A Reduced-order NLTE Kinetic Model for Radiating Plasmas of Outer Envelopes of Stellar Atmospheres

, , and

Published 2017 April 3 © 2017. The American Astronomical Society. All rights reserved.
, , Citation Alessandro Munafò et al 2017 ApJ 838 126 DOI 10.3847/1538-4357/aa602e

Download Article PDF
DownloadArticle ePub

You need an eReader or compatible software to experience the benefits of the ePub3 file format.

0004-637X/838/2/126

Abstract

The present work proposes a self-consistent reduced-order NLTE kinetic model for radiating plasmas found in the outer layers of stellar atmospheres. A detailed collisional-radiative kinetic mechanism is constructed by leveraging the most up-to-date set of ab initio and experimental data available in the literature. This constitutes the starting point for the derivation of a reduced-order model, obtained by lumping the bound energy states into groups. In order to determine the needed thermo-physical group properties, uniform and Maxwell–Boltzmann energy distributions are used to reconstruct the energy population of each group. Finally, the reduced set of governing equations for the material gas and the radiation field is obtained based on the moment method. Applications consider the steady flow across a shock wave in partially ionized hydrogen. The results clearly demonstrate that adopting a Maxwell–Boltzmann grouping allows, on the one hand, for a substantial reduction of the number of unknowns and, on the other, to maintain accuracy for both gas and radiation quantities. Also, it is observed that, when neglecting line radiation, the use of two groups already leads to a very accurate resolution of the photo-ionization precursor, internal relaxation, and radiative cooling regions. The inclusion of line radiation requires adopting just one additional group to account for optically thin losses in the α, β, and γ lines of the Balmer and Paschen series. This trend has been observed for a wide range of shock wave velocities.

Export citation and abstract BibTeX RIS

1. Introduction

Stellar atmospheres are the outer gaseous layers of stars and are the locus of a broad domain of physical phenomena such as shock waves, winds, flares, coronal mass ejections, and magnetic reconnection (Priest 1982, Chapter 1; Hubeny & Mihalas 2014, Chapter 1). The accurate prediction of the behavior of stellar atmospheres is challenging for a variety of reasons. First, stellar atmospheres are dynamical objects whose state is continuously changing with time. This was recognized since the early days of observations from the asymmetric profiles of atomic lines in observed spectra, which revealed the existence of massive gas motions. Second, the low values of pressure and density are such that collisional rates among gas particles (e.g., atoms, molecules, free electrons) are not sufficient to ensure local thermodynamic equilibrium (LTE) (Hubeny & Mihalas 2014, Chapter 18). A further, and major, complexity comes from energy transfer by radiation, which plays a dominant role. Radiative transfer introduces a global coupling between the stellar material at all points in the atmosphere.

Within the context of a fluid description, the modeling of stellar atmospheres should be attacked from the governing equations of radiation hydrodynamics. Under conditions where Non-LTE (NLTE) prevails and relativistic effects can be ignored, these equations comprise the continuity equations for each gaseous species, the global momentum and energy equations, and the radiative transfer equation (RTE) (Oxenius 1986, Chapter 3). In the most accurate formulation, each bound state of atomic and molecular components (e.g., H(1s), H(2s)) is treated as a separate pseudo-species to allow for departures of the internal distribution function from local equilibrium (i.e., Maxwell–Boltzmann distribution). On the other hand, the translational degree of freedom is often assumed to be in equilibrium by prescribing a Maxwellian velocity distribution function at its own translational temperature. This approach is sometimes referred to as the State-to-State (StS) approach (Capitelli et al. 2000, Chapters 2–4; Nagnibeda & Kustova 2009, Chapter 2) and is equivalent to the full NLTE approach used in the theory of stellar atmospheres (with the only difference being the use of different translational temperatures). The radiation field is treated based on a line-by-line method to capture the finest details of radiative processes leading to line and continuous spectra. It is clear that this modeling strategy, despite its use, with substantial simplifications, in one-dimensional situations (Klein et al. 1976, 1978; Kneer & Nagakawa 1976; Carlsson & Stein 1992, 1995, 1997, 2002), becomes infeasible when moving to multidimensional configurations and when accounting for all possible opacity sources, not to mention the coupling with electromagnetic fields.

The present work is the first step in a long-term effort aimed at developing reduced-order models for astrophysical plasmas, with the final goal of enabling accurate predictions in unsteady NLTE multidimensional simulations. In this paper, a self-consistent and systematic method to reduce the complexity of the NLTE collisional-radiative kinetics is proposed. The formulation is based on the general Maximum Entropy (ME) principle framework recently developed by Liu et al. (2015), which has already been used with success in the study of collisional excitation, dissociation, and ionization in atomic and molecular gases (Panesi & Lani 2013; Munafò et al. 2014, 2015; Liu et al. 2015). The starting point is the set of StS governing equations for a two-temperature plasma, where the mass and energy source terms due to collisional and radiative processes are obtained self-consistently from the Boltzmann equation (i.e., kinetic equation for gas particles; Oxenius 1986, Chapter 3). Kinetic and thermodynamic data are taken from the most recent and accurate ab initio calculations and/or experiments. The complexity reduction of the NLTE kinetic mechanism is achieved by lumping the bound energy states into groups. Different grouping strategies are investigated, such as uniform and Maxwell–Boltzmann. The reduced set of governing equations is obtained by taking moments with respect to the internal energy of the StS governing equations. The grouping is not applied, for the moment, to the radiative transfer problem, which is still treated in a line-by-line fashion.

Applications consider the steady flow across radiative shock waves in partially ionized hydrogen. Hydrogen was selected as it represents the most abundant element in stellar atmospheres and the interstellar medium (Hubeny & Mihalas 2014, Chapter 1). A standing shock was chosen as the test case since (i) its numerical calculation requires much less computational time compared to time-dependent simulations (Carlsson & Stein 1992), and (ii) shock waves are common in astrophysical flows and play important roles, for instance, in pulsating stars, accretion disks, and the chromosphere of the Sun (Priest 1982, Chapter 6). Hence, their accurate prediction is of great interest, especially for comparison with observations. At the same time, if a reduced-order model already performs poorly when applied to a standing shock, it is almost certain that the same situation (if not worse) will occur under time-dependent conditions.

Radiative shock waves have been investigated for a long time (and still are) and were among the first fluid mechanical benchmarks to be studied based on the methods of NLTE radiation hydrodynamics (Mihalas & Mihalas 1984, Chapter 8). Some examples are the early works by Whitney & Skalafuris (1963), Skalafuris (1965, 1968a, 1968b), Murty (1971), and Clarke and coworkers (Clarke & Ferrari 1965; Farnsworth & Clarke 1971; Foley & Clarke 1973), and the more recent papers by Fadeyev & Gillet (1998, 2000, 2001), Fadeyev et al. (2002), Panesi et al. (2009, 2011), Kapper & Cambier (2011a), Kapper & Cambier (2011b), and Capitelli et al. (2013). Within the context of the inviscid flow approximation (e.g., no electron heat conduction), the structure of a radiative shock wave in an atomic plasma can be, in general, subdivided into three distinct regions (see Figure 1): (i) a radiative precursor ahead of the gasdynamic jump, (ii) an internal relaxation region behind the shock dominated by collisional excitation and ionization, and (iii) a radiative relaxation (or cooling) region where the temperature decreases due to optically thin radiation produced by radiative recombination. The radiative precursor can be further divided into the (i) far precursor and (ii) near precursor (Foley & Clarke 1973; Libermann & Velikovich 1986). In the former, non-equilibrium excitation may occur due to absorption of resonant radiation in atomic line wings. The latter is characterized instead by photo-ionization from the ground state (e.g., Lyman continuum) by photons emitted in the hot layers of the radiative cooling region, which propagate upstream. The extent (and also the existence) of the aforementioned zones strongly depends on the shock velocity, and the gas pressure and temperature in the free stream (Foley & Clarke 1973).

Figure 1.

Figure 1. Schematic of the structure of a radiative shock wave in atomic hydrogen plasma in the absence of transport phenomena (e.g., electron heat conduction). The figure displays the most important kinetic processes in the (i) precursor (far and near), (ii) internal relaxation, and (iii) radiative cooling regions. The length of the different regions is not to scale and has been chosen just for illustrative purposes.

Standard image High-resolution image

The present paper is organized as follows. Section 2 describes the StS model for partially ionized hydrogen. Section 3 introduces the reduced-order modeling technique with the aim of simplifying the complexity of the NLTE kinetic mechanism. The numerical method for solving the governing equations of the material gas and the radiation field is outlined in Section 4. Computational results are discussed and analyzed in Section 5. Conclusions are summarized in Section 6. MKS units are used throughout the manuscript.

2. StS Physico-chemical Modeling

This section describes the StS model for partially ionized hydrogen adopted in this work. Section 2.1 introduces the notation used throughout the manuscript and the main simplifying assumptions. The computation of thermodynamic properties is discussed in Section 2.2. Section 2.3 describes the set of kinetic processes (collisional and radiative) and the related calculation of the mass/energy production terms for the material gas and the emission/absorption coefficients for the radiation field. The governing equations for the problem under investigation (i.e., radiative shock wave in a plane-parallel medium) are given in Section 2.4.

Thermodynamic and kinetic data for the model hydrogen atom (e.g., energy levels, oscillator strengths, Einstein coefficients) have been taken from the NIST atomic database (Kramida 2010; Kramida et al. 2014). The individual atomic substates (e.g., H(1s), H(2s), H(2p)) include levels up to n = 40 (i.e., Rydberg states) and have been treated as separate species (see Table 2 in Appendix A for more details). The total number of states contained in the database is equal to 94.7 Radiative transitions account for electric dipole, electric quadrupole, and magnetic dipole radiation (Herzberg 1963, Chapter 1). The total number of transitions (for which both experimental and computational data are available in the database) amounts to 374 and includes the whole set of hydrogen series (not only Lymann, Balmer, and Paschen).

2.1. Notation and Assumptions

The material gas under investigation is made up of three chemical components: free electrons ${e}^{-}$, hydrogen atoms ${\rm{H}}$, and protons ${{\rm{H}}}^{+}$. These components constitute the set ${ \mathcal C }=\{{e}^{-},{\rm{H}},{{\rm{H}}}^{+}\}$. The heavy-particle components (i.e., hydrogen atoms and protons) are stored in the subset ${{ \mathcal C }}_{{\rm{h}}}=\{{\rm{H}},{{\rm{H}}}^{+}\}$. The hydrogen bound states (sorted by increasing energy) are denoted by the global index i and stored in set ${ \mathcal I }$. Based on an StS approach, the bound states ${\rm{H}}(i)$ are treated as separate pseudo-species (Bates et al. 1962a, 1962b; Capitelli et al. 2000). The related statistical weights and formation energies are indicated, respectively, by the symbols ${g}_{i}$ and ${E}_{i}$. The corresponding quantities for the protons are ${g}_{+}\,=\,1$ and ${E}_{+}=13.59\,\mathrm{eV}$. The StS species are stored in the species set ${ \mathcal S }=\{{e}^{-},{\rm{H}}(i),{{\rm{H}}}^{+};i\in { \mathcal I }\}$.

In order to make the problem tractable, simplifying assumptions are introduced. For the adopted pressure and temperature conditions, it is possible to use Boltzmann statistics. Each species is modeled as a thermally perfect gas by disregarding collective plasma effects such as pressure ionization and multiple charge–charge interactions (Rogers 1974; Rogers et al. 1996; Hubeny & Mihalas 2014). Further, the velocity distribution function of each species is taken to be Maxwellian at its own translational temperature. For heavy particles, this temperature is taken to be a common heavy-particle temperature ${T}_{{\rm{h}}}$. This assumption is justified in light of (i) the efficient energy transfer in collisions between particles with similar masses and (ii) the large cross-section for resonant charge transfer in hydrogen–proton collisions (i.e., ${\rm{H}}+{{\rm{H}}}^{+}={{\rm{H}}}^{+}+{\rm{H}}$) (Mihalas & Mihalas 1984, Chapter 8). On the other hand, collisions between heavy particles and free electrons are energetically inefficient, due to the large mass disparity. This motivates the introduction of a separate free-electron temperature ${T}_{{\rm{e}}}$. The adopted StS model is therefore a two-temperature plasma model where thermal non-equilibrium effects between heavy particles and free electrons are taken into account through the macroscopic parameters ${T}_{{\rm{h}}}$ and ${T}_{{\rm{e}}}$, respectively (Appleton & Bray 1964; Murty 1971). In this work, it is further assumed that the plasma is locally neutral. Electromagnetic fields and transport phenomena are also neglected.

