoms at k -th level in excitation processes, and the next two terms corresponds to the deexcitation processes. In the last line, the summation is over all pairs of 3 pl, uq with lău by introducing the Kronecker delta function. Entropy and the H-Theorem 3.1 Entropy Np particles Sp “ kB log Wp , The Boltzmann entropy of a macroscopic system of where Wp of type p is dened as is the number of independent micro-states available and compatible with a given total energy of the system. The total phase space available to the system can be divided into unit cells, such that the number of possible arrangements of quantum states in the nth Np pnq particles distributed into Gp available cell is N Wp pnq “ We have assumed here a classical Gp p . Np ! approximation (Gp " Np ), since the studied conditions are such that no quantum statistical eects need to be considered. The Np ! factor is a correction for the fact that particles in the same unit phase space are indistinguishable. The number of particles being variable, the total number of possible arrangements for a given macro state (in units of Sp “ log ź Wp pnq “ n ÿ n 11 log N Gp p Np ! , kB “ 1 ) is (3.1) in which the dependence of N log N ´N , Gp Np and on n is implicit. Using the Stirling approximation logpN !q « the entropy (3.1) becomes Sp “ ´ ř ” ı Np N log ´ 1 . n p Gp (3.2) For a very large number of cells (continuous distributions), we can approximate the discrete summation by an integral, yielding ż Sp “ ´ „ ȷ Np pEq Np pEq log ´1 dE. Gp pEq We introduce now the denition of the occupation number variable. ζp pEq “ ζp , (3.3) a temperature-dependent, intensive Np pEq . Gp pEq For electrons, consider a small (or innitesimal) volume (3.4) d3 v d3 p in phase space. The number of possible micro-states in this phase space cell is Ge ” ge where ge “ 2 d3 x d3 p m3e 3 3 “ g d x d v, e h3 h3 is the spin degeneracy, and the remainder is the elementary volume in phase space. Assuming that the distributions are spatially homogeneous, we can replace volume 1. After a change of variable from velocity v to energy Ee d3 x by a spatial cell of (assuming isotropic distributions), one has « Ge ” ge in which get 4πm2e h3 ˆ 2Ee me ˙1{2 ff . dEe “ ge get dEe , is the energy-dependent degeneracy per unit volume due to the (3.5) translational modes only. Similarly, the number of particles in a unit spatial volume with energy in an interval of size dEe can be written as Ne pEe q “ F pvq d3 v “ f pEe q dEe The occupation number (3.4) becomes ζe pEe q “ It will be useful in the following to dene ş translational modes only ( x φpEqdE ” 1), Ne pEe q f pEe q “ . Ge pEe q ge get φpEe q (3.6) as a normalized distribution function for the and separate the occupation number into a component for internal modes only (ζ ) and one for translational modes; thus, for electrons, ζex “ ne {ge . From (3.3), we can write down the entropy for the free electrons: ˙ ˆ ż ż f pEe q ´1 dEe Se “ ´ f pEe q plog ζe ´1q dEe “ ´ f pEe q log ge get ˆ ˙ ż ζex φpEe q “ ´ne φpEe q log ´1 . get For atoms in a specied atomic state « Gk “ g k where gk k, the same denition applies. 4πMa2 h3 ˆ 2Ea Ma ˙1{2 ff dEa “ gk gat dEa , Ea , Ma are respectively the kinetic energy and mass of the Nk “ fk pEa q dEa ; note that since the atomic mass is the same for any atomic state k , use the same index (a) to describe the kinetic energy of the atom. Combining with (3.5), is now the level degeneracy, atom, and we can (3.7) 12 and performing a similar split between internal and translational modes, the occupation number for atoms at level k is Nk nk φpEa q φpEa q “ ζkx , “ Gk gk gat gat ζk “ (3.8) and the entropy for the atoms has a form similar to (3.7), i.e. Sa “ ÿ Sk ˆ ˙ ż ζ x φpEa q ´1 . Sk “ ´nk φpEa q log k t ga with k (3.9) Similarly, the occupation number for ions is ζ` “ x φpE q ζ` n` φpE` q a “ , t t g` g` g` (3.10) with the translational degeneracy 2 t “ 4πM` g` 3 ˆ h 2E` M` ˙1{2 . Because of the small electron/atom mass ratio, the kinetic energy of the atom is approximately constant during collisions, i.e. Ea M` « Ma and E` « Ea . Therefore t « gt g` a and again, we can use to describe the kinetic energy of the ions. The entropy for the ions is then: ż S` “ ´n` ˙ x φpE q ζ` a φpEa q log ´1 . gat ˆ (3.11) The total entropy of the system is the sum of that of its constituents, S “ Se ` Sa ` S` , with Se , Sa S0 , constant and S` dened by (3.7,3.9,3.11). (3.12) Note that the entropy is always dened up to a which is arbitrary, but commonly chosen such that SpT “ 0q “ 0. 3.2 Relation to H-theorem It is traditional to dene a H function 3 to describe the constrained evolution of the system towards H function includes the equilibrium distributions, * ż " f pEq eq ´ pf pEq ´ f pEqq dE H“ f pEq log eq f pEq * ÿ" nk eq ` nk log eq ´ pnk ´ nk q nk k ni ` ni log eq ´ pni ´ neq i q. ni equilibrium. The denition of a This expression is commonly referred to as the relative librium. We can also write it as the sum of partial (3.13) entropy, since it is identically zero at equi- H -function for each plasma component, which will be dened below. We can now make the correspondence with the physical entropy denitions (3.7,3.9,3.11). Using (2.5), (2.7) and (3.6), we can express the two distribution functions for the electrons as f pEe q “ ge get ζe pEe q, ge get ´Ee {T f eq pEe q “ neq e . e Ze pT q 3 (3.14a) (3.14b) In most cases, this function is equivalent to the negative entropy, and therefore is minimized at equilibrium. 13 where T is the equilibrium temperature of the complete thermodynamic system (electrons, ions and atomic states). Alternatively, we can write the equilibrium distribution for the normalized distribution function of translational modes only as gpt φ pEp q “ t e´Ep {T Zp eq Ztp “ with this expression being valid for all particles (e.g. ˆ 2πMp T h2 ˙3 Mp “ Me , Ma , M` ). 2 , Let us now compute the following expressions. First we remark that ζp “ np φp np e´Ep {T φp “ . gp gpt gp Ztp φeq p (3.15) It follows that ż ż np φp log ζp dEp “ np φp log where φp xEp y np ´ np . eq dEp ` np log t gp Zp T φp xEp y is the average thermal energy of particle p. (3.16) Since the collisions do not track the exchange of momentum, the velocity distribution functions for the atoms and ions remain the equilibrium, Maxwellian distributions; thus, for any atom in state k , φk Ñ φeq k , and the rst term on the right-hand-side of (3.16) vanishes. We also take notice of the following relation, using (2.8a) nk nk neq a ´Lk {T “ e . eq t gk Zk nk Za Therefore, nk log (3.17) Lk nk nk neq a “ n log ´nk , `n log k k eq t gk Zk Za T nk (3.18) Za “ Zxa Zta is the complete partition function. We also remark the following, neutrality (ne ” n` ) and the Saha equation (2.6b), „ eq eq eq ȷ n` ne na ne ne n` n` ne log “ ne log eq ` n` log eq ` ne log ` n` log t t t ge Ze g` Z` ne n` g` Ze neq a Za neq ne n` a “ ne log eq ` n` log eq ` n` log . Za ne n` where using charge (3.19) Combining these expressions, we arrive at the following ż ÿ φe n` nk nk log eq ` n` log eq eq dEe ` φe nk n` k „ ȷ eq na εtot ` ntot log ´ ´ ne ´na ´n` . Za T following H -functions * ż" f pEq He “ f pEq log eq ´ pf pEq ´ f eq pEqq dE, f pEq nk Hk “ nk log eq ´ pnk ´ neq k q, nk n` H` “ n` log eq ´ pn` ´ neq ` q. n` ř H “ He ` k Hk `H` is just the denition (3.13). It is p´Sq “ne φe log Let us now dene the The sum of all terms, (3.20) (3.21) then easy to see from (3.20) that p´Sq “ He ` ÿ k ȷ neq εtot a Hk ` H` ` ntot 1`log ` neq . e ´ Za T loooooooooooooooooomoooooooooooooooooon „ CpT q 14 (3.22) The last term involves a combination of conserved quantities (ntot , εtot ) and variables evaluated at equilibrium only. Since, according to lemma 2.1 there is a unique equilibrium state for the given conserved quantities, the last term in (3.22) is indeed a constant CpT q as indicated, for given initial conditions. This constant can be absorbed into the denition of the arbitrary S0 mentioned earlier, and the expression (3.22) is then formally equivalent to the negative of the system entropy. Therefore, the minimum of H is also the point of maximum entropy, i.e. the equilibrium solution. We point out that the equilibrium temperature closed system; it is equal to Te T is a unique and xed value for the complete, when the velocity distribution function for electrons has reached its equilibrium state along with the atomic and ionized states, but in general Te is a function of time; for example, one can initialize the system with a Maxwellian distribution at a temperature Te pt “ 0q, which is dierent from the nal, equilibrium value T. As the electrons start to exchange energy via the inelastic collisions with the atoms, and the distribution function will relax towards the equilibrium. The eective 4 T ptq Ñ T e temperature under which the system can actually reach as the system nears equilibrium; the conditions the exact equilibrium are discussed further below. 3.3 H-Theorems In this section we show the H-Theorems for the Boltzmann equations (2.19, 2.20, 2.21), based on the relative entropy. We rst summarize the results on moments conservations, while the proof is left in Appendix C. Proposition 3.1 (Conservation of density and energy.). (a) Suppose tf, nk , n` u is the solution to the ionization/recombination kinetic equations, i.e. (2.19, 2.20) with QeED “ QkED “ Q`ED “ 0. Then the total density and total energy are conserved, i.e., d d ntot “ εtot “ 0, dt dt where ntot and εtot are dened by (2.4a) and (2.4b). (b) Suppose tf, nk , n` u is the solution to the excitation/deexcitation kinetic equations, i.e. (2.19) (2.21) with QeIR “ QkIR “ Q`IR “ 0. Then the densities for each species and the total energy are conserved, i.e., ż d dt f pEq dE “ d d ÿ nk “ εtot “ 0, dt k dt where εtot is dened by (2.4b). (c) Suppose tf, nk , n` u is the solution to the ionization/recombination and excitation/deexcitation kinetic equations (2.19, 2.20, 2.21). Then the total density and total energy are conserved, d d ntot “ εtot “ 0. dt dt Now we give the stationary states: Proposition 3.2 (Stationary states). 8 • If the ionization and recombination processes are included, the stationary solutions tf 8 , n8 k , n` u are given by eq eq f 8 “ f eq , n8 n8 for k “ 1, . . . , kmax . (3.23) ` “ n` , k “ nk , eq eq eq Here f , nk and n` are the equilibria with temperature Te determined by the total number density ntot and the total energy εtot . 4 Dened as average thermal energy, without implying a Maxwellian distribution. 15 • If only the excitation and deexcitation processes are included, i.e. without the ionization and recombination processes, let ∆E be the greatest common divisor of ∆Elu 's, i.e. ∆E “ max t∆E ą 0 : ∆Elu is an integer multiplication of ∆E , for every l, uu . (3.24) 8 Then the stationary solutions tf 8 , n8 k , n` u are given by eq 8 eq 8 0 f “ f pE; Te qηpEq, n` “ n` , n8 for k “ 1, . . . , kmax , (3.25) k “ nk pTe q, where ηpEq is a periodic function in E ř with period ∆E , dened by j ηpEq “ ř f pE ` j∆E, t “ 0q f eq pE ` j∆E; Te q j . (3.26) with f pE, t “ 0q the initial electron distribution and Te the temperature of the stationary state determined ż ż from energy conservation Ef eq pE; Te qηpEq dE ` ÿ neq k pTe qLk “ Ef pE, t “ 0q dE ` k ÿ n0k Lk . (3.27) k where n0` is the initial ion density and neq k pT q is the k -level equilibrium atom density (2.8a) at temperature Te . The summation in (3.26) is over all integers j “ jpEq, such that E`j∆E ě 0. Remark 3.3. • The minimum amount of energy change for one electron in the excitation-deexcitation process is $ , & . ÿ ∆E “ min ∆E ą 0 : ∆E “ mlu ∆Elu , mlu PZ % l,u ∆E ηpEq. which is exactly the period periodicity of the function dened in (3.24). Here Z This is the mechanism generating the denotes the set of all positive (corresponding to excitation) and negative (corresponding to deexcitation) integers. • The non-equilibrium stationary state (3.25) depends on the periodic function ∆E . Every energy dierence ∆Elu must be an integer multiple of ∆E . ηpEq of period It follows that the non-equilibrium stationary state exists if and only if the ratio of any two energy dierences ∆Elu and ∆El1 u1 ∆Elu is a rational number. This is true for the Bohr model, since every IH . 0 very fast as the number of excitation levels increase. is a rational multiple of the ionization energy • The period 0, η ∆E decreases to As ∆E Ñ becomes a constant function and the stationary solutions are given by the Boltzmann distribution (2.5) and (2.8a), with temperature determined by energy conservation. The proposition 3.2 is a direct result of the following H-theorem on the relative entropy. We dene: Denition 3.4 . (Relative entropy) Dene the relative entropy of the system H “ He ` Ha ` H` , where He , Ha “ ř k Hk and H` (3.28) are the relative entropy of electrons, atoms and ions respectively ż f f log 8 dE ´ ne ` n8 e , f ˙ ÿˆ nk Ha “ nk log 8 ´ nk ` n8 , k n k k n` H` “ n` log 8 ´ n` ` n8 `, n` ş 8 “ f 8 dE ; tf 8 , n8 k , n` u are the the He “ where ne “ ş f dE and n8 e Proposition 3.2. 16 (3.29) stationary solutions given in ř p nk `n` q Note that is constant and ne “ n` , so that the total relative entropy (3.28) can be rewritten as ż H“ f log ˙ ÿˆ nk f n` 8 dE ` n log ` n ´ n ` n` log 8 . k k k 8 f8 n n ` k k (3.30) The H-theorem can be presented as follows: Theorem 3.5 (H-theorem). One always has He ě 0, Ha ě 0 and H` ě 0 in (3.29), hence H ě 0 in (3.28). If one further assumes the solution f to the kinetic system (2.19)(2.20)(2.21) is smooth, then for the entropy dened by (3.28)(3.29), one has dtd H ď 0. The equality holds if and only if H “ 0, which is equivalent to the fact that f pEq, nk and n` are the stationary solutions given in Proposition 3.2. This result is also valid for the solely ionization/recombination process (with ` e k QeED “ QkED “ Q` ED “ 0) or the excitation/deexcitation process (with QIR “ QIR “ QIR “ 0). The non-negativity of He , Ha and H` comes from the inequality y log xy ě py ´ xq, for x, y ą 0. We prove the decay of the relative entropy for dierent processes in the following sections. 3.3.1 Ionization and Recombination We rst prove the H-theorem for the ionization/recombination process, i.e., the equations (2.19) (2.20) with QeED “ QkED “ Q` ED “ 0. For any smooth function ż Bt ϕ “ ϕpEq, To start, we derive the weak form of the Boltzmann equations. one has for electrons, ż ϕ QeIR dE ÿij ` ˘ “ pϕ1 ` ϕ2 ´ ϕ0 q nk f0 v0 σ ion ´ n` f1 f2 v1 v2 σ rec dE0 dEt ϕf dE “ k “ ÿij ˆ pϕ1 ` ϕ2 ´ ϕ0 q k nk f0 n` f1 f2 eq eq ´ eq eq eq nk f0 n` f1 f2 ˙ (3.31) eq ion neq dE0 dEt , k f 0 v0 σ using the detailed balance relation (2.15). Similarly, for atoms and ions, ˙ n` f1 f2 nk f 0 eq eq ion ´ dE0 dEt , eq eq eq eq nk f0 v0 σ neq f n f f ` 1 2 k 0 ˙ ÿ ij ˆ nk f 0 n` f1 f2 eq eq ion “ ´ dE0 dEt . eq eq eq eq eq nk f0 v0 σ n f n f f ` 0 1 2 k k Bt nk “ QkIR “ ´ Bt n` “ Q` IR ij ˆ (3.32) (3.33) The following lemma, proven in Appendix B, will be used in the derivation of the H-theorem. eq Lemma 3.6. Suppose f pEq is a smooth function and f eq , tneq k u, n` are the equilibrium functions dened by (2.5)-(2.8b), sharing the same total number density and total energy with f , tnk u and n` . If nk f0 n` f1 f2 eq “ eq eq eq , neq f n ` f1 f2 k 0 for every E0 and Et , then f “ f eq , nk “ neq k , n` “ neq `. Now we are ready to give the H-theorem for the ionization/recombination process, 17 (3.34) Theorem 3.7 (H-theorems for ionization/recombination). Assume the solution f to the ionization/recombination kinetic equations ((2.19) (2.20) with QeED “ QkED “ Q`ED “ 0) is smooth. Then for the entropy H dened by (3.28)(3.29) with the stationary solution (3.23), one has dtd H ď 0. The equality holds if and only if H “ 0, which is equivalent to say f pEq, nk and n` are the equilibrium values. Proof. Utilizing the weak forms (3.31)-(3.33), a straightforward computation gives ż ÿ f nk n` dE ` Bt nk log eq ` Bt n` log eq ` 2Bt ntot eq f nk n` k ˆ ˆ ˙ ˆ ˙˙ ij ÿ nk f0 n` f1 f2 log “´ ´ log eq eq eq eq eq n f n ` f1 f2 k 0 k ˆ ˙ nk f0 n` f1 f2 eq eq ion ¨ dE0 dEt eq ´ eq eq eq nk f0 v0 σ neq f n f f ` 1 2 k 0 Bt H “ Bt f log ď 0. The equality holds if and only if nk f0 n` f1 f2 eq “ eq eq eq , neq f n ` f1 f2 k 0 for every E0 and Et . Lemma 3.6 then implies f “ f eq , which is equivalent to nk “ neq k , n` “ neq `, H “ 0. 3.3.2 Excitation and Deexcitation We then prove the H-theorem for the excitation/deexcitation process, i.e., the equations (2.19) (2.21) with QeIR “ QkIR “ Q` IR “ 0. Again we derive the weak form of the Boltzmann equations. ϕ “ ϕpEq, one has for electrons ij ϕf dE “ ϕ QeED dE ´ ¯ ÿż “ pϕ1 ´ ϕ0 q nl f0 v0 σ exc ´ nu f1 v1 σ dex dE0 For any smooth function ż Bt lău “ ÿż lău ˆ pϕ1 ´ ϕ0 q nu f1 nl f0 eq eq ´ eq eq nl f0 nu f1 ˙ (3.35) eq exc neq dE0 , l f0 v0 σ using the detailed balance relation (2.17). Similarly for atoms Bt nk “ QkED “ ˙ ÿ ż ˆ nl f0 nu f1 eq eq exc dE0 . ´ eq eq eq eq pδuk ´ δlk qnl f0 v0 σ n f n f u 0 1 l lău Note that for any sequence Bt ÿ tϕk u, ˙ ÿ ż ˆ nl f0 nu f 1 ÿ eq exc n k ϕk “ ϕk pδuk ´ δlk qneq dE0 eq eq ´ eq eq l f0 v0 σ n f n f u 0 1 l lău k ˙ ÿ ż ˆ nl f0 nu f 1 eq eq exc “ ´ dE0 . eq eq eq eq pϕu ´ ϕl qnl f0 v0 σ n f n f u 0 1 l lău (3.36) With only excitation/deexcitation, the equilibria f eq and neq k dened in Section 2.1 are not the only stationary solutions. The proof of this uses the following results. 18 (3.37) Lemma 3.8. (i). Suppose f is a solution to the excitation/deexcitation kinetic equations (2.19) (2.21) with initial value f pE, t “ 0q. Then d ÿ f pE ` j∆E, tq “ 0, dt j for any E ě 0, (3.38) where the summation is over all integers j such that pE ` j∆Eq ě 0. (ii). If f is in the form f pEq “ f eq pE; Te qηpE; Te q, for some temperature Te and some periodic function ηpE; Te q in E with the period ∆E , then η is given by (3.26). (iii). There is a unique value of Te satisfying the energy conservation (3.27). The detailed proof is given in Appendix D. The equation (3.38) describes the fact that an electron can only gain or lose an energy of integer multiplication of included. (ii) gives the formula of η , as a result of (3.38). ∆E with only excitation/deexcitation (iii) shows the uniqueness of the tempera- ture at stationary state, as an analogy to Lemma 2.1 where the ionization/recombination processes are included. Theorem 3.9. Assume the period ∆E dened in (3.24) exists. Suppose f pEq is a smooth function such that nu f1 nl f0 eq eq “ eq eq nl f 0 nu f1 (3.39) eq holds for every l, u and E0 , in which f eq “ f eq pE; Te q, tneq k u “ tnk pTe qu are of the form (2.5) and (2.8a)(2.8b) for the temperature Te determined from (3.27). Then f pEq “ f eq pE; Te qηpEq, nk “ neq k pTe q, (3.40) where ηpEq is the periodic function (3.26), with period ∆E . A detailed proof is given in Appendix D. Now we are ready to give the H-theorem for the excitation/deexcitation process, Theorem 3.10 (H-theorems for excitation/deexcitation). Assume the solution f to the excitation/deexcitation kinetic equations ((2.19) (2.21) with QeIR “ QkIR “ Q`IR “ 0) is smooth. Then for the entropy H dened by (3.28)(3.29) with the stationary solution (3.25), one has dtd H ď 0. The equality holds if and only if H “ 0, which is equivalent to say f and tnk u are given by (3.40). Proof. Note that with only excitation and deexcitation, one has ÿ pnk ´ n8 k q “ n` log k n` “ 0. n8 ` The entropy becomes ż H“ ˙ ÿˆ f pEq nk f pEq log eq dE ` nk log eq . f pE; T qηpEq nk pT q k 19 Utilizing the weak forms (3.35)(3.36), a straightforward computation gives ż ż ÿ nk Bt H “ Bt f log eq dE ` Bt nk log eq ` Bt f dE ` Bt nk f η nk k k ˆ ˙˙ ˆ ˙ ÿ ż ˆ ˆ nl f0 ˙ nu f1 nl f0 nu f1 eq eq exc dE0 ´ log “´ log eq eq eq eq eq eq ´ eq eq nl f0 v0 σ η η n f n f n f n f u u 0 1 0 1 0 1 l l lău ˆ ˙˙ ˆ ˙ ÿ ż ˆ ˆ nl f0 ˙ nu f1 nl f0 nu f1 eq eq exc log “´ ´ log dE0 eq eq eq eq eq eq ´ eq eq nl f0 v0 σ n f n f n f n f u 1 u 1 0 0 l l lău f ÿ ď 0. This inequality used ηpE1 q “ ηpE0 ´∆Elu q “ ηpE0 q, since ∆Elu l, u H “ 0. for every to ∆E , the period of η . nl f0 nu f1 eq eq “ eq eq , nl f0 nu f 1 is an integer multiplication of and E0 . By Theorem 3.9 and 3.8, f and tnk u The equality holds if and only if are given by (3.40), which is equivalent The stationary solution is unique for any specied initial data as a result of this theorem. 