2.2. Thermodynamics

In view of the assumptions introduced in the Section 2.1, the velocity distribution functions of free electrons, hydrogen bound states, and protons are

Equation (1)

Equation (2)

Equation (3)

where the quantities ns, $s\in { \mathcal S }$, and mc, $c\in { \mathcal C }$, denote, respectively, the species number densities and the mass of the chemical components. The symbols ${k}_{{\rm{B}}}$ stands for Boltzmann's constant, whereas the vectors ${{\boldsymbol{c}}}_{s}$, $s\in {\boldsymbol{s}}$, and ${\boldsymbol{v}}$ represent the molecular velocities and the bulk velocity of the material gas, respectively.

Thermodynamic properties (e.g., pressure, energy density) of the single species or the whole mixture can be computed based on suitable moments of the velocity distribution functions (Equations (1)–(3)) (Ferziger & Kaper 1972, Chapter 1). The gas pressure is obtained by taking the average of the sum of the diagonal terms of the pressure tensor and reads

Equation (4)

where the peculiar velocities are ${{\boldsymbol{C}}}_{s}={{\boldsymbol{c}}}_{s}-{\boldsymbol{v}}$, $s\in { \mathcal S }$. The evaluation of the integrals in Equation (4) leads to Dalton's law for a thermal non-equilibrium plasma:

Equation (5)

where the free-electron and heavy-particle partial pressures are, respectively, ${p}_{{\rm{e}}}={n}_{{\rm{e}}}{k}_{{\rm{B}}}{T}_{{\rm{e}}}$ and ${p}_{{\rm{h}}}={n}_{{\rm{h}}}{k}_{{\rm{B}}}{T}_{{\rm{h}}}$. Quantity ${n}_{{\rm{h}}}$ stands for the heavy-particle number density and is obtained by summing the contributions of the hydrogen bound states and protons, ${n}_{{\rm{h}}}={\sum }_{i\in { \mathcal I }}{n}_{i}+{n}_{+}$, where ${n}_{+}\,=\,{n}_{{\rm{e}}}$ in light of the assumed charge neutrality (see Section 2.1).

The heavy-particle and free-electron thermal energy densities are defined as

Equation (6)

Equation (7)

Using Equations (4) and (5), the integrals (6) and (7) can be evaluated in a straightforward manner to give

Equation (8)

The gas total energy density is obtained by summing the thermal contribution from free electrons and heavy particles, and the kinetic contribution of the gas as a whole, $\rho E=\rho {e}_{{\rm{h}}}+\rho {e}_{{\rm{e}}}+\rho {\boldsymbol{v}}\cdot {\boldsymbol{v}}/2$. The mass density is evaluated based on the species number densities as $\rho ={\sum }_{s\in { \mathcal S }}{n}_{s}\,{m}_{s}$. The total enthalpy density follows by adding the pressure to the total energy density, $\rho H=\rho E+p$.

Under complete thermodynamic equilibrium, all temperatures are equal (i.e., ${T}_{{\rm{h}}}={T}_{{\rm{e}}}=T$) and the population of hydrogen bound states is given by the Boltzmann distribution law:

Equation (9)

$i\in { \mathcal I }$, where the hydrogen number density and internal partition function are ${n}_{{\rm{H}}}={\sum }_{i\in { \mathcal I }}{n}_{i}$ and ${Z}_{{\rm{H}}}^{{\rm{i}}}(T)\ ={\sum }_{i\in { \mathcal I }}{g}_{i}\exp (-{E}_{i}/{k}_{{\rm{B}}}T)$, respectively. The hydrogen number density, ${n}_{{\rm{H}}}$, is related to that of free electrons and protons by Saha's equation and charge conservation (i.e., ${n}_{{\rm{e}}}={n}_{+}$):

Equation (10)

where the translational partition functions per unit volume are ${Z}_{c}^{{\rm{t}}}{(T)=(2\pi {m}_{c}{k}_{{\rm{B}}}T)}^{3/2}/{h}_{{\rm{P}}}^{3}$, $c\in { \mathcal C }$, where ${h}_{{\rm{P}}}$ is Planck's constant. The quantity ${g}_{{\rm{e}}}=2$ is the electron degeneracy accounting for its spin. Under a general NLTE situation, Equations (9) and (10) no longer hold. However, as shown in the results, Section 5, in the internal relaxation region behind the shock, the population of high-lying hydrogen bound states can be approximately described by a Boltzmann distribution (9) at its own internal temperature before equilibrium is reached. This motivates the introduction of reduced-order models, treated in Section 3.1, which represent the main subject of the present work.

Before ending the present discussion on thermodynamics, it should be recalled that, when studying the interaction between matter and radiation, one should in general account for the momentum and energy content of photons (Oxenius 1986, Chapter 7). The former plays an important role, for instance, in stellar interiors where up to 50% of the pressure may be due to radiation (Rogers et al. 1996). However, as shown in the work by Fadeyev & Gillet (1998, 2000), the radiant pressure and energy density are negligible compared to the corresponding gas values for the conditions adopted in this work.

2.3. Kinetics

The present section describes the NLTE kinetic model for atomic hydrogen plasmas used in this work. This StS model provides the basis for the reduced-order models developed in Section 3.1 and accounts for both collisional and radiative processes. The following subsections provide a description of their modeling. It is worth mentioning that, in the literature, other StS models for hydrogen plasmas exist such as the one developed by D'Ammando et al. (2010) and Colonna et al. (2012).

The calculation of the mass production and energy transfer terms due to collisional and radiative processes is obtained using the velocity distribution functions (1)–(3) in the moments of the Boltzmann equation for gas particles (for the details, the reader may consult Chapter 4 of the book by Oxenius 1986).

2.3.1. Collisional Processes

Collisional processes account for inelastic/elastic transitions, due to collisions between heavy particles and free electrons and comprise

  • (i)  
    excitation (e) and ionization (i) by electron impact:
    Equation (11)
  • (ii)  
    elastic (el) energy exchange between heavy particles and free electrons.

Excitation and ionization by heavy-particle impact have been neglected since, as shown in Section 5, the number of free electrons produced by photo-ionization in the precursor guarantees that the collisional kinetics is dominated by free electrons in the internal relaxation region.

The species production terms due to excitation and ionization are obtained by taking the zeroth-order moment (i.e., mass) of the collision operators for excitation and ionization in the Boltzmann equation for free electrons, hydrogen bound states, and protons (Oxenius 1986, Chapter 2):

Equation (12)

Equation (13)

$i\in { \mathcal I }$, where the acronym "col" (i.e., collisional) has been introduced to distinguish from the mass production terms due to radiative processes treated in Section 2.3.2. The endothermic rate coefficients for collisional excitation and ionization are evaluated based on the assumption of stationary heavy particles:

Equation (14)

Equation (15)

$i,j\in { \mathcal I }$, where the quantities ${\sigma }_{{ij}}^{{\rm{E}}}$ and ${\sigma }_{i}^{{\rm{I}}}$ denote, respectively, the total cross-section for excitation and ionization, with the corresponding threshold energies being defined as ${E}_{{ij}}={E}_{j}-{E}_{i}$ and ${E}_{+i}={E}_{+}-{E}_{i}$. The symbol ε stands for the free-electron energy. The exothermic rate coefficients for de-excitation and three-body recombination are related to those for excitation and ionization, respectively, via micro-reversibility (Oxenius 1986, Equations (2.3.10) and (2.3.16)):

Equation (16)

$i,j\in { \mathcal I }$. It is worth mentioning that the free-electron temperature, ${T}_{{\rm{e}}}$, is the only temperature appearing in the rate coefficients (14)–(16), due to the hypothesis of stationary heavy particles. In the present work, the cross-section for electron impact excitation and ionization have been modeled according to Drawin's semi-classical formula (Drawin 1963), with the absorption oscillator strengths taken from the NIST atomic database (Kramida 2010; Kramida et al. 2014).

The volumetric time rate of change of free-electron energy due to excitation, ionization, and elastic collisions with heavy particles can be written as ${{\rm{\Omega }}}_{{\rm{e}}}^{\mathrm{col}}={{\rm{\Omega }}}_{{\rm{e}}}^{{\rm{EL}}}+{{\rm{\Omega }}}_{{\rm{e}}}^{{\rm{E}}}+{{\rm{\Omega }}}_{{\rm{e}}}^{{\rm{I}}}$, where the individual contributions are obtained by taking the second-order moment (i.e., energy) of the related collision operators in the Boltzmann equation for free electrons:

Equation (17)

Equation (18)

where the effective collision frequency for elastic energy transfer in electron-heavy interactions reads (Petschek & Byron 1957; Desloge 1962; Mitchner & Kruger 1973):

Equation (19)

The quantities ${{\rm{\Omega }}}_{{\rm{e}}c}^{(1,1)}$ denote the electron-heavy collision integrals in the first-order Laguerre–Sonine polynomial expansion (Devoto 1966, 1967; Ferziger & Kaper 1972). The collision integrals ${{\rm{\Omega }}}_{{\rm{e}}c}^{(1,1)}$ are often referred to as diffusion cross-sections as they appear in the electron-heavy binary diffusion coefficients for ionized gas. For interactions between free electrons and hydrogen, the collision integral ${{\rm{\Omega }}}_{{\rm{e}}{\rm{H}}}^{(1,1)}$ has been evaluated based on the work by Bruno et al. (2010). For interactions between free electrons and protons, a screened Coulomb potential has been used to evaluate ${{\rm{\Omega }}}_{{\rm{e}}{{\rm{H}}}^{+}}^{(1,1)}$ (Devoto 1966, 1967; Bruno et al. 2010).

The functional form of the excitation and ionization energy transfer terms in Equation (18) reflect the physical fact that, under the assumption of stationary heavy particles, the only active source/sink of free-electron energy in inelastic and ionizing collisions is the change of heavy-particle formation energies.

2.3.2. Radiative Processes

The modeling of radiation in plasmas requires accounting for transitions characterized by absorption, emission, and scattering of light. Since the main focus of this work is the complexity reduction of the NLTE kinetics, scattering is not taken into account.

The radiative processes leading to emission and absorption of light can be subdivided into three groups: bound–bound (bb), bound–free/free–bound (bf/fb), and free–free (ff) (Oxenius 1986, Chapter 1; Hubeny & Mihalas 2014, Chapter 5). The radiative processes considered in this work are

  • (i)  
    spontaneous emission/absorption and induced emission (bb):
    Equation (20)
  • (ii)  
    photo-ionization and spontaneous/induced radiative recombination (bf/fb):
    Equation (21)
  • (iii)  
    spontaneous/induced and inverse bremsstrahlung for protons (ff).

bb transitions—The monochromatic emission and absorption coefficients due to bb radiation are (Oxenius 1986, Equations (3.2.9) and (3.2.15)):

Equation (22)

where the wavelength associated with the transition $j\to i$ is ${\lambda }_{{ij}}={h}_{{\rm{P}}}c/{E}_{{ji}}$, with c being the speed of light and the energy difference ${E}_{{ji}}$ having the same definition as in Equations (14) and (16) (i.e., ${E}_{{ji}}={E}_{j}-{E}_{i}$). The symbols ${A}_{{ji}}$, ${B}_{{ji}}$, and ${B}_{{ij}}$ (with $i\lt j$) denote, respectively, the Einstein coefficients for spontaneous emission, induced emission, and absorption. The Einstein coefficients are not independent of each other and satisfy the Einstein–Milne relations (Oxenius 1986, Equations (1.6.31) and (1.6.32)):

Equation (23)

$i,j\in { \mathcal I }$. The functions ${\phi }_{\lambda }^{{ji}}$ and ${\psi }_{\lambda }^{{ij}}$ in Equation (22) are, respectively, the monochromatic line emission and absorption profiles. In this work, it has been assumed that the emission and absorption profiles coincide (i.e., complete redistribution) and are described using a Voigt function by accounting for natural, Doppler, and collisional broadening (Hubeny & Mihalas 2014, Chapter 8). Doppler broadening has been evaluated at the heavy-particle temperature ${T}_{{\rm{h}}}$ as explained in Hubeny & Mihalas (2014), Equation (8.18). Collisional broadening accounts for collisions between charged particles (i.e., Stark broadening) and has been computed at the free-electron temperature ${T}_{{\rm{e}}}$ based on the work of Cowley (1971). The effects of resonance and pressure broadening and Doppler shifts due to macroscopic gas motion have not been taken into account. In order to speed up the calculations, the numerical evaluation of the Voigt function (i.e., convolution between Gaussian and Lorentzian profiles) has been accomplished using the curve fit by Whiting (1968).