3.3.3 Full Kinetic Equations Finally consider the full kinetic equations including all four processes (2.19)(2.20)(2.21). We are ready to give: Proof of Theorem 3.5. The proof of the non-negativity of H is the same as the one in Theorem 3.7. Then it is straightforward to obtain ˙˙ n` f1 f2 ´ log Bt H “ ´ log eq eq neq ` f1 f2 k ˆ ˙ nk f 0 n` f1 f2 eq eq ion ¨ ´ dE0 dEt eq eq eq eq nk f0 v0 σ neq f n f f ` 1 2 k 0 ˆ ˙˙ ˆ ˙ ÿ ż ˆ ˆ nl f0 ˙ nu f1 nu f1 nl f0 eq eq exc ´ log ´ log dE0 eq eq eq eq eq eq ´ eq eq nl f0 v0 σ n f n f n f n f u u 0 1 0 1 l l lău ÿij ˆ ˆ nk f0 eq neq k f0 ˙ ˆ ď 0. The equality holds if and only if both (3.34) and (3.39) hold. Then by (3.34) and Lemma 3.6, f “ f eq , nk “ neq k , which is consistent with (3.39) with the periodic function 4 n` “ neq `, η “ 1. Idealized Energetic Cross Sections This section describes semi-classical cross sections, which have explicit analytic formulas. 20 4.1 Ionization For ionization events pk, E0 , Et q Ñ pE1 , E2 q, experimental measurements and numerical calculations of the dierential cross section have been performed (e.g., [20], [21]) but the results are rather complicated. As a simpler alternative, the semi-classical cross section [8] is used. This is sucient for demonstrating the Monte-Carlo procedure, while highlighting some key diculties due to the scaling of the cross-sections with energy an issue which remains for any physically realistic crosssection model. In the semi-classical approximation, we have the dierential cross section 2 σkion pE0 ; E1 , E2 q “ 4πa20 IH 1 1 1B ďE ďE . E0 Et2 k t 0 (4.1) 1 ” 1 if the conditional argument is satised, ” 0 otherwise. The total ionization cross section given initial atomic energy level k , and electron energy E0 , is ż E0 ion σtot pk, E0 q “ σkion pE0 ; E1 , E2 q dEt Bk (4.2) ˆ ˙ 1 1 2 2 1 1E0 ąBk . “ 4πa0 IH ´ E0 Bk E0 where for a “ E0 from (2.11) is ion nk vpE0 qσtot pk, E0 q “ 2 4πa20 IH The corresponding ionization rate per electron ion rtot pk, E0 q c In simulation of ionization from level the energy transfer Et k, 2 nk ? me E0 ˆ 1 1 ´ Bk E0 ˙ 1E0 ąBk . (4.3) E0 is sampled according to (4.2), then Bk ă Et ă E0 according to the probability the incident energy is chosen from the energy interval density pion pEt q “ σkion pE0 ; E1 , E2 q E0 Bk 1 1B ďE ďE . “ ion E0 ´ Bk Et2 k t 0 σtot pk, E0 q Note that this expression diverges rapidly when the excess energy (4.4) E0 ´Bk Ñ 0. 4.2 Recombination The cross section for a recombination event using pE1 , E2 q Ñ pk, E0 q is obtained from the relation (2.16), Et “ E2 ` Bk , 1 gk E0 ion σ pE0 ; E1 , E2 q ´3 16πme h g` E1 E2 k 2 4πa20 IH 1 gk 1 “ . ´3 16πme h g` E1 E2 pE2 ` Bk q2 σkrec pE1 , E2 ; E0 q “ E1 , E2 from (2.10) is rec n` vpE1 qvpE2 qσk pE1 , E2 ; E0 q 2 a20 IH gk n 1 ? ` ? . 2 ´3 2me h E1 E2 g` pE2 ` Bk q2 (4.5) The corresponding recombination rate per pair of electrons r rec pk, E1 , E2 q “ “ Both cross-sections are divergent as E1 Ñ 0 allowed range of energy transferred, i.e. near E2 Ñ 0, conditions threshold (Et » Bk ) or or 21 (4.6) obtained at either end of the near maximum (Et » E0 ). 4.3 Excitation pl, E0 q Ñ pu, E1 q, the semi-classical ˆ ˙ 1 1 1 exc 2 2 σlu pE0 ; E1 q “ p4πa0 IH qp3ful q ´ 1E0 ą∆Elu , E0 ∆Elu E0 For the energetic cross section of excitation cross section is [8] (4.7) where the absorption oscillator strength is ful “ 32 1 1 ? 5 3 ´2 . 3π 3 l u pl ´ u´2 q3 The corresponding excitation rate per electron c 2 rexc pl, u, E0 q “ 12πa20 IH ful (4.8) E0 from (2.10) is ˆ ˙ 2 nl 1 1 ? 1E0 ą∆Elu . ´ me E0 ∆Elu E0 (4.9) 4.4 Deexcitation For the cross section of the deexcitation pu, E1 q Ñ pl, E0 q, the Klein-Rosseland relation (2.18) is used: gl E0 exc σ pE0 ; E1 q gu E1 lu ˆ ˙ gl 1 1 1 2 2 “ 12πa0 IH ful ´ . gu E1 ∆Elu E1 `∆Elu dex σul pE1 ; E0 q “ E1 from (2.10) is ˆ ˙ 2 gl 1 1 1 nu ? ´ . me gu E1 ∆Elu E1 `∆Elu (4.10) The corresponding deexcitation rate per electron c 2 rdex pu, l, E1 q “ 12πa20 IH ful 5 (4.11) Monte Carlo Particle Method Using the rates detailed in the previous sections, we now formulate a Monte Carlo simulation method for the system (2.19)-(2.20), which describes the ionization, recombination, excitation and deexcitation interactions between electrons and atoms/ions in plasma dynamics. The simulation is performed using a method like that of Direct Simulation Monte Carlo (DSMC) [22] designed for binary elastic collisions. Coulomb collisions and other physical processes are not included here, but will be added in future work. There are several diculties in simulation of excitation/deexcitation and ionization/recombination: 1. There are many processes to be simulated, including pkmax pkmax ´ 1qq excitation/deexcitation processes and (2kmax ) ionization/recombination processes. One needs to pick the processes and sample corresponding electrons with the right rates. 2. The recombination rate depends in a singular way on the electron energies, so that ecient and 3. The rates for dierent interactions are widely varying, so that it is dicult to eciently include correct sampling of the electrons for recombination events is dicult. all of the interactions. In this section we mainly focus on dealing with the rst two diculties. The third involves the multiscale properties which will be investigated in future work. 5.1 The Framework of Algorithm We employ a Kinetic Monte Carlo (KMC) method to deal with the rst diculty. The method can be summarized as 22 Step 1. Find upper bounds for the total rates of each of the four collision types, denoted as with the superscript R̃ “ Step 2. ř j “ exc, dex, ion, rec. r̃j The total rate of all types of collisions is given by j j r̃ . j Choose collision type with rate r̃j {R̃. Update the time with t “ t ` ∆t, where ∆t “ 1{R̃, so that there is only one virtual collision in one time step. Step 3. For the selected collision type, choose the participating particles, including electrons and atom levels before and/or after the collision, with the rate used in computing the upper bound r̃j . Step 4. Denote this rate to be r̃Pj , with the subscript P representing the particles. For the selected collision type and particles, compute the actual rate j j Accept and perform the collision with probability rP {r̃P . rPj of this collision. 5.2 Sample Electrons from a Singular Distribution In step 3 of the framework, electrons for excitation, deexcitation and ionization processes are uniformly chosen from the numerical particles. For the recombination process, the actual rate uniformly chosen rPj scales ´1{2 , if two electrons with energy E and E are like pE1 E2 q from the numerical 1 2 rec extremely large and the acceptance/rejection procedure particles. This makes the upper bound r̃ in step 4 very inecient. This is the second diculty mentioned above. To avoid this singularity, we choose electrons ´1{2 as scales like E E Ñ 0. E with rate ppEq, in which the distribution As described in the following sections, the two electrons E1 and ppEq E2 in the recombination are chosen with probabilities 1 “? , E 1 ÿ gl p2 pEq “ ? . E l pE ` Bl q2 p1 pEq (5.1) Now consider the following model problem. Suppose we want to simulate some reaction process over a xed time interval r0, tmax s, with the reaction rate R. Here R such as (2.12). For each collision one needs to choose an electron Ej with tEj , j “ 1, . . . , Ne u. Then the total number of collisions Rtmax {Ne “ Rtmax Ne {ne “ OpNe q, where ne is the physical density of electron, Ne is number of numerical particles and Ne is the eective number. from Ne pvolume ¨ timeq´1 , probability pj “ ppEj q is in the unit of numerical electrons with energy needed is the We propose four dierent methods here for this model problem and give a numerical comparison in Section 6.4. 5.2.1 Acceptance/Rejection Method For this method, rst compute pmax “ max pj , (5.2) 1ďjďNe then uniformly choose an electron is accepted. The value pmax Ej and accept it with rate pj pmax . Keep sampling until one electron can be updated after each collision. This method is used for the excitation, deexcitation and ionization. But it is very inecient for the recombination due to the singularity as is at equilibrium. At equilibrium for small E ă E˚ E Ñ 0. To see this, consider the case E˚ , using equation (2.5), the fraction when the system of electrons with is P pE ă E˚ q “ C ż E˚ ? 3{2 Ee´E{Te dE “ OpE˚ q, 0 23 (5.3) so that ˆ P ˙ min Ej ď E˚ 1ďjďNe Hence the minimum of E is ´2{3 OpNe q 3{2 “ 1 ´ p1 ´ P pE ă E˚ qqNe “ OpNe E˚ q. and rAR “ To simulate over r0, tmax s, pmax “ ?1 E 1{3 “ OpNe q. The acceptance rate is 1 “ OpNe´1{3 q. pmax the total cost is CAR “ O ˆ Ne rAR ˙ “ OpNe4{3 q. 5.2.2 Direct Search Method A direct search method is to sample a number r # j “ min ξ, s.t. ξ ÿ pi ě r i“1 Here the summation s“ r0, tmax s Ne ÿ and look for pi . (5.4) i“1 řNe i“1 pi can be updated after each collision. In this method the singularity is not a problem. However the searching for over r0, 1s + uniformly from j takes OpNe q operations. The total cost to simulate is Cdirect “ OpNe2 q. 5.2.3 Binary Search Method j in (5.4). Divide r0, Ne s into two equal size subdivisions and rst locate which division Ej is in. If this idea is applied iteratively to each subdivision, the total cost can be reduced to Oplog2 Ne q. This is the widely used ř binary search method. For this method the sum s “ pi in each subdivision can be updated after each collision. A total number of Oplog2 Ne q updates is need for each collision. For completeness a detailed algorithm is given in Appendix E. The total cost to simulate over r0, tmax s is Cbinary “ OpNe log2 Ne q. The direct search method can be accelerated by a binary search method to nd 5.2.4 Reduced Rejection Method The recently developed reduced rejection method [14] can be applied, by combining with the Marsaglia table method [23]. See Section 6 in [14] for a detailed description. There are two sources of computation costs. collisions. Then there are method is applied. OpKq Suppose the Marsaglia table is re-created after K particles in the L-region, where a regular acceptance/rejection According to Section 5.2.1, the cost of sampling one electron is the L-region. Hence the total cost of sampling OpNe q OpK 1{3 q in 1{3 q. A total number of particles is OpNe K OpNe {Kq Marsaglia tables are since the cost of creating one table is OpNe q, the resulting ` Nre-created, ˘ e total cost in this part is O N . Therefore the total cost of the reduced rejection method to e K simulate over r0, tmax s is ˆ ˙ Ne2 1{3 CRR “ O Ne K ` “ OpNe5{4 q, K if 3{4 K “ Ne . Note that to minimize the total cost, we need the costs in the two sources to be approximately equal, i.e., Ne K 1{3 « Ne2 K . In practice, we use a simple technique to make the cost of re-creating Marsaglia tables and the cost of sampling from L-regions approximately equal: 24 • • Step 1. Create the Marsaglia table. Record the computation time as Step 2. Perform collisions. Keep track of the total time tsample ttable . used in sampling particles for recombination. When tsample ą ttable , set tsample “ 0 and go to step 1. 5.3 Large Scale Variation in the Rates This section illustrates the large variation in the rates of dierent processes. Ecient multiscale algorithms will be investigated in future works. The total rate of excitation from level R exc ż pl, uq “ l to level u at equilibrium is rexc pl, u, E0 qf eq pE0 q dE0 “ 3ful nl Cpne , Te qψ Cpne , Te q is a ˙ 1 1 ´y ´ e dy. x y where ful is the absolute oscillatory strength (4.8), ż8ˆ ψpxq “ x ˆ ∆Elu Te function of ne ˙ , and Te , and One can easily check ˆ ˙ 1 , ψpxq “ O x as x Ñ 0. For large l, the value of ful is maximized at ∆Elu is minimized at the same Therefore, since u “ l ` 1, for which 4 ? l. fl`1,l « 3π 3 post-collisional level u “ l ` 1, 2 ∆El,l`1 « 3 IH . l nl “ Opl2 q, Rexc pl, l ` 1q “ nl fl`1,l ¨ O ˆ 1 ∆El,l`1 ˙ Similarly, at equilibrium the total rate of ionization from level Rion plq “ nl ¨ O ˆ 1 Bl ˙ l “ Opl6 q. (5.5) is “ nl ¨ Opl2 q “ Opl4 q. (5.6) (5.5) and (5.6) indicate that there are two dierent scales in the excitation/ionization problem. First, the scales of excitation and ionization vary dramatically as the atom level 4 of ionization increases as Opl q and the rate of excitation increases as for typical values of l increases. The rate 104 „ 106 Opl6 q, a range of kmax “ 10. Second, the rate of excitation is much larger than that of ionization for large excitation level, with ratio Rexc pl, l ` 1q “ Opl2 q. Rion plq Some numerical examples are presented in Section 6.5. 25 5.4 Particle/Continuum Representation The plasma is described by a discrete set of particles for the electron diestribution continuum description for the atoms nk ptq and ions n` ptq. f pE, tq and a f pE, tq The electron distribution function is represented by the following discrete distribution f pE, tq “ Ne Ne ÿ δpE ´ Ej ptqq. (5.7) j“1 The number units of f Ne “ Ne ptq of electrons is not constant because of ionization and recombination. The are (number of particles) per energy per volume, and the constant Ne is the eective number which represents the number of physical particles per numerical particle per volume. 5.5 A No Time Counter DSMC method This section describes in detail a Direct Simulation Monte Carlo method without time counter. The ntot and εtot method follows the framework illustrated in Section 5.1. It preserves the total electron density (including the free and bound electrons) and the total energy εtot , due to the fact that ntot are preserved in each collision. Moreover, since the algorithm satises detailed balance, if initially the system is in an equilibrium state, then it is always in equilibrium. This is numerically tested in Section 6.1. Step 1 : Decide the rates of virtual collisions. First nd upper bounds on the total rates of collisions of each type. The upper bounds need to be not too large, while at the same time easy to compute. The actual total rates of collisions in each type can be derived by taking the summation of the rates (4.3), (4.6), (4.9), (4.11) over all electrons and atom levels: rexc “ Ne ÿÿÿ rexc pl, u, Ek q l uąl k c ÿ ÿ ÿ nl 2 ? ful Ne “ me Ek l uąl k ÿÿÿ rdex “ Ne rdex pu, l, Ek q 2 12πa20 IH ˆ 1 1 ´ ∆Elu Ek ˙ 1Ek ą∆Elu , u lău k c “ 2 12πa20 IH rion “ Ne ÿÿ l ion pl, E q rtot k k c 2 “ 4πa20 IH 2 rrec “ Ne ? ÿ ÿ ÿ nu g l 2 ful Ek Ne , me gu ∆Elu p∆Elu ` Ek q u lău k ÿÿ ÿ ÿ nl 2 ? Ne me Ek l k ˆ 1 1 ´ Bl Ek ˙ 1Ek ąBl , rrec pl, Ei , Ej q i‰j l ¸ ÿ ÿ 1 ÿ gl gl 1 a “ ´ Ei l pEi ` Bl q2 Ei Ej l pEi ` Bl q2 i,j i ˜ ˜ ¸ ¸ ÿ 1 ÿ ÿ 1 ÿ ÿ g a20 n` 1 g l l a ? “ I2 N 2 ´ . 4m2e h´3 g` H e j Ei l pEi ` Bl q2 Ei l pEi ` Bl q2 Ej i i a20 n` I2 N 2 4m2e h´3 g` H e ˜ ÿ (5.8) 26 Note that these are numerical approximations of (2.12). They have the following upper bounds: 2 ˜ “ 12πa20 IH rexc ď rexc ˜ “ 12πa2 I 2 rdex ď rdex 0 H c ÿ ÿ 2 2 ful ? Ne Ne nl , me 3 3 p∆Elu q3{2 l uąl c ÿ ÿ gl 2 ful Ne Ne nu , me g 2p∆Elu q3{2 u lău u c ÿ 2 2 nl ? Ne Ne , 3{2 me pB q 3 3 l l ˜ ¸ 2 ÿ ÿ 1 ÿ 1 g a n 2 ` l 0 ˜ “ a ? rrec ď rrec I2 N 2 , 4m2e h´3 g` me H e j Ei l pEi ` Bl q2 Ej i ˜ “ rion ď rion using the inequalities 2 4πa20 IH 2 ? a´3{2 , for 3 3 x´1{2 p1{a ´ 1{xq ď x ě a, and ? 2 x apa`xq ď a´3{2 . Then ˜ ` rion ˜ ` rrec ˜ ` rdex ˜ R̃ “ rexc is an upper bound on the total rate of all collisions. Evaluation of R̃ requires total computational cost of Oppkmax q2 ` Ne q at beginning, with kmax the number of excitation levels. Since only one or two electrons and one or two atom levels are updated after each collision, it is easy to update R̃ in each step. Step 2 : Update time and choose collision type. For a time step ∆t, there are R̃∆t{Ne virtual collisions. For the space homogeneous problem, one can take ∆t “ Ne . R̃ This eliminates the time error. An equivalent choice is to take Ne lnp1{uq, R̃ r0, 1s. This has ∆t “ with u a number uniformly sampled from been widely used in the simulation of chemical reactions since Gillespie's work [17] Then update the time t “ t ` ∆t. r̃j , R̃ The collision type is chosen with probability for j P texc, dex, ion, recu. Step 3 and 4 : Perform one virtual collision We combined the steps 3 and 4 in Section 5.1. For the dierent collision types, the algorithms are slightly dierent. • For excitation: Step 3.1 Choose a free electron E0 uniformly. Step 3.2 Choose a prior-collisional level l and post-collisional level u ą l with rate ř l by the following method: Denote nl ful 3{2 p∆Elu q ř ful nl , uąl p∆E q3{2 lu alu “ ful p∆Elu q3{2 and bl “ ř uąl alu . talu u and tbl u are pre-computed. Then apply the direct search method described in Section 5.2.2) to choose rst Opkmax q. l nb with rate ř l l l nl bl , then 27 u with rate alu bl . The total cost in this step is Step 3.3 Sample a number q r0, 1s. If ¯ ´ E10 1E0 ą∆Elu uniformly from qď ?1 E0 ´ 1 ∆Elu 2 1 ? 3 3 p∆Elu q3{2 , accept the collision and perform the excitation pl, E0 q Ñ pu, E1 “ E0 ´ ∆Elu q. Otherwise reject it. • For deexcitation: Step 3.1 Choose a free electron E1 uniformly. Step 3.2 Choose a prior-collisional level u and post-collisional level l ă u with rate nu ggul 2p∆Eful q3{2 lu , ř ř gl ful n u u lău gu 2p∆E q3{2 lu by a method similar to Step 3.2 in the case of excitation. Step 3.3 Sample a number q r0, 1s. uniformly from ? E1 ∆Elu `E1 1 2p∆Elu q1{2 qď If , accept the collision and perform the deexcitation pu, E1 q Ñ pl, E0 “ E1 ` ∆Elu q. Otherwise reject it. • For ionization: Step 3.1 Choose a free electron E0 uniformly. Step 3.2 Choose a prior-collisional level l with rate nl pBl q3{2 nl l pBl q3{2 ř Step 3.3 Sample a number q . r0, 1s. If ¯ ´ E10 1E0 ąBl uniformly from qď ?1 E0 ´ 1 Bl 2 1 ? 3 3 pBl q3{2 Et E0 Bk 1 1B ďE ďE . E0 ´ Bk Et2 k t 0 accept the collision and sample the transfer energy , according to and perform the ionization pl, E0 q Ñ pE1 “ E0 ´ Et , E2 “ Et ´ Bl q. Otherwise reject it. • For recombination: Step 3.1 Pick up one free electron with with ?1 Ej ř 1 , ? j Ej and denote it by E1 . Pick up a second electron with rate ?1 Ei ř ´ i 28 gl l pEi `Bl q2 ř ?1 Ei gl l pEi `Bl q2 ř ¯, and denote it by E2 . As described in Section 5.2, this step can be accomplished by several dierent methods. In the numerical section part, we apply four dierent methods and compare their eciencies. Step 3.2 If the two electrons are the same, reject the collision. Otherwise, choose the l post-collisional level with rate gl pE2 `Bl q2 ř gl l pE2 `Bl q2 . Step 3.3 Perform the recombination, pE1 , E2 q Ñ pl, E0 “ E1 ` E2 ` Bl q. 6 Computational Results This section presents computational results from the simulation method described above. 5 atomic levels are included, unless specied. All the results except Section 6.4 are obtained using the binary search method in recombination. However all the other methods described in Section 5.2 give the same results. 6.1 Equilibrium initial data The rst results are for a distribution of electrons, neutral atoms and ions that are in the equilibrium Te “ 4.31 eV and number 100, 000 computational particles are used distribution given by (2.5), (2.8a) and (2.8b), with initial temperature of density ntot “ 1028 m´3 for all (free and bound) electrons. to represent the electron distribution function. The model for the neutral atoms consists of ve levels. The left of Figure 1 shows the time evolution of the atom density right shows the electron energy distribution function f pEq nk , for dierent level k; the at various times. The plot on the top is for a simulation that includes excitation/deexcitation alone with no ionization/recombination; while that on the bottom is for ionization/recombination with no excitation/deexcitation. In this simulation the time step is ∆t “ 6.5´11 for the ionization/recombination. ps for the excitation/deexcitation, and ∆t “ 4.5´10 8 The simulation is performed for about 10 time steps. ps These results show that the simulation method does preserve the equilibrium distribution as required. The uctuations in the results are due to the statistical error from a nite number of computational particles. 