Before proceeding further, it is worth spending some words on the definition of the emission and absorption coefficients given in Equation (22). In the aforementioned expression, induced/stimulated emission is treated as a negative absorption. However, according to the principles of electrodynamics, the true monochromatic emission coefficient should account for both the spontaneous and the induced/stimulated contributions (Oxenius 1986, Chapter 3). Hence, the expressions given in Equation (22) should be interpreted as the phenomenological emission and absorption coefficients. The treatment of induced/stimulated emission as a negative absorption is adopted only with the purpose of isolating the plasma optical properties, which do not explicitly depend on the local radiation field (i.e., monochromatic intensity). The convenience of this approach becomes clear when writing down the RTE (see Section 2.4). In the present work, the same convention is also used in the definition of the plasma optical properties due to bf/fb and ff radiation.

bb transitions lead to a change in the occupation numbers of hydrogen bound states, for which the related mass production terms are (Oxenius 1986, Equation (2.4.13)):

Equation (24)

$i\in { \mathcal I }$, where the mass production (superscript m) Einstein coefficients, defined by ${A}_{{ji}}^{{\rm{m}}}={A}_{{ji}}$ (with $i\lt j$) and ${B}_{{ij}}^{{\rm{m}}}={B}_{{ji}}\,{\lambda }_{{ij}}^{2}/c$, have been introduced for convenience. The quantity ${{\rm{\Phi }}}_{{ij}}$ is defined by the line-shape integral:

Equation (25)

$i,j\in { \mathcal I }$, where the symbol ${J}_{\lambda }$ denotes the average monochromatic intensity. The former is obtained by the integration of the (directionally dependent) monochromatic intensity ${I}_{\lambda }$ over all directions, ${J}_{\lambda }=1/4\pi \oint {I}_{\lambda }d{\rm{\Omega }}$, with Ω being the solid angle (Hubeny & Mihalas 2014, Chapter 11). The changes in the occupation numbers of hydrogen bound states are accompanied by a transfer of energy between the material gas and the radiation field. The volumetric time rate of loss of matter energy due to bb transitions is obtained by integrating the quantity ${\varepsilon }_{\lambda }^{{\rm{BB}}}-{I}_{\lambda }\,{\kappa }_{\lambda }^{{\rm{BB}}}$ over all wavelengths and directions (Oxenius 1986, Chapter 3). Performing the required integrations, one obtains

Equation (26)

where the energy transfer (e) Einstein coefficients are defined as ${A}_{{ji}}^{{\rm{e}}}={E}_{{ji}}\,{A}_{{ji}}$ (with $i\lt j$) and ${B}_{{ji}}^{{\rm{e}}}={h}_{{\rm{P}}}{\lambda }_{{ij}}{B}_{{ji}}$.

BF/FB transitions—The monochromatic emission and absorption coefficients due to bf and fb radiation are (Oxenius 1986, Equations (3.2.28) and (3.2.32)):

Equation (27)

Equation (28)

where the ionization threshold is defined, as in Equations (15) and (16), via the relation ${E}_{+i}={E}_{+}-{E}_{i}$. The quantity ${\sigma }^{{\rm{PI}}}$ denotes the total photo-ionization cross-section. In the present work, the former has been evaluated based on Kramer's formula (Zel'dovich & Raizer 1967, Equation (5.34); Hubeny & Mihalas 2014, Equation (7.84)).

The mass production terms for free electrons, protons, and hydrogen bound states due to photo-ionization and radiative recombination can be expressed as (Oxenius 1986, Equations (2.4.22) and (2.4.27)):

Equation (29)

Equation (30)

$i\in { \mathcal I }$, where the rate coefficients for photo-ionization and radiative recombination (accounting for both spontaneous (s) and induced (i) contributions) are

Equation (31)

Equation (32)

Equation (33)

$i\in { \mathcal I }$, with the threshold wavelength for photo-ionization/radiative recombination being ${\lambda }_{+i}={h}_{{\rm{P}}}c/{E}_{+i}$.

By analogy with the procedure outlined above for bb radiation, the volumetric time rate of loss of matter energy due to bf/fb radiation is obtained through the integration of the quantity ${\varepsilon }_{\lambda }^{{\rm{FB}}}-{I}_{\lambda }\,{\kappa }_{\lambda }^{{\rm{BF}}}$ over all wavelengths and directions:

Equation (34)

It should be noted that the volumetric energy loss term (34) refers to the whole gas, which includes free electrons and heavy particles. The adoption of an additional energy equation for free electrons (see Section 2.4) requires the evaluation of the corresponding quantity ${{\rm{\Omega }}}^{{\rm{BF}}/{\rm{FB}}}$ for the free-electron gas alone. This is accomplished in a straightforward manner by taking the second-order moment of the collision operator for photo-ionization/radiative recombination of the Boltzmann equation for free electrons. Under the assumption of Maxwellian free electrons at temperature ${T}_{{\rm{e}}}$ and stationary heavy particles, the final result is

Equation (35)

ff transitions—The monochromatic emission and absorption coefficients due to ff radiation produced by encounters between free electrons and protons are (Oxenius 1986, Equations (3.2.41) and (3.2.42)):

Equation (36)

where the quantities ${\epsilon }_{0}$ and ${q}_{{\rm{e}}}$ denote, respectively, the vacuum permittivity and the electron charge.

The volumetric time rate of loss of matter energy due to ff radiation is obtained, as done before, by integrating quantity ${\varepsilon }_{\lambda }^{{\rm{FF}}}-{I}_{\lambda }\,{\kappa }_{\lambda }^{{\rm{FF}}}$ over all wavelengths and directions:

Equation (37)

By collecting the results through Equations (22)–(37), it is possible to write down the mass production terms due to radiative transitions for free electrons, protons, and hydrogen bound states as ${\omega }_{{\rm{e}}}^{\mathrm{rad}}={\omega }_{{\rm{e}}}^{{\rm{BF}}/{\rm{FB}}}$, ${\omega }_{+}^{\mathrm{rad}}={\omega }_{+}^{{\rm{BF}}/{\rm{FB}}}$, and ${\omega }_{i}^{\mathrm{rad}}={\omega }_{i}^{{\rm{BB}}}+{\omega }_{i}^{{\rm{BF}}/{\rm{FB}}}$, respectively. The corresponding energy loss rates for the whole gas and free electrons alone are ${{\rm{\Omega }}}^{\mathrm{rad}}={{\rm{\Omega }}}^{{\rm{BB}}}+{{\rm{\Omega }}}^{{\rm{BF}}/{\rm{FB}}}+{{\rm{\Omega }}}^{{\rm{FF}}}$ and ${{\rm{\Omega }}}_{{\rm{e}}}^{\mathrm{rad}}={{\rm{\Omega }}}_{{\rm{e}}}^{{\rm{BF}}/{\rm{FB}}}+{{\rm{\Omega }}}^{{\rm{FF}}}$, respectively.

2.4. Governing Equations

The steady flow across a normal shock wave of a two-temperature radiating plasma is governed by the species continuity equations, the global momentum and energy equations, and the free-electron energy equation. In the absence of transport phenomena, the former set of equations reads (Zel'dovich & Raizer 1967, Equation (7.30); Murty 1971, Equation (1); see also Appendix C):

Equation (38)

$s\in { \mathcal S }$, where the mass production terms are given by the sum of the collisional and radiative contributions, ${\omega }_{s}={\omega }_{s}^{\mathrm{col}}+{\omega }_{s}^{\mathrm{rad}}$. The quantity u stands for the flow velocity measured in the shock wave reference frame. Gravity terms have not been taken into account as the length scales of the radiative shock waves investigated in this work (e.g., 101–102 m) are much smaller than those typical of stellar atmospheres (e.g., 106 m).

It is known that electron heat conduction, neglected in Equation (38), may play an important role in shaping the temperature profile in the precursor (i.e., conduction precursor) (Jukes 1957; Shafranov 1957; Imshennik 1962; Jaffrin & Probstein 1964; Fadeyev et al. 2002). However, for the standing shocks studied in this work, electron heat conduction is influenced little by the dynamics of excited electronic states,8 and thus is expected to play a minor role compared to radiation and chemistry on the accuracy of a reduced-order NLTE model.

It is worth noticing that, by assuming zero velocity and adding gravity terms in Equation (38), one may retrieve the structural equations for the material gas within a one-dimensional plane-parallel (and static) stellar atmosphere. As a matter of fact, under the above assumption, the species continuity equations reduce to ${\omega }_{s}=0$, which is precisely the condition of statistical equilibrium, whereas the global momentum and energy equations reduce to the requirements of hydrostatic and radiative equilibrium, respectively (i.e., $\partial p/\partial x=-\rho g$ and ${{\rm{\Omega }}}^{\mathrm{rad}}=0$).

For a radiating gas, the flow-governing Equations (38) must be coupled with the RTE (Mihalas & Mihalas 1984, Chapters 7 and 8). The former can be thought of as the kinetic equation for a photon gas and describes the evolution in space and time of a radiation field, due to emission, absorption, and scattering of light. In the case of a plane-parallel non-scattering medium under steady-state conditions, the RTE reads

Equation (39)

where the quantity $\mu \in [-1,1]$ stands for the cosine of the angle between the line of sight and the x-axis. The (total) emission and absorption coefficients in the RTE (39) are obtained by summing the individual contributions due to the bb, bf/fb, and ff transitions as ${\varepsilon }_{\lambda }^{}={\varepsilon }_{\lambda }^{{\rm{BB}}}+{\varepsilon }_{\lambda }^{{\rm{FB}}}+{\varepsilon }_{\lambda }^{{\rm{FF}}}$ and ${\kappa }_{\lambda }^{}={\kappa }_{\lambda }^{{\rm{BB}}}+{\kappa }_{\lambda }^{{\rm{BF}}}+{\kappa }_{\lambda }^{{\rm{FF}}}$, respectively. In radiative transfer problems within plane-parallel media, it is often convenient to transform the RTE (39) to a second-order differential equation as proposed by Feautrier (1964):

Equation (40)

where the quantity ${P}_{\lambda \mu }$ is defined as ${P}_{\lambda \mu }=({I}_{\lambda \mu }+{I}_{\lambda -\mu })/2$, and the monochromatic source function is given by the ratio of emission and absorption coefficients, ${S}_{\lambda }={\varepsilon }_{\lambda }^{}/{\kappa }_{\lambda }^{}$. The symbol $d{\tau }_{\lambda }$ denotes the infinitesimal monochromatic optical thickness increment, $d{\tau }_{\lambda }={\kappa }_{\lambda }^{}\,{dx}$. For the RTE (40), the range of the angular variable μ is restricted to $[0,1]$ (Hubeny & Mihalas 2014, Chapter 11). When written in terms of the newly introduced unknown ${P}_{\lambda \mu }$, the average monochromatic intensity becomes simply

Equation (41)

3. Reduced-order Modeling

In order to simplify the complexity of the NLTE kinetic mechanism, the energy levels (i.e., species) of the StS model described in Section 2 are lumped into groups. The governing equations for the reduced-order model are obtained using a moment method after prescribing a distribution within each group. The general procedure is explained below.