6.2 Non-equilibrium initial data The second simulation results, shown in Figure 2 - 7, are for initial data for which all of the neutral atoms are distributed at the ground level, with total number density na “ 5 ˆ 1027 m´3 . The electrons and ions are initially in an equilibrium distribution given by (2.5) and (2.8b), with initial temperature of Te “ 50000K “ 4.31 eV and number density ne “ 5 ˆ 1027 m´3 . 1, 000, 000 computational particles are used to represent the electron distribution function. The model for the neutral atoms consists of ve levels. In Figures 2, 3, 4, only excitation/deexcitation processes are included. The plots in the left column of Figure 2 show the density of the ve dierent atomic levels (blue bars) at various times, demonstrating convergence to equilibrium (white bars) given by (2.8a). The plots in the right column show the density distribution of free electrons (blue lines) at various times, demonstrating convergence to equilibrium (red dotted lines) given by (2.5). Note that the atomic level population inversion 29 27 44 x 10 k=5 16 x 10 t = 0.002825 ps t = 0.005682 ps t = 0.008539 ps t = 0.011395 ps t = 0.014254 ps t = 0.017111 ps t = 0.019968 ps 14 k=1 EEDF / m−3 J−1 12 k=4 k=3 10 8 6 4 k=2 0.005 2 0.01 time / ps 0.015 0 0.02 27 0 0.2 0.4 0.6 0.8 1 energy / IH 1.2 1.4 1.6 1.8 2 44 x 10 16 k=5 x 10 t = 0.002423 ps t = 0.005053 ps t = 0.007686 ps t = 0.010319 ps t = 0.012945 ps t = 0.015571 ps t = 0.018201 ps 14 k=1 EEDF / m−3 J−1 12 k=4 k=3 10 8 6 4 k=2 2 0.005 0.01 time / ps 0.015 0.02 0 0 0.2 0.4 0.6 0.8 1 energy / IH 1.2 1.4 1.6 1.8 2 Figure 1: Atomic (left) and electron (right) density function at various times starting from equilibrium, with excitation/deexcitation only (top), and ionization/recombination only (bottom) 100000 samples for free electrons. 30 is expected, because the level degeneracy grows faster than the exponential drop o due to ∆E (except for between level 1 & 2) since the higher levels are close together. Figure 3 shows the time evolution of the electron temperature (top) ş8 εe 2 0 Ef pEq dE ş “ Te “ p3{2qne 3 08 f pEq dE and the normalized relative entropy (bottom) H“ where H H ntot Teeq “ 27959K determined from the Te decays to Teeq as time evolves, due is dened by (3.13), with the equilibrium temperature density and energy conservation. The electron temperature to the fact that the excitations from ground level to excited levels dominate in the evolution. The normalized relative entropy function (negative entropy) is monotonically decreasing to 0, except for uctuations due to the nite number of sampled particles which scales like Figure 4 shows the time evolution of various ratios. In the top we show type of virtual collision; in the middle collisions; in the bottom ? 1{ N . r̃j , the fraction of each R̃ rj , the ratio of each type of accepted collision over all virtual R̃ rj , the acceptance rate of each type. Here r̃j j P texc, dexu. In the middle gure, the two ratios become equal as time evolves, meaning that the two processes get balanced. In the bottom gure, the acceptance rates of each virtual collision are above 50%, showing the high eciency of our method. consecutive collisions. Each point in these gures is computed by taking average over 50000 Note that compared to the entropy decay in Figure 3, these ratios reach steady states in a much shorter time. This is because the collisions involving high levels (in this example, level 4 and 5) have much larger rates, and hence dominate in computing these ratios. The contribution from the low levels is very small. As the high levels can reach equilibrium in a very short time, these ratios reach steady states very fast. On the other hand, the contribution from the low levels in computing the entropy is not negligible, leading to the slow decay of the entropy in Figure 3. Note that with only excitation/deexcitation, the stationary state of the electron distribution diers from the equilibrium by a periodic function the period of η η, according to (3.40). However with 5 excitation levels is ∆E “ 1 IH 32 42 52 “ IH , 3600 which is too small to be observed in this experiment. The relaxation towards this stationary state does not preclude the generation of physical entropy, as shown by Figure 3, since according to the discussion of Section 3.2, the evolution of the system can be described by either form of entropy, but it does imply that by allowing only transitions with nite energy gaps, the nal value of does not correspond to the absolute minimum (maximum of physical entropy). H This eect was also observed in a non-stochastic solver [18], and is not an artifact of our numerical approach, but rather a physical feature of the system, indicating breakdown of ergodicity for system with only excitation/deexcitation. In Figure 5, 6, 7, similarly results are obtained with only ionization and recombination. In contrast to the previous case of excitation/deexcitation, there is a continuum of energy that can be exchanged with the free electrons. Therefore in this case the stationary state of the electron distribution is exactly the equilibrium state. Figure 6 shows the time evolution of the degree of ionization (top), the electron temperature (middle) and the normalized relative entropy (bottom), with 31 H dened by (3.13). More recombination atom distribution / m−3 27 t = 0.00 ps 8 x 10 45 8 actual equilibrium 6 4 2 2 1 2 3 Levels 4 0 5 27 t = 0.02 ps 8 8 6 6 4 4 2 2 1 2 3 Levels 4 0 5 27 t = 0.50 ps 8 Figure 2: 0 0.5 1 Energy / IH 1.5 2 0.5 1 Energy / IH 1.5 2 0.5 1 Energy / IH 1.5 2 x 10 0 45 x 10 8 6 6 4 4 2 2 0 actual equilibrium 45 x 10 0 electron distribution / m−3J −1 6 4 0 x 10 1 2 3 Levels 4 0 5 x 10 0 Snapshots of time evolution of atom density (left) and electron distribution (right), compared to the equilibrium (white bars and red dotted lines), starting from all neutral atoms at the ground level. Only excitation/deexcitation processes are included. collisions happen than the ionization, leading to the decay of the degree of ionization. In the early time, the electron temperature increases due to the fact that electron with lower energy is easied combined with an ion. This can also be observed in the second snapshot of EEDF in Figure 5. As the number of low energy electrons decreases, the high energy electrons start to dominate. More high energy electrons are combined with ions, leading to the decrease of the electron temperature. The relative entropy is monotonically decreasing to 0, except for the statistical uctuations. Note that compared to the excitation/deexcitation collisions (see Figure 3), the ionization/recombination collisions take longer time to reach equilibrium. This is due to the fact that the collision rates for excitation/deexcitation are larger than those of ionization/recombination (see Figure 13 for an example). Again Figure 7 shows the acceptance rate of each virtual ionization collision is above acceptance rate of each virtual recombination collision is almost 50%. The 100%, since we are using the binary search method for the recombination process and the only rejected collisions in our algorithm (Step 3.2) are those with two identical incident electrons. 32 4 electron temperature / K 5 x 10 4.5 4 3.5 3 temperature at equilibrium 2.5 0 0.01 0.02 0.03 0.04 0.05 time / ps relative entropy 0.06 0.07 0.08 0.09 0.1 0 0.01 0.02 0.03 0.04 0.05 time / ps 0.06 0.07 0.08 0.09 0.1 0 10 −1 10 −2 10 −3 10 −4 10 Figure 3: Time evolution of the electron temperature (top) and the normalized relative entropy H (bottom), starting from all neutral atoms at the ground level. processes are included. 33 Only excitation/deexcitation ratio of virtual collisions 1 excitation deexcitation 0.5 0 0 0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035 0.004 0.0045 0.005 time / ps ratio of actual collisions 0 0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035 0.004 0.0045 0.005 time / ps acceptance rate 0 0.0005 0.001 0.0015 0.002 0.0025 0.003 0.0035 0.004 0.0045 0.005 time / ps 1 0.5 0 1 0.5 0 Figure 4: Time evolution of the fraction of each type of virtual collision (top), the ratio of each type of accepted collision over all virtual collisions (middle) and acceptance rate (bottom), starting from all neutral atoms at the ground level. Only excitation/deexcitation processes are included. 34 −3 27 t = 0.00 ps 8 45 atom distribution / m x 10 4 actual equilibrium 6 2 2 1 1 2 3 Levels 4 0 5 t = 0.02 ps 4 6 3 4 2 2 1 1 2 3 Levels 4 0 5 t = 0.50 ps Figure 5: 1 Energy / I 1.5 2 1 Energy / I 1.5 2 1 Energy / IH 1.5 2 x 10 0 0.5 45 x 10 4 6 3 4 2 2 1 0 0.5 H 27 8 0 45 x 10 0 actual equilibrium H 27 8 −3 −1 electron distribution / m J 3 4 0 x 10 1 2 3 Levels 4 0 5 x 10 0 0.5 Snapshots of time evolution of atom density (left) and electron distribution (right), compared to the equilibrium (white bars and red dotted lines), starting from all neutral atoms at the ground level. Only ionization/recombination processes are included. 35 0.5 0.45 degree of ionization 0.4 0.35 0.3 0.25 0.2 0 0.01 0.02 0.03 0.04 0.05 time / ps 0.06 0.07 0.08 0.09 0.1 0.01 0.02 0.03 0.04 0.05 time / ps 0.06 0.07 0.08 0.09 0.1 0.3 0.35 0.4 0.45 0.5 4 x 10 9.5 9 electron temperature / K 8.5 8 7.5 7 6.5 6 5.5 5 4.5 0 relative entropy 0 10 −1 10 −2 10 −3 10 −4 10 −5 10 0 0.05 0.1 0.15 0.2 0.25 time / ps Figure 6: Time evolution of the degree of ionization (top), the electron temperature (middle) and the normalized relative entropy H (bottom), starting from all neutral atoms at the ground level. Only ionization/recombination processes are included. 