3.1. Level Grouping: ME Model

Following the work by Liu et al. (2015), the logarithm of the normalized population within a given group k is written as a polynomial in the internal energy ${E}_{i}$:

Equation (42)

$i\in {{ \mathcal I }}_{k}$, $k\in { \mathcal K }$, where the sets ${{ \mathcal I }}_{k}$ and ${ \mathcal K }$ denote, respectively, the energy levels within group k and the group indices. In the present work, terms that are second and higher order in energy are neglected. Under these circumstances, one needs to determine only the quantities ${\tilde{\alpha }}_{k}$ and ${\tilde{\beta }}_{k}$. These are related to the group populations, ${\tilde{n}}_{k}$, and average energies, ${\tilde{E}}_{k}$, by the following moment constraints on particle number and energy (Liu et al. 2015):

Equation (43)

$k\in { \mathcal K }$. Substituting Equation (42) into the moment constraints (43) gives the implicit relation linking ${\tilde{\alpha }}_{k}$ and ${\tilde{\beta }}_{k}$ to the group populations and energies:

Equation (44)

$k\in { \mathcal K }$. After introducing, for the sake of convenience, group temperatures ${T}_{k}=-1/{k}_{{\rm{B}}}{\tilde{\beta }}_{k}$ and partition functions ${\tilde{Z}}_{k}^{}({T}_{k})\ ={\sum }_{i\in {{ \mathcal I }}_{k}}{g}_{i}\exp ({\tilde{\beta }}_{k}{E}_{i})$, it is possible to re-write Equation (44) as

Equation (45)

$k\in { \mathcal K }$. Substituting the first of Equation (45) into Equation (42) leads finally to

Equation (46)

$i\in {{ \mathcal I }}_{k}$, $k\in { \mathcal K }$. Equation (46) shows that retaining terms up to first order in energy in Equation (43) is equivalent to assuming a Maxwell–Boltzmann distribution within each group. The Maxwell–Boltzmann distribution is the LTE distribution for which the entropy is maximum (Groot & Mazur 2011, Chapter 1). This is the reason that reduced-order models developed based on Equation (42) are called ME models (Liu et al. 2015). The model corresponding to Equation (46) is the Maximum Entropy Linear (MEL) model as only linear terms are retained in the energy polynomial (40). The MEL model reduces to the Maximum Entropy Uniform (MEU) model when taking the limit of infinite group temperatures (i.e., ${\tilde{\beta }}_{k}=0$) or, equivalently, when retaining only the zeroth-order energy term in Equation (42). In this case, Equation (46) reduces to

Equation (47)

$i\in {{ \mathcal I }}_{k}$, $k\in { \mathcal K }$, where the group partition function is now the sum of the statistical weights of the energy levels within the group, ${\tilde{Z}}_{k}={\sum }_{i\in {{ \mathcal I }}_{k}}{g}_{i}$. It is worth mentioning that the MEU distribution (47) does not allow equilibrium (i.e., Maxwell–Boltzmann distribution) to be retrieved.

3.2. Moment Equations

The governing equations for the ME model are obtained by taking the moments with respect to the energy ${E}_{i}$ of the species continuity equations (Liu et al. 2015). As explained in Section (3.1), in the present work only the zeroth and first-order moments are needed:

Equation (48)

$k\in { \mathcal K }$. Using the moment constraints (43), Equation (48) becomes

Equation (49)

$k\in { \mathcal K }$, where the group mass production and energy transfer terms are defined as ${\tilde{\omega }}_{k}={\sum }_{i\in {{ \mathcal I }}_{k}}{\omega }_{i}$ and ${\tilde{{\rm{\Omega }}}}_{k}={\sum }_{i\in {{ \mathcal I }}_{k}}{\omega }_{i}{E}_{i}$, respectively. In the case of the MEU model, the group number densities ${\tilde{n}}_{k}$ are the only unknowns. Thus, only the first of Equation (49) is needed. The complete set of flow-governing equations for the MEU/MEL models is obtained based on those for the StS model (38) by replacing the species continuity equations for the hydrogen bound states with the moment equations (49). The related expressions for thermodynamic properties, mass/energy production terms, and emission/absorption coefficients are given in Appendix B.

The MEU and MEL models have already been successfully applied to the study of collisional excitation, dissociation, and ionization in atomic and molecular gases (Panesi & Lani 2013; Munafò et al. 2014, 2015; Liu et al. 2015). In that work, the determination of the states contained in a given group can be based, for instance, on an even subdivision of the internal energy ladder. This is justified by the fact that the rate coefficients for inelastic collisional processes (e.g., electron impact ionization) are larger for states with similar energy. The inclusion of radiative transitions (in particular bb radiation) completely changes the picture for states that are close in energy but are strongly coupled through radiative transitions. Such states would be inaccurately modeled by being placed within the same energy group.

4. Computational Method

The self-consistent solution of the governing equations for the material gas (38) and the radiation field (40) is, in general, challenging, due to the non-local nature of radiation, which introduces a global coupling between the solution at all points. In mathematical terms, the inclusion of radiation transforms the non-radiating shock flow problem, which is an initial value problem, to a mixed initial-boundary value problem. In view of this, one has to resort to an iterative approach for numerical solutions. Various techniques have been proposed to solve radiation hydrodynamics problems (Mihalas & Mihalas 1984, Chapter 6). For the present study, the method of global iterations developed in the series of papers by Fadeyev & Gillet (1998, 2000, 2001) and Fadeyev et al. (2002) is employed. The former is essentially a lambda iteration method, where the flow-governing equations at iteration n are solved using the radiative rates from iteration $n-1$. To speed up the calculations, the solution is usually restarted from a previously computed one with a slightly different free-stream velocity. A similar approach has been developed independently by Panesi & Huo (2011) to study ionization phenomena in air for atmospheric entry flows. It is known that the lambda iteration may suffer from a slow convergence rate when accounting for thick continua (e.g., Lyman continuum) and lines (Hubeny & Mihalas 2014, Chapter 12). To overcome these deficiencies, methods such as the complete linearization by Auer & Mihalas (1969a, 1969b, 1969c) or the accelerated lambda iteration by Ribicky & Hummer (1991, 1992) have been proposed. These techniques outperform the conventional lambda iteration. However, they come at the price of a more complex and lengthy implementation (especially for the complete linearization). Since the purpose of this work is to reduce the complexity of the NLTE kinetic mechanism, it was decided that a simpler method such as the one by Fadeyev & Gillet (1998) be adopted. Moreover, as shown by the results in Section 5, the Balmer and Paschen lines are mostly in emission throughout the shock layer, which is a favorable condition when using a method resembling lambda iteration.

4.1. Spatial, Wavelength, and Angular Grids

For convenience, the shock, which is treated as a discontinuous surface, is placed at x = 0. The left and right boundaries are placed at $x=-{x}_{{\rm{L}}}$ and $x={x}_{{\rm{R}}}$, respectively. The lengths ${x}_{{\rm{L}}}$ and ${x}_{{\rm{R}}}$ are set to values of the order of 102–103 m in order to include the whole extent of the precursor and radiative relaxation regions, respectively. In order to properly resolve the smaller scales of the internal relaxation region (≃10−2–101 m), an exponential stretching is applied to reduce the grid size around the shock location.

The wavelength domain is discretized as suggested by Fadeyev & Gillet (1998, 2000). When accounting only for continuum radiation, the procedure goes as follows. After prescribing the minimum and maximum wavelengths (${\lambda }_{\min }$ and ${\lambda }_{\max }$, respectively), the interval $[{\lambda }_{\min },{\lambda }_{\max }]$ is divided into sub-intervals determined by the photo-ionization thresholds ${\lambda }_{+i}$. Each of these sub-intervals is then discretized using Gauss–Legendre quadrature points. This procedure is slightly modified to account for line radiation by adding additional sub-intervals for each atomic line. This is motivated by the rapid variation of emission and absorption coefficients (and source functions as well) over the lines. For this reason, the wavelength domain close to an atomic line l is discretized by adopting a Gauss–Legendre or uniform grid over the interval $[{\lambda }_{l}-\delta {\lambda }_{l},{\lambda }_{l}+\delta {\lambda }_{l}]$, where the width $\delta {\lambda }_{l}$ is set to $5$$10\,\mathring{\rm A} $. The remaining sub-intervals are discretized as done for continuum radiation.

The angular variable μ is also discretized by using Gauss–Legendre quadrature points.

4.2. Numerical Solution of the Governing Equations

The flow-governing equations (38) are solved using a space-marching approach in both the pre-shock and the post-shock regions. This requires the specification of initial conditions. For the pre-shock region, LTE conditions are assumed during the first iteration. However, the gas in the far precursor is not in LTE due to non-equilibrium excitation caused by absorption of resonant radiation in atomic line wings (Murty 1968; Foley & Clarke 1973). To take this into account, the occupation numbers of free electrons, protons, and hydrogen bound states are computed (starting from the second iteration) by solving the statistical equilibrium equations (Hubeny & Mihalas 2014, Chapter 9). These equations are obtained by setting to zero the convective term in the species continuity equations (i.e., ${\omega }_{s}=0$) and are solved iteratively by a Newton–Raphson procedure. Once the shock location is reached, the integration of Equation (38) is stopped and the Rankine–Hugoniot jump relations (Zel'dovich & Raizer 1967, Chapter 1) are applied to determine the kinematic and thermo-chemical state of the gas just behind the shock. This provides the initial solution for the post-shock region. The jump relations are solved under the assumption of frozen kinetics and by neglecting the effects of radiant energy fluxes (Marshak 1958). Free electrons are assumed isothermal within the shock. Alternatively, one could consider a slightly more accurate method by treating the compression of free electrons as isentropic (Zel'dovich & Raizer 1967, Chapter 7; Fadeyev & Gillet 1998). Preliminary calculations indicated, however, that this second approach does not lead to appreciable improvements compared to the first one (isothermal compression). In the present work, the flow-governing equations (38) are numerically integrated by using a fifth-order Backward Differentiation Formula method (Gear 1971, Chapter 7) implemented in the lsode library for stiff initial value problems (Radhakrishnan & Hindmarsh 1993). For convenience, the numerical integration is performed by using the mass fractions, velocity, and temperatures as solution variables (Panesi et al. 2009, 2011; Munafò et al. 2014, 2015). This is motivated by the particularly simple form assumed by the flow-governing equations (38) when rewritten in terms of these variables.

Radiative transfer is treated by using Feautrier's method (in the improved version proposed by Ribicky & Hummer 1991) because of its straightforward implementation for one-dimensional plane-parallel problems. The Feautrier form of the RTE (40) is discretized using a conventional second-order finite difference method under the assumption of no incoming radiation from both boundaries (Auer 1967; Hubeny & Mihalas 2014). For a non-scattering medium where Doppler shifts due to bulk motions are neglected, the above procedure leads to a set of uncoupled tridiagonal systems of equations (one for each discrete angle-wavelength point). These tridiagonal systems are solved using the elimination scheme proposed by Ribicky & Hummer (1991) for the sake of better numerical conditioning.

5. Results

The present section describes the results obtained by applying the StS and ME models to radiative shock waves propagating through a partially ionized atomic hydrogen plasma. The discussion is organized into two parts. The first, Section 5.1, illustrates the general features of the test case under investigation. In the second, Section 5.2, the StS predictions are systematically compared with those of the reduced-order MEU and MEL models to assess the accuracy. The comparison is performed on both gas and radiation quantities.

The free-stream temperature, ${T}_{\infty }$, and pressure, ${p}_{\infty }$, are set to $5000\,K$ and $5\,{Pa}$, respectively. The former corresponds to a number density, n, and mass density, ${\rho }_{\infty }$, of 7.24 × 1019 m−3 and 1.21 × 10−7 kg m−3. The free-stream velocity, ${u}_{\infty }$, is varied between a minimum of 40 km s−1 and a maximum of 70 km s−1 in steps of 5 km s−1. The adopted free-stream conditions are typical of atmospheres of pulsating stars (e.g., Cepheid variables) and are similar to the ones used by Fadeyev & Gillet (1998, 2000).

In all of the calculations, the precursor and radiative relaxation lengths are set to, respectively, 2000 and 500 m. The minimum and maximum wavelengths are taken at $600\,\mathring{\rm A} $ and $20000\,\mathring{\rm A} $, respectively. In this work, the number of atomic lines considered is nine, namely, the α, β, and γ lines of the Lyman, Balmer, and Paschen series.

Preliminary convergence studies have been performed on the spatial, wavelength, and angular grids. For the spatial grid, calculations indicated that using 1690 nodes, with a maximum grid size of 5 m at the boundaries and a minimum of 1 × 10−3 m close to the shock, lead to accurate predictions. For the wavelength domain, it has been observed that placing more than 16–32 Gauss–Legendre nodes in each photo-ionization sub-interval did not result in appreciable improvements when accounting only for continuum radiation. This outcome is not surprising and is due to the smoothness of the continuum emission and absorption coefficients as functions of wavelength/frequency. The inclusion of line radiation required the adoption of a higher number of nodes. In the present work, up to 300 wavelength nodes (distributed uniformly) have been used for each atomic line. Finally, for the angular grid, results have demonstrated that 32 Gauss–Legendre nodes were sufficient to resolve the directional dependence of the radiation field.

5.1. General Features of the StS Predictions

Figure 2 shows the evolution of the temperatures and electron mole fraction across the internal relaxation and the near-precursor regions for a shock propagating at 40 km s−1. The absorption of Lyman continuum radiation photo-ionizes the gas in the free stream and leads to an increase in the electron concentration, which is about 1% at the shock location. The spatial extent of the near-precursor region is two orders of magnitude larger than that of the internal relaxation region. The temperature evolution within the latter shows the typical structure observed in atomic plasmas (Zel'dovich & Raizer 1967, Chapter 7), where an initial sudden increase of the free-electron temperature is followed by a region where global ionization is indirectly controlled by elastic energy transfer in electron-heavy collisions. This may be explained as follows. Right behind the gasdynamic jump, the temperature of free electrons is of the order of 1–2 eV. The ionization of hydrogen in the internal relaxation region occurs via a two-step process: (i) excitation of high-lying electronic states and (ii) subsequent ionization from these states. Direct ionization from the ground state is also possible, although this may occur only in collisions with very energetic electrons (i.e., 10–20 eV). In both situations, free electrons lose on average 10 eV to produce a proton. Thus, the formation of a proton requires an energy equal to the thermal energy of 10 electrons. It is clear that if the free-electron energy were not replenished, their temperature would drop very rapidly. This will in turn hinder ionization as the rate of this process is proportional to the Boltzmann factor $\exp (-{E}_{+i}/{k}_{{\rm{B}}}{T}_{{\rm{e}}})$. The energy lost by free electrons in excitation/ionization collisions is replenished via elastic collisions with hot heavy particles. The inefficient energy transfer in these interactions (due to the large mass disparity) limits the amount of energy recovered by free electrons, and thus indirectly controls macroscopic ionization. Within the internal relaxation region, ionization through electron–atom collisions stimulates the further ionization of the gas, leading to avalanche (or cascade) processes (see Equation (11)). The electron avalanche causes an exponential increase in the free-electron concentration and is accompanied, in the final stages, by a rapid drop of the heavy-particle temperature (see Figure 2(b)). The avalanche terminates when the amount of protons produced is such that three-body recombination balances macroscopic ionization. After thermal equilibrium is reached, the plasma enters the radiative relaxation region where the degree of ionization decreases due to radiative recombination (see Figure 2(a)).

Figure 2.

Figure 2. (a) Electron mole fraction and (b) temperatures switching on/off the radiative transitions. In (a) the solid line is with radiation, the dashed line without radiation. In (b) the solid line is ${T}_{{\rm{h}}}$ with radiation, the dashed line is ${T}_{{\rm{e}}}$ with radiation, the dotted–dashed line is ${T}_{{\rm{h}}}$ without radiation, and the dotted line is ${T}_{{\rm{e}}}$ without radiation (StS model; ${u}_{\infty }=40\,\,\mathrm{km}\,{{\rm{s}}}^{-1}$).

Standard image High-resolution image

Figure 2 also reports the temperature and electron mole fraction when neglecting radiative processes. This is done to show the effects of radiation on the shock structure. Neglecting radiative transitions (in particular photo-ionization) leads to a larger internal relaxation region, due to the smaller concentration of free electrons at the shock location. In the present case, the length of the above zone is almost five times larger. Once ionization is completed, the gas properties (e.g., chemical composition, temperatures) maintain their post-shock LTE values.

To investigate the number of iterations needed to obtain a converged solution, the L2 norm of the relative error on the heavy-particle translational temperature has been monitored (see Figure 3). The relative error is defined as

Equation (50)

where the symbol Tnj denotes the temperature at the discrete node j and global iteration n, and N = 1690 stands for the number of spatial nodes. For both situations analyzed in Figure 3 (with and without bb radiation), a converged solution is essentially obtained after 40 lambda iterations. As expected, the rate of convergence is slower when accounting for line radiation. Similar trends to those shown in Figure 3 have been observed for the other cases investigated in the paper.

Figure 3.

Figure 3. Time history of the normalized L2 norm of the relative error (${\mathrm{log}}_{10}$ scale) on the heavy-particle translational temperature: unbroken line with bound–bound radiation, dashed line without bound–bound radiation (StS model; ${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1}$).

Standard image High-resolution image

Increasing the shock speed leads to more energetic photons from the radiative cooling region. This causes a larger degree of ionization in the precursor (see Figure 4(a)), which has, in turn, the effect of reducing the length of the internal relaxation region by more than one order of magnitude as revealed by the evolution of the heavy-particle translational temperature shown in Figure 4(b). To help in highlighting the above reduction in length, a logarithmic scale is used for the x-axis in Figure 4(b). For free-stream velocities larger than 60 km s−1, the hydrogen plasma in the precursor is fully ionized. Under these circumstances, when the gas enters the radiative cooling region, the electron mole fraction remains almost constant at the beginning, and then decreases due to radiative recombination. The extent of the region of (almost) constant degree of ionization increases with the shock speed and is reflected in the plateau observed for the heavy-particle temperature in Figure 4(b). At large shock speeds, the ionization avalanche is followed by a fast cooling zone, which is slowed by the onset of radiative recombination. This is demonstrated in Figure 5 which superimposes the electron mole fraction and temperature evolution at 40 and 70 km s−1.

Figure 4.

Figure 4. (a) Electron mole fraction and (b) heavy-particle temperature for different free-stream velocities: line with circle ${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1}$, line with square ${u}_{\infty }=45\,\mathrm{km}\,{{\rm{s}}}^{-1}$, line with triangle ${u}_{\infty }=50\,\mathrm{km}\,{{\rm{s}}}^{-1}$, line with diamond ${u}_{\infty }=55\,\mathrm{km}\,{{\rm{s}}}^{-1}$, dashed line ${u}_{\infty }=60\,\mathrm{km}\,{{\rm{s}}}^{-1}$, dotted–dashed line ${u}_{\infty }=65\,\mathrm{km}\,{{\rm{s}}}^{-1}$, and dotted line ${u}_{\infty }=70\,\mathrm{km}\,{{\rm{s}}}^{-1}$ (StS model).

Standard image High-resolution image
Figure 5.

Figure 5. Temperatures and electron mole fraction for (a) ${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1}$ and (b) ${u}_{\infty }=70\,\mathrm{km}\,{{\rm{s}}}^{-1}$: solid line ${T}_{{\rm{h}}}$, dashed line ${T}_{{\rm{e}}}$, dotted–dashed line ${X}_{{\rm{e}}}$ (StS model).

Standard image High-resolution image

The pronounced temperature drop at 70 km s−1 (Figure 5(b)) is due to the absence of neutral absorbers, which makes the plasma completely transparent to continuum radiation (see Equation (34)). The temperature decrease favors radiative recombination as the rate coefficient for this process is larger at low temperatures (see Equations (32) and (33)). Once the amount of ground-state hydrogen atoms is large enough, the radiation emitted in the Lyman continuum is partially re-absorbed, causing the temperature inflection point observed in Figure 5(b). At lower speeds (see Figure 5(a)), the above temperature drop is not observed, as the shock strength is not sufficient to fully ionize the incoming gas. The trends shown in Figures 4 and 5 are similar to those reported in the work by Fadeyev & Gillet (1998, 2000).

The evolution of the free-electron temperature in Figure 5 shows a non-monotonic behavior in the precursor. The initial rise is due to energy deposition by photo-ionization (see Equation (35)). During this stage, elastic collisions between charged particles are unable to bring heavy particles and free electrons into equilibrium due to the low degree of ionization. This trend continues until the further increase of ionization causes elastic energy transfer to dominate over photo-ionization. As a result, the free-electron temperature reaches a maximum and then relaxes toward the heavy-particle temperature. The observed precursor behavior of the free-electron temperature is consistent with the findings of Foley & Clarke (1973), who investigated radiative shock waves in helium and argon plasmas. The only significant difference from the above reference is that, in the present work, free electrons are essentially in thermal equilibrium with heavy particles at the shock location. This is most probably due to the more efficient energy transfer in electron–atom collisions caused by the lower atomic weight of hydrogen (which is, respectively, 4 and 40 times lighter than helium and argon). This fact also has a strong influence on the length of the internal relaxation region (where elastic collisions play a crucial role), as recognized in the early theoretical and experimental work by Belozerov & Measures (1969).

The precursor heating for the whole gas becomes substantial only at large shock speeds, as indicated by the rise of the heavy-particle temperature and pressure in Figures 6(a) and (b), and by the volumetric radiant loss term plotted in Figure 6(c). It is interesting to note that at large speeds (i.e., 70 km s−1), the free-electron temperature reaches a local minimum, and then rises again before reaching the shock. The observed behavior is due to the combined effect of the energy absorbed by radiation, which tends to increase the thermal energy of heavy particles, and elastic electron–proton collisions, which tend to keep the heavy-particle and free-electron temperatures together. The absorption of radiation in the precursor, in addition to altering the temperature and concentration profiles with respect to their free-stream values, also modifies the post-shock conditions due to the decrease and increase, respectively, of the Mach number and the pressure of the incoming gas. This is shown in detail in Table 1, which shows the pre- and post-shock conditions with and without radiation. At large speeds, the absorption of radiation reduces the post-shock temperature by more than 20%. As opposed to pressure and temperature, the mass density is quite insensitive to precursor radiation (see Table 1). This is a consequence of the (near) constancy of the flow velocity (not provided in Table 1), which can be explained by recalling that the absorbed radiation goes mainly into the thermal energy of the material gas. For a standing shock, global mass conservation requires that mass flux $\rho u$ must be constant, so that the density cannot vary appreciably when the flow velocity is changed by a small amount.

Figure 6.

Figure 6. Evolution of (a) the heavy-particle and free-electron temperatures, (b) the gas pressure, and (c) the volumetric radiative loss term in the precursor for different shock speeds. In (a), the solid line is ${T}_{{\rm{h}}}$ for ${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1}$, the dashed line is ${T}_{{\rm{e}}}$ for ${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1}$, the dotted–dashed line is ${T}_{{\rm{h}}}$ for ${u}_{\infty }=70\,\mathrm{km}\,{{\rm{s}}}^{-1}$, and the dotted line is ${T}_{{\rm{e}}}$ for ${u}_{\infty }=70\,\mathrm{km}\,{{\rm{s}}}^{-1}$. In (b) and (c), the solid line ${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1}$, and the dashed line ${u}_{\infty }=70\,\mathrm{km}\,{{\rm{s}}}^{-1}$ (StS model).

Standard image High-resolution image

Table 1.  Pressure, Temperature, and Density at the Pre- and Post-shock Locations Switching Radiation On/Off  (StS Model)

  Without Radiation With Radiation
  Post-shock Pre-shock Post-shock
${u}_{\infty }\,(\mathrm{km}\,{{\rm{s}}}^{-1})$ $p\,(\mathrm{Pa})$ $T\,({\rm{K}})$ $\rho \,(\mathrm{kg}\,{{\rm{m}}}^{-3})$ $p\,(\mathrm{Pa})$ $T\,({\rm{K}})$ $\rho \,(\mathrm{kg}\,{{\rm{m}}}^{-3})$ $p\,(\mathrm{Pa})$ $T\,({\rm{K}})$ $\rho \,(\mathrm{kg}\,{{\rm{m}}}^{-3})$
40 144.086 40683.75 4.29 × 10−7 5.087 5034.509 1.21 × 10−7 142.79 40331.96 4.28 × 10−7
45 182.69 50347.42 4.40 × 10−7 5.22 5082.56 1.21 × 10−7 178.46 49209.53 4.38 × 10−7
50 225.83 6114.46 4.48 × 10−7 5.46 5171.12 1.21 × 10−7 214.91 58231.51 4.45 × 10−7
55 273.52 73076.09 4.54 × 10−7 5.89 5320.73 1.21 × 10−7 249.15 66616.85 4.45 × 10−7
60 325.74 86142.43 4.58 × 10−7 6.56 5548.98 1.22 × 10−7 278.84 73744.47 4.52 × 10−7
65 382.51 100343.87 4.62 × 10−7 7.50 5892.28 1.22 × 10−7 306.054 80208.88 4.54 × 10−7
70 443.82 115680.72 4.65 × 10−7 8.61 6322.32 1.22 × 10−7 333.62 86745.66 4.54 × 10−7

Note. The pre-shock pressure, temperature, and density are not provided in the case without radiation as these quantities are given by their free-stream LTE values (${p}_{\infty }=5\,{Pa}$, ${T}_{\infty }=5000\,K$, ${\rho }_{\infty }=1.21e-7\,\mathrm{kg}\,{{\rm{m}}}^{-3}$).