36 ratio of virtual collisions ionization recombination 1 0.5 0 0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012 0.0014 0.0016 0.0018 0.002 time / ps ratio of actual collisions 0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012 0.0014 0.0016 0.0018 0.002 time / ps acceptance rate 0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012 0.0014 0.0016 0.0018 0.002 time / ps 1 0.5 0 1 0.5 0 Figure 7: Time evolution of the fraction of each type of virtual collision (top), the ratio of each type of accepted collision over all virtual collisions (middle) and acceptance rate (bottom), starting from all neutral atoms at the ground level. Only ionization/recombination processes are included. 37 6.3 The stationary states without ionization/recombination In this section we compute the stationary states with only excitation/deexcitation processes. The model includes only two atomic levels. The period of the stationary function Initially the free electrons are ηpEq in (3.40) is 3 ∆E “ ∆E12 “ L2 ´ L1 “ IH . 4 placed near E “ ∆E with distribution 2 f pE, 0q “ ? with σ0 “ IH 10 and ne “ 0.9ntot , where the total density distribution is The time evolution is shown in Figure 8. period of ´ pE´∆Eq ne 2 2σ0 e , 2πσ0 (6.1) ntot “ 1028 m´3 . The stationary electron distribution clearly shows the 3 4 IH . Figure 9 shows the time evolution of the normalized entropy H“ H ntot , where H is the entropy dened by (3.28)(3.29) with the stationary solutions given by (3.25). This veries the convergence to the theoretical result. Furthermore, one can see from Figure 9 that the relaxation is much slower than the one in Figure 3. This is due to the fact that the collision between the electron and the higher atomic level has a larger rate. In Figure 3, ve atomic levels are included; while in Figure 9 only two atomic levels included. We have also solved a similar problem with three atomic levels and observed the period of 1 36 IH in the stationary solution. In the next we start with the same initial data, but with all the processes (excitation/deexcitation and ionization/recombination) included. The time evolutions of the atomic distribution, with only two levels, and the electron distribution are shown in Figure 10. With the ionization/recombination f odicity in electron distribution f processes included, the EEDF decays to the Maxwell-Boltzmann distribution. However the perican still be observed at the early time of evolution, due to the fact that the collision rates of excitation/deexcitation is higher than those of ionization/recombination (see Figure 13 for an example). 6.4 Comparison of dierent methods for sampling electrons in recombination We apply the four methods described in Section 5.2 to the excitation/ionization problem and make a comparison. The rst test problem is the simulation of a system in equilibrium. Initially all atoms are ionized and free electrons are given by a Maxwellian. ntot “ 1028 m´3 and the temperature of free electron Te “ 4.31 ´3 eV. The degree of ionization stays around 38.04%. The simulation is performed until time t “ 10 6 ps, which requires around 2 ˆ 10 time steps for Ne “ 100000. We take the total electron density We solve this problem with numbers of simulation particles for electrons to be Ne “ 781, 1562, 3150, 6250, 12500, 25000, 50000, 100000, 200000, 400000, 800000 respectively for the initial free electrons. We track the time spent on the processes related to the sampling of electrons for recombination collisions. For reduced rejection method this includes the rebuilding of the Marsaglia tables and the sampling process. For the binary search method, this includes the updating of the summation tables and the sampling process. For direct search method, only the sampling process are timed. 38 atom distribution / m−3 26 t = 0.00 ps 8 x 10 2 6 1.5 4 1 2 0.5 0 1 electron distribution / m−3J −1 46 actual stationary 0 2 x 10 actual stationary 0 0.5 1 1.5 2 Energy / IH 2.5 3 3.5 1 1.5 2 Energy / IH 2.5 3 3.5 1 1.5 2 Energy / IH 2.5 3 3.5 Levels 26 t = 0.02 ps 8 46 x 10 2 6 1.5 4 1 2 0.5 0 1 0 2 x 10 0 0.5 Levels 26 t = 0.52 ps 8 46 x 10 2 6 1.5 4 1 2 0.5 0 1 0 2 x 10 0 0.5 Levels Figure 8: Snapshots of time evolution of atom density (left) and electron distribution (right), compared to the stationary states (white bars and red dashed lines), starting from the initial data (6.1). Only excitation/deexcitation processes are included. 200000 samples are used for free electrons. relative entropy 0 10 −1 10 −2 10 −3 10 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 time / ps Figure 9: Time evolution of the normalized relative entropy H dened by (3.28)(3.29), with the initial data (6.1). Only excitation/deexcitation processes are included. free electrons. 39 200000 samples are used for 27 t = 0.000 ps 2 x 10 atom distribution / m−3 46 2 actual stationary 1.5 1 0.5 0.5 1 electron distribution / m−3J −1 actual stationary 1.5 1 0 x 10 0 2 0 1 2 Energy / IH 3 1 2 Energy / IH 3 1 2 Energy / IH 3 Levels 27 t = 0.002 ps 2 46 x 10 2 1.5 1.5 1 1 0.5 0.5 0 1 0 2 x 10 0 Levels 27 t = 0.010 ps 2 46 x 10 2 1.5 1.5 1 1 0.5 0.5 0 1 0 2 Levels Figure 10: x 10 0 Snapshots of time evolution of atom density (left) and electron distribution (right), compared to the stationary states (white bars and red dashed lines), starting from the initial data (6.1). All processes are included. 1000000 samples are used for free electrons. 40 time spent on sampling for recombination 5 time spent on sampling for recombination 4 10 10 regular rejection direct search reduced rejection binary search 4 10 3 10 regular rejection direct search reduced rejection binary search 3 10 slope = 1.35 slope = 1.85 2 10 slope = 1.90 slope = 1.18 10 cputime / s cputime / s slope = 1.27 slope = 1.24 2 1 10 1 10 slope = 1.11 0 10 slope = 0.96 0 −1 10 10 −1 −2 10 10 −2 −3 10 3 4 10 5 10 10 number of initial samples 10 6 10 3 4 10 5 6 10 10 number of initial samples 10 Figure 11: The time cost for recombination collisions when various numbers of samples are used for the initial free electron distribution. Left: the system stays in equilibrium for all the time. Right: the equilibration process with all atoms ionized at beginning. The results are shown in the left of Figure 11. As expected the direct search method needs computations. It is getting extremely slow when more than 106 samples are used. OpNe2 q The accep- 1.35 q computation, which is very close to the theoretical results. tance/rejection method needs OpNe However, due to the large upper bound caused by the singularity of collision rate as E Ñ 0, the acceptance rate in the acceptance/rejection method is very small, leading to ineciency of the acceptance/rejection method. The reduced rejection method is faster than the acceptance/rejection OpNe1.25 q computation. The cost for binary search method is OpNe q, which agrees with the theoretical result OpNe log Ne q. Note that the binary method is about 10 times faster than method and needs the reduced rejection method, although these two methods have almost the same convergence order. A similar test is applied on the simulation of a equilibration process. electron density ntot “ Again we take the total 1028 m´3 and the temperature of free electron Te “ 4.31 eV. In this case the 44.50%. We simulate the decaying process until t “ 10´3 ps, steps for Ne “ 100000. At the end time, the degree of ionization is degree of ionization at equilibrium is which requires around around 106 time 48.18%, indicating that the system is close to the equilibrium. The simulation is performed t “ 10´3 ps, which requires around 2 ˆ 106 time steps for Ne “ 100000. The results are until time shown in the right of Figure 11. We observe similar eciencies as in the equilibrium test. 6.5 The large scale variation This section numerically checks the large scale variation in the rates illustrated in Section 5.3. Figure 12 shows the rates for excitation and deexcitation at equilibrium. Here 10 atomic levels are Ñ lq gives the rate of excitation or deexcitation (for k ă l or k ą l respectively) l. The rates with k “ l are not dened and left blank in the gure. For each 6 level, the entries closest to the diagonal have the largest rates. These rates vary over a range of 10 exc dex for dierent levels. These rates are obtained by computing each terms in r and r in (5.8). included. ratepk from level k to level Figure 13 shows the total rates from/to each level. The blue circle shows the total rate of excitation/deexcitation from a given level. The red cross shows the total rate of excitation/deexcitation to a given level. The black square shows the total rate of ionization from a given level. The magenta plus shows the total rate of recombination to a given level. As predicted in Section 5.3, the rates 41 log10 rate ( k −> l ) 23 atom level l after collision 10 9 22 8 21 7 20 6 19 5 18 4 17 3 2 16 1 15 1 2 3 4 5 6 7 atom level k before collision 8 9 10 Figure 12: The total rates for excitation/deexcitation at equilibrium. The rates with k“l are not dened and left blank in the gure. of excitation/deexcitation vary over dierent levels. For high levels like 106 while those of ionization/recombination vary over 104 for k “ 9, 10, the rates of excitation/deexcitation are 100 times larger than those of ionization/recombination. Note that the total excitation/deexciation rate to/from level 10 is smaller that that of level 9. This is due to the fact that we only include 10 levels. If the more levels are included, we will have the opposite relation. 7 Conclusions and Future Work This paper presents an idealized model for excitation/deexcitation and ionization/recombination by electron impact in plasma. The energetic cross sections for the model are all given by analytic formulas and they are consistent with detailed balance. A Boltzmann type equation for the energy distribution function is then formulated and an H -theorem is derived. A conservative and equilibrium-preserving Direct Simulation Monte Carlo method is also developed for this system. The rates of recombination has a singularity as the energy of electron goes to 0. To overcome this diculty, four dierent methods are proposed and compared, among which the binary search method and the reduced rejection method give the highest eciency. A simulation method suers from two diculties: First, there are a large number of kinetic processes (i.e., excitation and deexcitation between any two atomic levels) with widely varying rates. Fair representation of all of these processes can require very small time steps. Second, random uctuations can exhaust the nite number of simulation particles in some state. In subsequent work, we plan to develop coarse-graining and a multi-scale simulation strategy, for example the level grouping 42 24 10 22 total rates 10 20 10 excitation/deexcitation from this level excitation/deexcitation to this level ionization from this level recombination to this level 18 10 16 10 1 2 3 4 5 6 7 8 9 10 levels Figure 13: The total rates for all collisions at equilibrium. method developed in [24], to reduce the eect of these two computational bottlenecks. The model and method here has been developed under a number of simplifying assumptions: The energy levels come from the Bohr model of hydrogen; only Hydrogen atoms and ions are included; Coulomb collisions have been neglected. Generalization to more realistic models will be further investigated. References [1] M. M. Becker, D. Lohagen, Derivation of moment equations for the theoretical description of electrons in non-thermal plasmas, Adv. Pure Math., vol. 3, pp. 343352, 2013. [2] R. E. Robson, K. F. Ness, Velocity distribution function and transport coecients of electron swarms in gases: spherical-harmonics decomposition of Boltzmann's equation, Phys. Rev. A, vol. 33, pp. 20682077, 1986. [3] R.E. Robson, R. Winkler, F. Sigeneger, Multiterm spherical tensor representation of Boltzmann's equation for a non-hydrodynamic weakly ionized plasma, Phys. Rev. E, vol. 65, p. 056410, 2002. [4] G. J. M. Hagelaar, L. C. Pitchford, Solving the Boltzmann equation to obtain electron transport coecients and rate coecients for uid models, Plasma Sources Sci. Techn., vol. 14, pp. 722733, 2005. [5] S. Longo, Monte-Carlo models of electron and ion transport in non-equilibrium plasmas, Plasma Sources Sci. Techn., vol. 9, pp. 468476, 2000. [6] S. Dujko, R. D. White, Z. Lj. Petrovic, Monte-Carlo studies of non-conservative electron J. Phys D, vol. 41, p. 245205, 2008. P. L. Bartlett and A. T. Stelbovics, Ionization of atoms by electron impact, Nuclear Instruments and Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and transport in the steady-state Townsend experiment, [7] 43 Associated Equipment, vol. 619, pp. 1 6, 2010. Frontiers in radiation physics and applica- tions:Proceedings of the 11th International Symposium on Radiation Physics. [8] Y. B. Zel'dovich and Y. P. Raizer, Phenomena. Physics of Shock Waves and High-Temperature Hydrodynamic New York: Dover Publications, Inc., 2002. [9] K. Steiner, Weighted particle methods solving kinetic equations for dilute ionized gases, in Flow Simulation with High-Performance Computers II (E. Hirschel, ed.), vol. 48 of Notes on Numerical Fluid Mechanics (NNFM), pp. 322339, Vieweg+Teubner Verlag, 1996. [10] T. Nedelea, B. Briehl, and H. M. Urbassek, Simulation study of the eect of recombination processes in an expanding plasma, Journal of Plasma Physics, vol. 71, pp. 589609, 10 2005. [11] A. Barmin and A. Kozlov, Structure of a steady-state ionization front in the plasma accelerator Fluid Dynamics, vol. 48, no. 4, pp. 556566, 2013. channel, [12] T. E. Glover, J. K. Crane, M. D. Perry, R. W. Lee, and R. W. Falcone, Measurement of velocity distributions and recombination kinetics in tunnel-ionized helium plasmas, Phys. Rev. Lett., vol. 75, pp. 445448, Jul 1995. [13] K. Nanbu, S. Yonemura, and A. Sasaki, Particle modeling of ionization and three-body recombination in fully ionized plasmas, Japan J. Appl. Phys., vol. 38, pp. 44604464, 1999. Comm. Comp. [14] F. Barekat and R. Caisch, Simulation with uctuating and singular rates, Phys., vol. 16, pp. 287306, 2014. The theory of stochastic processes. [15] D. R. Cox and H. D. Miller, Boca Raton, FL : Chapman and Hall/CRC, 1965. "First CRC reprint 2001" verso t.p. [16] W. M. Young and E. W. Elcock, Monte Carlo studies of vacancy migration in binary ordered alloys: I, Proceedings of the Physical Society, vol. 89, no. 3, p. 735, 1966. [17] D. T. Gillespie, A general method for numerically simulating the stochastic time evolution of coupled chemical reactions, Journal of Computational Physics, vol. 22, pp. 403434, 1976. [18] J.-L. Cambier and R. Caisch, A conservative scheme for the non-Maxwellian kinetics of preprint. Kinetic Theory of Particles and Photons. Berlin: collisional-radiative plasma with sub-cell resolution, [19] J. T. Oxenius, Springer-Verlag, 1986. [20] M. Baertschy, T. Rescigno, W. Isaacs, X. Li, and C. McCurdy, Electron-impact ionization of atomic hydrogen, Phys. Rev. A, vol. 63, p. 022712, 2001. The Theory of Atomic Collisions. [21] N. Mott and H. Massey, 1965. [22] G. Bird, Molecular Gas Dynamics. London: Oxford University Press, London: Oxford University Press, 1976. [23] G. Marsaglia, Generating discrete random variables in a computer, Commun. ACM, vol. 6, pp. 3738, Jan. 1963. [24] H. P. Le, A. R. Karagozian, and J.-L. Cambier, Complexity reduction of collisional-radiative kinetics for atomic plasma, Physics of Plasmas (1994-present), vol. 20, no. 12, pp. , 2013. 44 Appendix A Proof of Lemma 2.1 In Appendix A-D we provide the proof of several lemmas and propositions given in Section 3. For The reader's convenience, we restate these results in the appendices. The proof of Lemma 2.1 is given in this appendix. Lemma 2.1. Given ntot ą 0 εtot ą 0 the total ne , na and nk , @k the total number of free and bound electrons, and energy of the system, there are unique values of temperature Te and densities satisfying ne ` na “ ntot , ÿ eq 3 ne Te ` nk Lk ` neq ` L` “ εtot , 2 k with neq k Proof. neq ` “ ne and obtained using (2.6a)-(2.6b). For ease of notation, the sux eq is ignored, with the understanding that all quantities computed here are at thermodynamic equilibrium. For a given with temperature Te ntot , the total energy at equilibrium is ÿ 3 εpTe q “ ne Te ` nk Lk ` n` L` , 2 k with nk and n` given by (2.8a)(2.8b). We only need to show ε is monotonic in Te . Charge neutrality gives ntot ´ na “ ne “ n` “ n1{2 a γpTe q, with γpTe q “ CTe3{4 pZxa q´1{2 e´L` {p2Te q , where ´ ˘ ¯1{2 ` e 3{2 C “ 2g` 2πm h2 is a constant. The function γ can be shown to be monotonic in ř ´Lk {Te dγ 3 γ L` γ γ k gk Lk e ř “ ` ´ . ´Lk {Te dTe 4 Te Te 2Te 2Te2 k gk e We dene by x¨y the averaging over the atomic partition function Zxa , Te : (A.1) i.e. for ϕ “ pϕ1 , . . . , ϕkmax q, dene ř ´Lk {Te ÿ nk k ϕk gk e xϕy “ ř “ ϕk . ´Lk {Te na k gk e k The last term in (A.1) can be recognized as xLy. Since L` ą Lk , @k majorant for this expression „ ȷ dγ γ 3 L` xLy 3γ “ ` ´ ě ą 0. dTe 2Te 2 Te Te 4Te The solution for the neutral atom density is then na1{2 “ ´ γ 1a 2 ` γ `4ntot , 2 2 na “ ´a 4n2tot γ 2 `4ntot ` γ 45 ¯2 , (A.2) and γ ě 0, we can nd a with the following variation dna ´2na “a 2 dγ γ ` 4ntot ` γ 1` a ¸ γ ă 0. γ 2 ` 4ntot Te increases, and since dne ne monotonically increases with Te , i.e. dTe ě 0. Therefore, na ˜ monotonically decreases as ne “ ntot ´ na , this implies that The variation of the energy can be computed as follows dε 3 “ ne ` dTe 2 ˆ 3 Te ` L` 2 ˙ dne ÿ dnk ` Lk . dTe dTe k (A.3) The last term is not a-priori trivial; we expect the population of excited states to increase with (note that the ground state, being depleted, does not contribute to the sum since L1 ” 0), Te but it is entirely possible that the rate of changes of some levels increase while others decrease, especially as ionization proceeds preferentially from the upper states. Let us consider this variation, dnk dna x ´1 ´Lk {Te nk x ´1 ÿ nk Lk “ pZa q gk e ´ 2 pZa q gk Lk e´Lk {Te ` . dTe dTe Te Te2 k Identifying again the averaging over the atomic partition function Zxa , we obtain nk dna nk dnk “ ` 2 pLk ´ xLyq . dTe na dTe Te Inserting into (A.3), ¸ ÿ nk Lk dne ÿ nk ` ˘ 3 Te ` L` ´ ` L2k ´ Lk xLy 2 2 na dTe Te k k ˆ ˙ ˘ 3 dne na ` 3 Te ` L` ´ xLy ` 2 xL2 y ´ xLy2 . “ ne ` 2 2 dTe Te 3 dε “ ne ` dTe 2 ˜ The second term is nonnegative since L` ą Lk for any k and dne dTe ě 0. The third term is nonnegative due to Cauchy-Schwarz inequality. Therefore at equilibrium, the total energy is also monotonically increasing as a function of Te . This completes the proof of Lemma 2.1. Appendix B Proof of Lemma 3.6 This appendix gives the proof of Lemma 3.6. Lemma 3.6. Suppose f pEq is a smooth function and eq f eq , tneq k u, n` are the equilibrium functions dened by (2.5)-(2.8b), sharing the same total number density and total energy with n` . f , tnk u and If nk f0 n` f1 f2 eq eq “ eq eq eq , nk f0 n` f 1 f 2 for every E0 and Et , then f “ f eq , Proof. nk “ neq k , n` “ neq `. Denote Ck “ nk , neq k C` “ 46 n` , neq ` f f˜ “ eq . f (B.1) Then Ck f˜pE0 q “ C` f˜pE0 ´ Et qf˜pEt ´ Bk q, for any k , E0 Et satisfying Bk ď Et ď E0 . Et on both sides to obtain 1 ˜ ˜ ´f pE0 ´ Et qf pEt ´ Bk q ` f˜pE0 ´ Et qf˜1 pEt ´ Bk q “ 0, Take the derivative with respect to f˜1 pE0 ´ Et q f˜1 pEt ´ Bk q “ . f˜pE0 ´ Et q f˜pEt ´ Bk q Denote ϕ “ log f˜, then ϕ1 pE0 ´ Et q “ ϕ1 pEt ´ Bk q. This formula is valid for any where both α and β E0 , therefore ϕ1 is a constant, f˜ “ eαE`β , so that f˜ has the form are constants. Then Ck “ C` e´αBk `β “ C` e´αL` `β eαLk . Denote c0 “ C` e´αL` `β , then Ck “ c0 eαLk , This means that f , nk and n` in (B.1) are at equilibrium with temperature ˆ where Te is the temperature of C` “ c0 eαL` ´β . f eq . 