Download table as:  ASCIITypeset image

Before moving to the comparison with the reduced-order ME models discussed in Section 5.2, it is worth briefly analyzing the effect of including/excluding bb transitions from the calculations. Figure 7 shows a sample of this investigation for the temperatures and electron mole fraction profiles obtained for a shock propagating at 40 km s−1. The inclusion of line radiation has the effect of shrinking the internal relaxation region and speeding up radiative recombination. The observed behavior originates from emission in the Balmer and Paschen lines, which are optically thin in the post-shock region, as shown in Figure 8, which shows the average monochromatic intensity at $x=2\,{\rm{m}}$ and $x=10\,{\rm{m}}$. The effect of an optically thin line (in particular the ${H}_{\alpha }$ line) is to provide an efficient channel to depopulate the higher bound electronic states. This is accompanied by a loss of energy which, in turn, induces the faster temperature drop observed in Figure 7(b). At the same time, in the radiative cooling region, the depopulation of high-lying states due to line emission indirectly favors radiative recombination because it makes the quantity $({n}_{i}\,{k}_{i}^{{\rm{PI}}}-{n}_{{\rm{e}}}\,{n}_{+}\,{k}_{i}^{{\rm{RR}}})$ more negative for these states (see Equation (29)). For large shock speeds (e.g., 60 km s−1), the effects of bbtransitions are of less importance, due to the larger amount of photo-ionization in the precursor. This general trend has been observed in all cases investigated in this work.

Figure 7.

Figure 7. (a) Electron mole fraction and (b) temperatures switching bound–bound radiation on/off. In (a), the solid line is with bound–bound radiation, the dashed line is without bound–bound radiation. In (b), the solid line is ${T}_{{\rm{h}}}$ with bound–bound radiation, the dashed line is ${T}_{{\rm{e}}}$ with bound–bound radiation, the dotted–dashed line is ${T}_{{\rm{h}}}$ without bound–bound radiation, and the dotted line is ${T}_{{\rm{e}}}$ without bound–bound radiation (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ StS model).

Standard image High-resolution image
Figure 8.

Figure 8. Average monochromatic intensity at the locations (a) $x=2\,{\rm{m}}$ (internal relaxation region) and (b) $x=10\,{\rm{m}}$ (radiative cooling region) behind the shock switching on/off the bound–bound radiation: solid line is with bound–bound radiation, dashed line is without bound–bound radiation. The dotted–dashed line represents the Planck function at the free-electron temperature, ${B}_{\lambda }({T}_{{\rm{e}}})$, when including bound–bound radiation (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ StS model).

Standard image High-resolution image

5.2. Comparison Between the StS and the ME Models

After analyzing the general features of the radiative shock waves considered in this work, predictions obtained by means of the ME models were systematically compared with the StS results. The comparison was first performed by excluding line radiation (Section 5.3), which was then added back into the final runs (Section 5.4).

Before discussing the results, it is worth saying a few words on how the grouping was done in practice. As anticipated at the end of Section 3.2, one should avoid lumping together states that are coupled by strong radiative transitions. This is particularly true for optically thin lines (e.g., ${\rm{H}}\alpha $, ${\rm{H}}\beta $). For this reason, the first and second energy groups (when two or more are used) were assigned to the n = 1 and n = 2 states, respectively, where n stands for the principal quantum number (Pauling & Wilson 1935, Chapter 5; Hubeny & Mihalas 2014, Chapter 7). No internal temperatures are used for these groups in the case of the MEL model, allowing for a further reduction of the number of unknowns. The higher states ($n\geqslant 3$) are lumped based on Equation (46) (or Equation (47) for the MEU model). In what follows, the notations MEU(k) and MEL(k) are used to indicate the reduced-order model used and the number of groups, k.

5.3. Continuum Radiation Only

Figure 9 shows the temperature evolution for a shock propagating at 40 km s−1 when using the MEL(3) model. The (only) internal temperature, indicated by the dotted–dashed line, refers to the states with $n\geqslant 3$. In the precursor, the high-lying levels are not in thermal equilibrium with free electrons since ${T}_{{\rm{i}}}\ne {T}_{{\rm{e}}}$. This behavior is due to the concurrent action of photo-ionization from high-lying states and de-excitation collisions (Foley & Clarke 1973). The internal temperature replicates the observed non-monotonic trend of that of the free electrons. When the incoming gas reaches the shock, the high-lying states are almost in thermal equilibrium with free electrons.

Figure 9.

Figure 9. Temperature evolution across (a) the precursor and (b) the internal relaxation region for the MEL(3) model: solid line ${T}_{{\rm{h}}}$, dashed line ${T}_{{\rm{e}}}$, dotted–dashed line ${T}_{{\rm{i}}}$ (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ no bound–bound radiation).

Standard image High-resolution image

The results of Figure 9 were compared with the StS predictions in Figure 10 for the electron mole fraction and the heavy-particle and free-electron temperatures. The MEU(3) solution is also reported. The MEL(3) model is in excellent agreement with the StS results. The two solutions are essentially indistinguishable. These results indicate that using three energy groups plus one internal temperature is sufficient to achieve an accurate prediction of radiative shock waves in hydrogen plasmas. In practical terms, this means that the number of unknowns is reduced by two orders of magnitude, which makes the MEL model very attractive for Computational Fluid Dynamics (CFDs) applications. On the other hand, the MEU(3) predictions are in strong disagreement with the StS results.

Figure 10.

Figure 10. Comparison between the StS, MEU(3), and MEL(3) model predictions for (a) the electron mole fraction and (b) the heavy-particle and free-electron temperatures. In (a), the solid line is the StS model, the dashed line is the MEU(3) model, the dotted–dashed line is the MEL(3) model. In (b) the solid line is the ${T}_{{\rm{h}}}$ of the StS model, the dashed line is the ${T}_{{\rm{h}}}$ of the MEU(3) model, the dotted–dashed line is the ${T}_{{\rm{h}}}$ of the MEL(3) model, the line with the circles is the ${T}_{{\rm{e}}}$ of the StS model, the dashed line with squares is the ${T}_{{\rm{e}}}$ of the MEU(3) model, and the dotted–dashed line with the triangles is the ${T}_{{\rm{e}}}$ of the MEL(3) model (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ no bound–bound radiation).

Standard image High-resolution image

To gain more insight from the comparison of Figure 10, the normalized populations of the internal energy states, ni/gi, have been extracted at the locations $x=-10,2$ and $20\,{\rm{m}}$ (see Figures 1113). The first location (Figure 11) refers to the near precursor, the second (Figure 12) to the internal relaxation region, and the third (Figure 13) to the radiative cooling region. In the same figures, the corresponding average monochromatic intensity is also provided. For the MEL and MEU models, the population distributions have been reconstructed based on the group number densities and internal temperatures using Equations (46) and (47), respectively. In the present work, the normalized populations, ni/gi, have been preferred over the conventional population ratios (Hubeny & Mihalas 2014, Chapter 9), ${n}_{i}/{n}_{i}^{* }$ (where the symbol ${n}_{i}^{* }$ denotes the Saha equilibrium (10) occupation numbers), for the following reason. The population ratios give information only about the deviation from the local Saha distribution. On the other hand, the inspection of the ratio ni/gi on a semi-log plot (as done in Figures 1113) allows for a deeper understanding of NLTE effects since it shows by how much the energy population deviates from the local Boltzmann distribution. This is immediately realized by taking the log of Equation (46):

Equation (51)

$k\in {{ \mathcal I }}_{k}$, $k\in { \mathcal K }$. Equation (51) shows that in the plot $\mathrm{ln}({n}_{i}/{g}_{i})$ versus Ei, the equilibrium population follows a straight line whose slope is indirectly proportional to the local temperature. This provides an effective way to quantify the departure from the local equilibrium distribution. This fact is of particular importance for high-lying states, which are the ones where most of the ionization occurs. The results in Figures 1113 show that the superior description of the MEL model lies in its ability to (almost) replicate the StS behavior of the high-lying states. This fact is of great importance, especially for the ionization within the internal relaxation region. The MEU model gives a poor description of the population dynamics due to the assumption of infinite internal temperatures. As already stated in Section 3.1, this hypothesis does not allow the retrieval of the equilibrium (i.e., Boltzmann distribution) and has a strong impact on the temperature evolution in the radiative relaxation region.

Figure 11.

Figure 11. Comparison between the StS, MEU(3), and MEL(3) model predictions for (a) the population of H bound levels and (b) the average monochromatic intensity at the location $x=-10\,{\rm{m}}$ ahead of the shock (near-precursor region): line with circles, StS model; line with squares, MEU(3) model; line with triangles, MEL(3) model. The dotted–dashed line in (b) represents the Planck function at the free-electron temperature, ${B}_{\lambda }({T}_{{\rm{e}}})$, for the StS model (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ no bound–bound radiation).

Standard image High-resolution image
Figure 12.

Figure 12. Comparison between the StS, MEU(3), and MEL(3) model predictions for (a) the population of H bound levels and (b) the average monochromatic intensity at the location $x=2\,{\rm{m}}$ behind the shock (internal relaxation region): line with circles, StS model; line with triangles, MEU(3) model; line with squares, MEL(3) model. The dotted–dashed line in (b) represents the Planck function at the free-electron temperature, ${B}_{\lambda }({T}_{{\rm{e}}})$, for the StS model (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ no bound–bound radiation).

Standard image High-resolution image
Figure 13.

Figure 13. Comparison between the StS, MEU(3), and MEL(3) model predictions for (a) the population of H bound levels and (b) the average monochromatic intensity at the location $x=20\,{\rm{m}}$ behind the shock (radiative cooling region): line with circles, StS model; line with squares, MEU(3) model; line with triangles, MEL(3) model. The dotted–dashed line in (b) represents the Planck function at the free-electron temperature, ${B}_{\lambda }({T}_{{\rm{e}}})$, for the StS model (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ no bound–bound radiation).

Standard image High-resolution image

In this zone, where the degree of ionization decreases due to radiative recombination, the collisional rate among the gas particles is large enough to maintain equilibrium at the local free-electron temperature (see Figure 13(a)). The same holds true for the radiation field only in the Lyman continuum, as shown by the observed departure from the local Planck function, ${B}_{\lambda }({T}_{{\rm{e}}})$, in the Balmer and Paschen continuum (see Figure 13(b)). The use of a uniform grouping prevents obtaining a Boltzmann distribution and leads to a higher population of the high-lying states and an underpredicted population for the n = 2 and n = 3 states (see Figure 13(a)). This produces lower temperatures compared to the MEL(3) and StS solutions. The observed difference is not negligible and is also enhanced by the large statistical weight of the high-lying states.9 The poor description of the level dynamics by the MEU(3) model also has an adverse effect on the predicted radiative signature of the plasma. This is demonstrated by the comparison in terms of the average monochromatic intensity given in Figures 11(b), 12(b), and 13(b). At all locations, the difference is higher at the Lyman and Balmer edges. The above findings on the MEU(3) and MEL(3) models have also been obtained when analyzing cases at larger shock speeds. For the sake of brevity, these results have not been added to the manuscript.

Before including the effects of line radiation on the reduced-order models (see Section 5.4), it was decided that the sensitivity of the solution to the number of groups will be investigated. The study included both the MEU and MEL models.

Figure 14 shows the free-electron mole fraction and all temperatures for the MEL(1), MEL(2), and MEL(3) models. When all the hydrogen bound states are lumped into one group (i.e., MEL(1) model), it is not possible to account for the under-population experienced by the high-lying states in the precursor and within the internal relaxation region. As a matter of fact, the MEL(1) model forces the high-lying states to be in thermal equilibrium with free electrons, resulting in a larger ionization rate (see Figures 14(a) and (b)). It is worth mentioning that, despite the use of one group, the MEL(1) model is still a NLTE model since (i) the hydrogen number density is not prescribed via Saha's equation (10), and (ii) the group internal temperature ${T}_{{\rm{i}}}$ is allowed to depart from the translational temperature of heavy particles and free electrons. The inclusion of one additional group (i.e., MEL(2) model) for the states with $n\geqslant 2$ greatly improves the solution and leads essentially to the same result obtained with the MEL(3) model. The MEL(3) model, however, gives a superior description of the population distribution as demonstrated by the comparison between the three models in Figure 15. It is worth noting that the lower accuracy for the population of the n = 2 and n = 3 states for the MEL(2) model plays a negligible role on gas quantities such as chemical composition and temperatures. This is no longer true when line radiation is taken into account (see Section 5.4). One may conclude that using only two energy groups is already enough to obtain an accurate prediction when neglecting line radiation.

Figure 14.

Figure 14. Comparison between the StS and the MEL(1), (a) and (b), MEL(2), (c) and (d), and the MEL(3), (e) and (f), model predictions for the electron mole fraction, and the heavy-particle and free-electron temperatures. In (a), (c), and (e), the solid line is the StS model, and the dashed line is the MEL(i) model. In (b), (d), and (f), the solid line is the ${T}_{{\rm{h}}}$ of the StS model, the dashed line is the ${T}_{{\rm{e}}}$ of the StS model, the line with circles is the ${T}_{{\rm{h}}}$ of the MEL(i) model, and the line with squares is the ${T}_{{\rm{e}}}$ of the MEL(i) model (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ no bound–bound radiation).