1 ´α Te ˙´1 , From Lemma 2.1, the equilibrium temperature is unique, hence α “ 0. Then the conservation of number density gives f˜ ” 1, β“0 and Ck “ 1, c0 “ 1, so that C` “ 1, which nishes the proof. Appendix C Proof of Proposition 3.1 Proposition 3.1. (a) Suppose tf, nk , n` u is the solution to the ionization/recombination kinetic QeED “ QkED “ Q` ED “ 0. Then the total density and total equations, i.e. (2.19, 2.20) with energy are conserved, i.e., d d ntot “ εtot “ 0, dt dt ntot and εtot are dened by (2.4a) and (2.4b). Suppose tf, nk , n` u is the solution to the excitation/deexcitation kinetic equations, i.e. (2.19) ` e k (2.21) with QIR “ QIR “ QIR “ 0. Then the densities for each species and the total energy are where (b) conserved, i.e., d dt where εtot ż f pEq dE “ d ÿ d nk “ εtot “ 0, dt k dt is dened by (2.4b). 47 (c) Suppose tf, nk , n` u is the solution to the ionization/recombination and excitation/deexcitation kinetic equations (2.19, 2.20, 2.21). Then the total density and total energy are conserved, d d ntot “ εtot “ 0. dt dt Proof. For case (a), where only the ionization/recombinaiton processes are included, take ϕ“1 in (3.31) to obtain ˙ ÿ ij ˆ nk f0 n` f1 f2 eq eq ion Bt ne “ dE0 dEt . eq eq ´ eq eq eq nk f0 v0 σ n f n f f ` 0 1 2 k k Combining this with (3.32) leads to the conservation of total density of electrons ÿ d d d ntot “ ne ` nk “ 0. dt dt dt Now take ϕ“E in (3.31) to obtain ż Ef pEq dE “ Bt ÿij ˆ pE1 ` E2 ´ E0 q k nk f0 n` f1 f2 eq eq ´ eq eq eq nk f0 n` f1 f2 ˙ eq ion neq dE0 dEt . k f0 v0 σ Combining this with (3.32) and (3.33) leads to ż ÿ d d d d εtot “ Ef pEq dE ` Lk nk ` Lion n` dt dt dt dt k ˆ ˙ ij ÿ nk f0 n` f 1 f 2 eq eq ion “ pE1 `E2 ´E0 ´Lk `Lion q dE0 dEt . eq eq ´ eq eq eq nk f0 v0 σ n f n f f ` 0 1 2 k k Noting that d dt εtot E1 ` E2 ` Bk “ E0 from energy conservation at the microscopic level, which yields “ 0. For case (b), where only the excitation/deexcitataiton processes are included, take ϕ“1 and (3.37) to obtain d dt Now take ż f pEq dE “ d ÿ nk “ 0. dt k ϕ “ E in (3.35) to obtain ˆ ˙ ż ÿż nl f0 nu f1 eq eq exc Bt Ef pEq dE “ pE1 ´ E0 q dE0 . eq eq ´ eq eq nl f0 v0 σ n f n f u 0 1 l lău ϕk “ Lk in (3.37) gives ˙ ż ˆ ÿ ÿ d nl f 0 nu f1 eq eq exc ´ nk Lk “ dE0 . eq eq eq eq pLu ´ Ll qnl f0 v0 σ dt k n f n f u 0 1 l lău As for the energy of atoms, taking Therefore, ż ÿ d d d d εtot “ Ef pEq dE ` Lk nk ` Lion n` dt dt dt dt k ˙ ˆ ÿż nu f 1 nl f0 eq eq exc “ pE1 ´ E0 ` Lu ´ Ll q dE0 eq eq ´ eq eq nl f0 v0 σ n f n f u 0 1 l lău “ 0. The last equal sign comes from microscopic energy conservation, 48 E1 “ E0 ´∆Elu . in (3.35) For case (c), where all processes are included, the conservations of moments come from the fact that the IR operators and the ED operators conserve the moments separately. Appendix D Proof of Lemma 3.8 and Theorem 3.9 This appendix provides the proof of Lemma 3.8 and Theorem 3.9. Lemma 3.8. f is a solution to the excitation/deexcitation f pE, t “ 0q. Then d ÿ f pE ` j∆E, tq “ 0, for any E ě 0, dt j (i). Suppose (2.21) with initial value where the summation is over all integers (ii). f If function such that pE ` j∆Eq ě 0. f pEq “ f eq pE; Te qηpE; Te q, for some temperature Te E with the period ∆E , then η is given by (3.26). is in the form ηpE; Te q in (iii). There is a unique value of Proof. j kinetic equations (2.19) (i). For any Ẽ ą 0, Te and some periodic satisfying the energy conservation (3.27). take ϕpEq “ 8 ÿ δpE ´ Ẽ ´ j∆Eq (D.1) j“0 in (3.35) to obtain (3.38), using the fact that ϕpE1 q “ ϕpE0 ´ ∆Elu q “ ϕpE0 q, since ϕ (ii). If is a periodic function with period f ∆E ∆Elu and is an integer multiplication of ∆E . is in the form of (3.40), multiply both sides of (3.40) by (D.1) and integrate over tE ą 0u to obtain ż ÿ δpE ´ Ẽ ´ j∆Eqf pEq dE “ ż ÿ j δpE ´ Ẽ ´ j∆Eqf eq pEqηpEq dE, j ÿ f pẼ ` j∆Eq “ ÿ f eq pẼ ` j∆EqηpẼq. j j Therefore, ř ηpẼq “ ř j j ř f pẼ ` j∆Eq f eq pẼ ` j∆Eq “ f pẼ ` j∆E, t “ 0q . ř eq j f pẼ ` j∆Eq j (D.2) The last equality comes from (3.38). na and ne are constant in time. The total energy ( eq nk pTe q at temperature Te is (iii). With only excitation/deexcitation included, for f“ f eq pE; T e qηpE; Te q and ż εpTe q “ ␣ Ef eq pE; Te qηpE; Te q dE ` ÿ neq k pTe qLk ` n` L` . k For a given total energy ε is monotonic in εtot , to prove εpTe q “ εtot Te . 49 has a unique solution Te , we only need to show First notice that where x¨y d dTe pn` L` q “ 0 and from the proof of Lemma 2.1, ÿ ÿ ˘ d d eq na ` neq L “ Lk nk “ 2 xL2 y ´ xLy2 ě 0, k k dTe k dTe Te k is dened in (A.2). Next one can check 1 d eq f “ f eq dTe Te ˆ E 3 ´ T 2 ˙ , and ÿ d ´η d η “ ř eq f eq pE ` j∆Eq dTe dT f pE ` j∆Eq e j j ˆ ˙ ÿ ´η{Te E ` j∆E 3 eq “ ř eq ´ f pE ` j∆Eq . Te 2 j f pE ` j∆Eq j Hence, for f “ f eq pE; Te qηpE; Te q, after simplication, d d eq d f “η f ` f eq η dTe dTe dTe ř eq f eq η∆E j jf pE ` j∆Eq ř , “´ ¨ eq Te2 j f pE ` j∆Eq where the summation is over all integers r¨s j “ jpEq E ` j∆E ě 0, such that i.e. jě´ “ E ∆E , with ‰ representing the integer part. Then for the electron energy ż εe pTe q “ one has d ∆E εe “ ´ 2 dTe Te Here the For any ż ř j Ef pEqηpEq ř eq Ef eq pE; Te qηpE; Te q dE, jf eq pE ` j∆Eq dE f eq pE ` j∆Eq ř ż eq ∆E ÿ pj0 `1q∆E j jf pE ` j∆Eq eq ř dE “´ 2 Ef pEqηpEq eq Te j j0 ∆E j f pE ` j∆Eq 0 ř ż eq ∆E ∆E ÿ j jf pE ` j0 ∆E ` j∆Eq eq ř “´ 2 dE. pE ` j0 ∆Eqf pE ` j0 ∆EqηpEq eq Te 0 j f pE ` j∆Eq j0 ” ı “ E ‰ E`j0 ∆E summation in the numerator is over all integers j ě ´ “ ´j0 ´ ∆E . ∆E E P r0, ∆Eq, j denote ÿ Aj0 pEq “ jf eq pE ` j0 ∆E ` j∆Eq jě´j0 ´rE{∆Es ÿ “ pj 1 ´ j0 qf eq pE ` j 1 ∆Eq j 1 ě´rE{∆Es ÿ “ ÿ j 1 f eq pE ` j 1 ∆Eq ´ j0 j 1 ě´rE{∆Es f eq pE ` j 1 ∆Eq. j 1 ě´rE{∆Es Let «ř j0˚ “ j0˚ pEq “ j 1 ě´rE{∆Es j ř 1 f eq pE j 1 ě´rE{∆Es f 50 eq pE ` j 1 ∆Eq ` j 1 ∆Eq ff , then Aj0 pEq ď 0, for j0 ď j0˚ , and Aj0 pEq ą 0, for j0 ą j0˚ . Therefore ř ż eq d ∆E ∆E ÿ j jf pE ` j0 ∆E ` j∆Eq ˚ eq ř eq εe ě ´ 2 pE ` j0 ∆Eqf pE ` j0 ∆EqηpEq dE dTe Te 0 j f pE ` j∆Eq j0 ż ∆E ÿ d pE ` j0˚ ∆Eq “ f pE ` j0 ∆Eq dE dT e 0 j0 ż ∆E d ÿ “ pE ` j0˚ ∆Eq f pE ` j0 ∆Eq dE dTe j 0 0 ż ∆E ÿ d f pE ` j0 ∆E, t “ 0q dE. “ pE ` j0˚ ∆Eq dT e j 0 0 d ř d We have used (3.38) in the last step. Note that j0 f pE ` j0 ∆E, t “ 0q “ 0, one has dTe εe “ 0. dTe Therefore d d d ÿ eq d ε“ εe ` nk L k ` pn` L` q ě 0. dTe dTe dTe k dTe This ends the proof. Theorem 3.9. every l, u. Assume there is Suppose f pEq ∆E ą 0, such that ∆Elu is an integer multiplication of ∆E , for is a smooth function such that nu f1 nl f0 eq eq “ eq eq nl f 0 nu f1 holds for every l, u and E0 , in which (2.8a)(2.8b) for the temperature Te eq f eq “ f eq pE; Te q, tneq k u “ tnk pTe qu are of the form (2.5) and determined from (3.27). Then f pEq “ f eq pE; Te qηpEq, nk “ neq k pTe q, where ηpEq Proof. is the periodic function (3.26), with period ∆E . Denote Ck “ nk , neq k f f˜ “ eq . f Then Cl f˜pE0 q “ Cu f˜pE0 ´ ∆Elu q, for any l, u and E0 . E0 on both 1 1 ˜ ˜ Cl f pE0 q “ Cu f pE0 ´ ∆Elu q, Take the derivative with respect to sides to obtain f˜1 pE0 q f˜1 pE0 ´ ∆Elu q “ . f˜pE0 q f˜pE0 ´ ∆Elu q Denote ψ “ log f˜, then ψ 1 pE0 q “ ψ 1 pE0 ´ ∆Elu q. l, u and E0 . Therefore ψ 1 is a periodic function l and u, the period of ψ 1 is ∆E dened in (3.24). This formula is valid for any Since this is true for every 51 with period ∆Elu . Then ψpE0 q “ ψpE0 ´ ∆Eq ` C , for some constant C. Let α “ C{∆E , one has ψpE0 q ´ αE0 “ ψpE0 ´ ∆Eq ´ αpE0 ´ ∆Eq. Denote ηpEq “ exppψpEq ´ αEq, then η is a periodic function with period ∆E and f pEq is given by f pEq “ f pE; Te qf˜pEq “ f pE; Te q exppψpEqq “ eαE f eq pE; Te qηpEq. eq eq (D.3) Moreover, from Cl f˜pE0 q “ Cu f˜pE0 ´ ∆Elu q one has f˜pE0 ´ ∆Elu q eαLl Cl 1 “ “ α∆E “ αLu . lu Cu e e f˜pE0 q Choose a constant c0 such that C1 “ c0 eαL1 , nk “ then Ck neq k pTe q Ck “ c0 eαLk “ for any level k. Hence c0 eαLk neq k pTe q. p1{Te ´ αq´1 . From Lemma 3.8 (iii), the temperature is uniquely determined from the energy conservation, hence α “ 0. Combined with (D.3), f pEq and nk are in the form of (3.40) with temperature This completes the proof. Appendix E Algorithm for binary search method This appendix provides the detailed algorithm for the application of binary search method in the recombination problem. E.1 Padding by zeros For a given set tvk , k “ 1, . . . , N u, rst enlarge its size to N1 by padding zeros. N1 ě N . The method is easier to formulate for N1 , minimum power of 2 satisfying Here N1 is the a power of two, and padding by zeros also reserves some slots for possible new electrons generated in the ionization collisions. In the following we assume N is a power of 2 and N1 “ N , without loss of generality. E.2 Building a table Let L “ log2 N . Build a table to store the pre-computed summations, L´l j2ÿ sl,j “ vk , with 0 ď l ď L, 1 ď j ď 2l . k“pj´1q2L´l `1 Note that sL,j “ vj , there is no need to store 1 ` 2 ` ¨¨¨ ` 2 Moreover, the table tsl,j u sL,j . The total memory used for storing this table is L´1 L “ 2 ´ 1 “ N ´ 1. can be computed backwardly sl,j “ sl`1,2j´1 ` sl`1,2j . Hence we only need to carry out N summations. 52 E.3 Sampling a particle Here we describe the algorithm for sampling. r uniformly from ř p0, 1q k “ 1, q “ s0,1 “ j vj for l “ 1, . . . , L do q1 Ð sl,2k´1 p Ð qq1 if r ď p then k Ð 2k ´ 1 q Ð q1 r Ð pr Sample Let else k Ð 2k q Ð q ´ q1 r´p r Ð 1´p end if end for The variable k gives the solution for (5.4). The computational cost is OpLq “ Oplog2 N q. E.4 Update the table when some vk changes If vk changes, there is exactly one entry in a total number of L tsl,j u to be updated, for each l. k“ L ÿ βl 2l , l“0 one only needs to update sl,αl :“ sl,αl ` vknew ´ vkold , where α0 “ 1 l “ 0, . . . , L, and αl “ 2αl´1 ` βL`1´l ´ 1. This cost is also Hence we need to update entries in the table. More precisely, if we write the binary expansion of Oplog2 N q. 53 k,
© Copyright 2025 Paperzz