Standard image High-resolution image
Figure 15.

Figure 15. Comparison between the StS and the MEL(i) model predictions for the population of H bound levels at the location $x=2\,{\rm{m}}$ behind the shock (internal relaxation region): line with circles, StS model; line with squares, MEL(1) model; line with triangles, MEL(2) model; and line with stars, MEL(3) model (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ no bound–bound radiation).

Standard image High-resolution image

Figure 16 illustrates the results of the sensitivity study for the MEU model. In analogy with the MEL model, increasing the number of groups improves the solution. However, in the present case, this is achieved with a larger number of groups (11), which makes the MEU model less attractive for CFD applications. The use of 11 groups allows the accurate prediction of both temperature and species concentrations in the precursor and internal relaxation region. However, some discrepancies with the StS results still appear in the radiative cooling region. As demonstrated earlier, these are caused by the intrinsic defect of the MEU model, namely, the impossibility of retrieving a Boltzmann distribution for the bound states. In view of these results, the MEU model has not been considered in the calculations performed, including the effects of line radiation as discussed in Section 5.4.

Figure 16.

Figure 16. Comparison between the StS and the MEU(i) model predictions for (a) the electron mole fraction and (b) the population of H bound levels at the location $x=2\,{\rm{m}}$ behind the shock: line with circles, StS model; line with squares, MEU(3) model; line with triangles, MEU(7) model; and line with stars, MEU(11) model (${u}_{\infty }=40\,\mathrm{km}\,{{\rm{s}}}^{-1};$ no bound–bound radiation).

Standard image High-resolution image

5.4. Line and Continuum Radiation

The sensitivity analysis at the end of Section 5.3 has shown that two energy groups are sufficient to obtain an accurate solution when accounting only for continuum radiation. To test whether this finding is retrieved when atomic lines are added, the StS and MEL(2) solutions have been compared in Figure 17 for the electron mole fraction and heavy-particle and free-electron temperatures. In the same image, the MEL(3) solution is also shown. Overall, the MEL(2) solution now shows a sensible departure from the StS prediction, the difference being larger in the internal relaxation region (see Figures 17(a) and (b)). In the precursor, the MEL(2) model slightly overestimates photo-ionization, while in the radiative cooling region the disagreement with the StS solution is barely noticeable. The MEL(3) model is again in excellent agreement with the StS prediction (see Figures 17(c) and (d)). The performance degradation of the MEL(2) model is not surprising and is the result of grouping together the n = 2 and n = 3 states. This annihilates the effect of radiative decay through the optically thin ${\rm{H}}\alpha $ line, which, as shown in Figure 7, has a non-negligible impact on the internal relaxation and radiative cooling regions. On the other hand, in the MEL(3) model, the n = 2 and n = 3 states are grouped separately, thus allowing for a more accurate prediction of bb radiative losses. The above qualitative arguments are confirmed when comparing the StS and MEL(2-3) population distributions as done in Figure 18. The above picture refers to the location $x=2\,{\rm{m}}$ behind the shock (internal relaxation region), and for the sake of completeness, also shows results obtained when neglecting line radiation. It is readily seen that, when the n = 2 and n = 3 states are lumped together, the accuracy of the reconstructed distribution strongly degrades. In particular, the population of the above states (and also those close to the ionization limit) is underpredicted by at least one order of magnitude. It is worth noting that the observed departure between the StS and MEL(3) predictions for the population of the n = 3 and n = 4 states (see Figure 18(b)) has essentially no effect on the solution accuracy. This can be explained by recalling that in general, the Einstein coefficient for the ${\rm{H}}\alpha $ line is one order of magnitude larger than of the ${\rm{H}}\beta $ and ${\rm{H}}\gamma $ lines.

Figure 17.

Figure 17. Comparison between the StS and the MEL(2), (a) and (b), and the MEL(3), (c) and (d), model predictions for the electron mole fraction, and the heavy-particle and free-electron temperatures. In (a) and (c), the solid line is the StS model, and the dashed line is the MEL(i) model. In (b) and (d), the solid line is the ${T}_{{\rm{h}}}$ of the StS model, the dashed line is the ${T}_{{\rm{e}}}$ of the StS model, the line with circles is the ${T}_{{\rm{h}}}$ of the MEL(i) model, and the line with squares is the ${T}_{{\rm{e}}}$ of the MEL(i) model (${u}_{\infty }=40\,\,\mathrm{km}\,{{\rm{s}}}^{-1}$).

Standard image High-resolution image
Figure 18.

Figure 18. Comparison between the StS, MEL(2), and MEL(3) model predictions for the population of H bound levels at the location $x=2\,{\rm{m}}$ behind the shock (internal relaxation region): line with circles, StS model; line with squares, MEL(2) model; and line with triangles, MEL(3) model. In (a) without bound–bound radiation, in (b) with bound–bound radiation (${u}_{\infty }=40\,\,\mathrm{km}\,{{\rm{s}}}^{-1}$).

Standard image High-resolution image

Figure 19 compares the average monochromatic intensity around the ${\rm{H}}\alpha $, ${\rm{H}}\beta $, and ${\rm{H}}\gamma $ lines predicted by the StS and MEL(2-3) models at the location $x=10\,{\rm{m}}$ (radiative cooling region). Despite the good agreement observed for gas quantities (such as the electron mole fraction shown in Figure 17), the MEL(2) model systematically overestimates the peak value of the intensity of all lines. The MEL(3) model is, on the other hand, in excellent agreement with the StS results.

Figure 19.

Figure 19. Comparison between the StS, MEL(2), and MEU(3) model predictions for the average monochromatic intensity near the (a) ${\rm{H}}\alpha $, (b) ${\rm{H}}\beta $, and (c) ${\rm{H}}\gamma $ atomic lines at the location $x=10\,{\rm{m}}$ behind the shock (radiative cooling region): solid line, StS model; dashed line, MEL(2) model; dotted–dashed line, MEL(3) model (${u}_{\infty }=40\,\,\mathrm{km}\,{{\rm{s}}}^{-1}$).

Standard image High-resolution image

The accuracy of the MEL(3) model has been further confirmed by repeating the calculations treated in this section at larger shock speeds. This is demonstrated in Figure 20, which compares the StS and MEL(3) electron mole fraction profiles for increasing shock speeds. In analogy with what is observed for the 40 km s−1 case, the two solutions overlap.

Figure 20.

Figure 20. Comparison between the StS and the MEL(3) model predictions for the electron mole fraction for (a) ${u}_{\infty }=40\,\,\mathrm{km}\,{{\rm{s}}}^{-1}$, (b) ${u}_{\infty }=45\,\,\mathrm{km}\,{{\rm{s}}}^{-1}$, (c) ${u}_{\infty }=50\,\,\mathrm{km}\,{{\rm{s}}}^{-1}$, and (d) ${u}_{\infty }=55\,\,\mathrm{km}\,{{\rm{s}}}^{-1}$: solid line, StS model; line with circles, MEL(3) model.

Standard image High-resolution image

6. Conclusions

A StS kinetic model for NLTE astrophysical hydrogen plasmas has been constructed by collecting the most up-to-date ab initio and/or experimental data. Based on the ME principle, the complexity of the StS kinetic model has been reduced by lumping the bound energy states of hydrogen into energy groups. Two different grouping strategies have been considered in this work: uniform (MEU model) and Maxwell–Boltzmann (MEL model). The reduced set of governing equations for the material gas have been obtained based on a moment method. The accuracy of the MEU/MEL models has been tested by means of a systematic comparison with the StS predictions. Applications considered the flow across radiative shock waves for conditions typical of pulsating stars. The results have shown that, with the use of only two to three energy groups, the MEL model is in excellent agreement with the StS predictions. To be more specific, two energy groups are already enough to achieve an accurate solution when neglecting line radiation. The inclusion of atomic lines requires the adoption of one additional group to account for the optically thin losses in the Balmer and Paschen lines. This makes the MEL very attractive for potential CFD applications. The MEU model, on the other hand, is less accurate than the MEL model and requires the adoption of a larger number of groups to achieve a fair agreement with the StS model. The persistent disagreement (even with a large number of groups) is due to the assumption of uniform distribution, which prevents retrieving a Boltzmann distribution in the radiative cooling region.

The research of A.M. was supported by the NASA Award grant NNX 14AN44G. The research of M.P. was supported by the NASA grant NNX15AQ57A. The views and conclusions contained herein are those of the authors and should not be interpreted as necessarily representing the official policies or endorsements, either expressed or implied, of NASA. The authors gratefully acknowledge Dr. A Wray and Dr. D Hathaway at NASA Ames Research Center for the useful scientific discussions and suggestions while preparing the manuscript.

Appendix A: Hydrogen Data

This Appendix provides the energy levels of the model hydrogen atom used in the present work.

Table 2.  Energy Level Data for the Model Hydrogen Atom Taken from the NIST Atomic Database (Kramida 2010; Kramida et al. 2014)

# Conf. Term J g $E\,(\mathrm{eV})$ # Conf. Term J g $E\,(\mathrm{eV})$
1 1s ${}^{2}{\rm{S}}$ 1/2 2 0 48 3/2 4 13.430553985
2 2p ${}^{2}{\rm{P}}^\circ $ 1/2 2 10.19880606470 49 9s ${}^{2}{\rm{S}}$ 1/2 2 13.430553535
3 3/2 4 10.19885142904 50 9d ${}^{2}{\rm{D}}$ 3/2 4 13.4305539831
4 2s ${}^{2}{\rm{S}}$ 1/2 2 10.19881043960433 51 5/2 6 13.4305541491
5 3p ${}^{2}{\rm{P}}^\circ $ 1/2 2 12.0874935577 52 10p ${}^{2}{\rm{P}}^\circ $ 1/2 2 13.462450945
6 3/2 4 12.0875069990 53 3/2 4 13.462451308
7 3s ${}^{2}{\rm{S}}$ 1/2 2 12.0874948597 54 10s ${}^{2}{\rm{S}}$ 1/2 2 13.462450981
8 3d ${}^{2}{\rm{D}}$ 3/2 4 12.0875069769 55 10d ${}^{2}{\rm{D}}$ 3/2 4 13.46245130616
9 5/2 6 12.0875114568 56 5/2 6 13.46245142716
10 4p ${}^{2}{\rm{P}}^\circ $ 1/2 2 12.74853233939 57 11p ${}^{2}{\rm{P}}^\circ $ 1/2 2 13.486051441
11 3/2 4 12.74853800981 58 3/2 4 13.486051714
12 4s ${}^{2}{\rm{S}}$ 1/2 2 12.74853288970 59 11s ${}^{2}{\rm{S}}$ 1/2 2 13.486051468
13 4d ${}^{2}{\rm{D}}$ 3/2 4 12.7485380014 60 11d ${}^{2}{\rm{D}}$ 3/2 4 13.4860517119
14 5/2 6 12.74853989060 61 5/2 6 13.4860518029
15 4f ${}^{2}{\rm{F}}^\circ $ 5/2 6 12.748539891 62 12p ${}^{2}{\rm{P}}^\circ $ 1/2 2 13.504001544
16 7/2 8 12.7485408334 63 3/2 4 13.504001754
17 5p ${}^{2}{\rm{P}}^\circ $ 1/2 2 13.05449807364 64 12s ${}^{2}{\rm{S}}$ 1/2 2 13.504001564
18 3/2 4 13.054500977 65 12d ${}^{2}{\rm{D}}$ 3/2 4 13.50400175254
19 5s ${}^{2}{\rm{S}}$ 1/2 2 13.054498354 66 5/2 6 13.504001822555
20 5d ${}^{2}{\rm{D}}$ 3/2 4 13.0545009646 67 13 338 13.51797117
21 5/2 6 13.0545019320 68 14 392 13.52905540
22 5f ${}^{2}{\rm{F}}^\circ $ 5/2 6 13.054501936840 69 15 450 13.53799760
23 7/2 8 13.054502416 70 16 512 13.54531613
24 5g ${}^{2}{\rm{G}}$ 7/2 8 13.054502419807 71 17 578 13.55138155
25 9/2 10 13.054502710137 72 18 648 13.55646443
26 6p ${}^{2}{\rm{P}}^\circ $ 1/2 2 13.22070135109 73 19 722 13.56076607
27 3/2 4 13.22070303125 74 20 800 13.56443874
28 6s ${}^{2}{\rm{S}}$ 1/2 2 13.22070151443 75 21 882 13.56759935
29 6d ${}^{2}{\rm{D}}$ 3/2 4 13.22070302852 76 22 968 13.57033883
30 5/2 6 13.22070358845 77 23 1058 13.57272882
31 6f ${}^{2}{\rm{F}}^\circ $ 5/2 6 13.220703588190 78 24 1152 13.57482634
32 7/2 8 13.220703868212 79 25 1250 13.57667721
33 6g ${}^{2}{\rm{G}}$ 7/2 8 13.220703867684 80 26 1352 13.57831866
34 9/2 10 13.220704035699 81 27 1458 13.57978113
35 6h ${}^{2}{\rm{H}}^\circ $ 9/2 10 13.220704035368 82 28 1568 13.58108972
36 11/2 12 13.220704147381 83 29 1682 13.58226529
37 7p ${}^{2}{\rm{P}}^\circ $ 1/2 2 13.320916535 84 30 1800 13.58332527
38 3/2 4 13.320917593 85 31 1922 13.58428434
39 7s ${}^{2}{\rm{S}}$ 1/2 2 13.320916638 86 32 2048 13.58515490
40 7d ${}^{2}{\rm{D}}$ 3/2 4 13.3209175916 87 33 2178 13.58594754
41 5/2 6 13.3209179442 88 34 2312 13.58667126
42 8p ${}^{2}{\rm{P}}^\circ $ 1/2 2 13.38595996642 89 35 2450 13.58733384
43 3/2 4 13.38596067523 90 36 2592 13.58794197
44 8s ${}^{2}{\rm{S}}$ 1/2 2 13.38596003538 91 37 2738 13.58850147
45 8d ${}^{2}{\rm{D}}$ 3/2 4 13.38596067408 92 38 2888 13.58901738
46 5/2 6 13.385960910363 93 39 3042 13.58949412
47 9p ${}^{2}{\rm{P}}^\circ $ 1/2 2 13.430553487 94 40 3200 13.58993555

Note. The database includes information on all substates of hydrogen only for $n\leqslant 5$. For the states with $6\leqslant n\leqslant 12$, only those with orbital quantum number $l=0,1,2$ (s, p and d, respectively) are provided. For higher states, no substate information is given, and their statistical weight is evaluated as $g=2{n}^{2}$.

Download table as:  ASCIITypeset image

Appendix B: Thermodynamics, Kinetics, and Radiation of the Reduced-order Models

This appendix provides the constitutive thermodynamic and kinetic relations for the reduced-order NLTE models developed in Section 3. These relations are obtained by substituting the group distributions (46)/(47) in the expressions for thermodynamic properties, source terms, and emission/absorption coefficients for the StS model. Only the MEL model is considered. The corresponding formulas for the MEU model can be readily obtained taking the limit of infinite internal temperatures (as already explained in Section 3.1). For the sake of a more concise notation, the following group averaging operators are introduced:

Equation (52)

$k,l\in { \mathcal K }$, where the symbols ${\alpha }_{i}$ and ${\beta }_{{ij}}$ stand for the quantities being averaged. To avoid ambiguity in the evaluation of the averages $ \langle\langle {\beta }_{{ij}}{ \rangle\rangle }_{}$, the dummy index of the internal sum is set equal to the first lower script (i.e., i in this case).

B.1. Thermodynamics

The gas pressure is always given by Dalton's law, $p={n}_{{\rm{e}}}{k}_{{\rm{B}}}{T}_{{\rm{e}}}+{n}_{{\rm{h}}}{k}_{{\rm{B}}}{T}_{{\rm{h}}}$, where the heavy-particle number density is obtained by summing the group and proton contributions, ${n}_{{\rm{h}}}={\sum }_{k\in { \mathcal K }}{\tilde{n}}_{k}+{n}_{+}$. In view of Equation (46), the heavy-particle thermal energy density (8) becomes

Equation (53)

B.2. Collisional Kinetics

The group mass production terms due to collisional excitation and ionization are obtained via the substitution of Equation (46) into Equation (13) and then summing the result obtained over all of the levels within each group (i.e., ${\tilde{\omega }}_{k}={\sum }_{i\in {{ \mathcal I }}_{k}}{\omega }_{i}$). After some algebraic manipulation, one obtains

Equation (54)

Equation (55)

$k\in { \mathcal K }$, where the group endothermic rate coefficients for excitation and ionization, and those for the related exothermic de-excitation and three-body recombination processes, are

Equation (56)

Equation (57)

$k,l\in { \mathcal K }$, where $k\lt l$ in Equation (57). It is worth noting that the group rate coefficients for ionization and three-body recombination (56) are related by a detailed balance relation among the groups. The same holds true for excitation/de-excitation only in the case of thermal equilibrium with free electrons (i.e., ${T}_{k}={T}_{{\rm{e}}},\forall \,k\in { \mathcal K }$).

By applying the same procedure as above, one arrives at the following relations for the group energy transfer source terms due to ionization and excitation:

Equation (58)

$k\in { \mathcal K }$, where the related transfer rates due to ionization, recombination, and excitation/de-excitation are

Equation (59)

Equation (60)

$k,l\in { \mathcal K }$. The use of the Boltzmann grouping relation (46) in Equation (18) allows the ionization and excitation energy transfer terms for the free-electron gas to be rewritten as

Equation (61)

where ${\omega }_{+}^{{\rm{I}}}={\omega }_{+}^{\mathrm{col}}$.

B.3. Radiative kinetics

bb transitions—The substitution of Equation (46) into Equation (22) allows the monochromatic emission and absorption coefficients due to bb radiation to be rewritten as

Equation (62)

where the group wavelength and temperature-dependent Einstein coefficients are defined as

Equation (63)

$k,l\in { \mathcal K }$, where the obvious symmetry relations satisfied by the line emission/absorption profile and wavelength (i.e., ${\phi }_{\lambda }^{{ij}}={\phi }_{\lambda }^{{ji}}$ and ${\lambda }_{{ij}}={\lambda }_{{ji}}$, respectively) are used. The application of the above procedure to Equations (24) and (26) gives, respectively, the group mass production rates and the volumetric rate of energy loss of matter energy due to bb radiation:

Equation (64)

Equation (65)

where the group mass production and energy transfer Einstein coefficients are given by the following Boltzmann averages:

Equation (66)

Equation (67)

Equation (68)

$k,l\in { \mathcal K }$, where $k\leqslant l$.

bf/fb transitions—The substitution of Equation (46) into Equations (27) and (28) allows the monochromatic emission and absorption coefficients due to bf and fb radiation to be rewritten as

Equation (69)

Equation (70)

where the quantity ${\tilde{\sigma }}_{k}^{{\rm{PI}}}$ stands for the temperature-dependent group photo-ionization cross-section. The former is computed via the Boltzmann average ${\tilde{\sigma }}_{k}^{{\rm{PI}}}=\langle {\sigma }_{i}^{{\rm{PI}}}{\rangle }_{k}$. The notation ${\tilde{\sigma }}_{k}^{{\rm{PI}}}(\lambda ,{T}_{{\rm{e}}})$ indicates that the Boltzmann average has to be evaluated at the free-electron temperature.

The mass production terms are

Equation (71)

Equation (72)

$k\in { \mathcal K }$, where the group rate coefficients for photo-ionization and radiative recombination read ${\tilde{k}}_{k}^{{\rm{PI}}}=\langle {k}_{i}^{{\rm{PI}}}{\rangle }_{k}$ and ${\tilde{k}}_{k}^{{\rm{RR}}}={\sum }_{i\in {{ \mathcal I }}_{k}}{k}_{i}^{{\rm{RR}}}$, respectively. By setting the elementary photo-ionization cross-sections to zero for wavelengths above the photo-ionization limit (i.e., ${\sigma }_{i}^{{\rm{PI}}}=0$ for $\lambda \gt {\lambda }_{+i}$), it is possible to extend the wavelength integrals to infinity and exchange the order between summation and integration when computing the rate coefficients ${\tilde{k}}_{k}^{{\rm{PI}}}$ and ${\tilde{k}}_{k}^{{\rm{RR}}}$. Performing these operations, one obtains

Equation (73)

Equation (74)

Equation (75)

$k\in { \mathcal K }$. Equations (73)–(75) show that the group rate coefficients for photo-ionization and radiative recombination are formally identical to those for the StS model (32) and (33) and that the former can be obtained from the latter by replacing the statistical weights with the group partition functions, and the elementary cross-sections with the group cross-sections.

The group energy transfer terms due to photo-ionization and radiative recombination are

Equation (76)

$k\in { \mathcal K }$, where energy transfer rates due to photo-ionization and radiative recombination are ${\tilde{G}}_{k}^{{\rm{PI}}}=\langle {E}_{i}\,{k}_{i}^{{\rm{PI}}}{\rangle }_{k}$ and ${\tilde{G}}_{k}^{{\rm{RR}}}\,={\sum }_{i\in {{ \mathcal I }}_{k}}{E}_{i}\,{k}_{i}^{{\rm{RR}}}$, respectively.

By repeating the above procedure, the net volumetric energy loss rates for the material gas (34) and free electrons (35) become

Equation (77)

Equation (78)

ff transitions—The emission and absorption coefficients, and the net volumetric energy loss rate due to ff transitions are not affected by the grouping as only charge–charge interactions are taken into account in this work. Hence, Equations (36) and (37) are used in the same way as for the StS model.

Appendix C: Free-electron Energy Governing Equation

This appendix provides a brief discussion on how to obtain the governing equation for free-electron energy as given in Equation (38). Before moving to the mathematical details, it is worth recalling that in a plasma, polarization effects (i.e., charge separation) appear due to gradients of macroscopic quantities. This creates polarization fields, which tend to restore charge neutrality and act against further charge separation (Zel'dovich & Raizer 2002, Chapter 7).

The starting point in obtaining the governing equation for the free-electron energy is the momentum equations for free electrons. In the absence of transport phenomena and under the assumption that the bulk velocity of free electrons is the same as that of the whole gas, the above equation reads (Zel'dovich & Raizer 2002, Chapter 7):

Equation (79)

where the symbol ${\boldsymbol{E}}$ denotes the polarization electric field, and the Lagrangian (or substantial) derivative operator is $D()/{Dt}\,=\partial ()/\partial t+{\boldsymbol{v}}\cdot {\rm{\nabla }}()$. Since electrons have a very small mass, one may safely neglect the inertia term in Equation (79) to obtain

Equation (80)

The complete equation for free-electron energy accounting for (i) the work of the polarization electric field and (ii) collisional and radiative energy losses is

Equation (81)

where ${v}^{2}={\boldsymbol{v}}\cdot {\boldsymbol{v}}$. In analogy with that done for the momentum equation (79), the kinetic energy terms, ${n}_{{\rm{e}}}\,{m}_{{\rm{e}}}{v}^{2}/2$, on the left-hand side of Equation (81) may be ignored to give

Equation (82)

where the free-electron energy density (8) has been introduced. The substitution of the expression for the polarization electric field (80) in Equation (82) leads to

Equation (83)

By expanding ${\rm{\nabla }}\cdot ({p}_{{\rm{e}}}\,{\boldsymbol{v}})={\rm{\nabla }}{p}_{{\rm{e}}}\cdot {\boldsymbol{v}}+{p}_{{\rm{e}}}{\rm{\nabla }}\cdot {\boldsymbol{v}}$, Equation (83) becomes

Equation (84)

Under the assumption of steady and one-dimensional flow, Equation (84) reduces to the free-electron energy equation as given in Equation (38).

Footnotes

  • In the NIST atomic database, information on the substates having $6\lt n\lt 13$ is only provided for those with orbital quantum number $l=0,1,2$ (s, p, and d states, respectively), whereas they are not given at all for the higher states (see Table 2 in Appendix A).

  • The calculation of electron transport properties and fluxes is often accomplished by accounting only for the effects of elastic collisions (Devoto 1966, 1967). When assuming that the electron-heavy collision integrals do not depend on the particular electronic state (as done in this work; see Equation (19)), it can be shown that the electron transport formulas do not show an explicit dependence on the population of excited electronic states.

  • The non-relativistic solution of Schrödinger's equation for an isolated hydrogen atom predicts that the energy and the statistical weight of a bound energy state with principal quantum number n are, respectively, ${E}_{n}=-{I}_{{\rm{H}}}/{n}^{2}$ and ${g}_{n}=2{n}^{2}$, where ${I}_{{\rm{H}}}$ is the hydrogen ionization potential (Pauling & Wilson 1935, Chapter 5). Hence, the statistical weight grows rapidly with increasing energy.

Please wait… references are loading.
10.3847/1538-4357/aa602e