Photoionization and Electron-Ion Recombination in
Photoionization and Electron-Ion Recombination in
Photoionization and Electron-Ion Recombination in
Review
Photoionization and Electron–Ion Recombination in
Astrophysical Plasmas
D. John Hillier
Department of Physics and Astronomy & Pittsburgh Particle Physics, Astrophysics and Cosmology Center
(PITT PACC), University of Pittsburgh, 3941 O’Hara Street, Pittsburgh, PA 15260, USA; hillier@pitt.edu;
Tel.: +1-412-624-9000
Abstract: Photoionization and its inverse, electron–ion recombination, are key processes that influ-
ence many astrophysical plasmas (and gasses), and the diagnostics that we use to analyze the plasmas.
In this review we provide a brief overview of the importance of photoionization and recombination
in astrophysics. We highlight how the data needed for spectral analyses, and the required accuracy,
varies considerably in different astrophysical environments. We then discuss photoionization pro-
cesses, highlighting resonances in their cross-sections. Next we discuss radiative recombination, and
low and high temperature dielectronic recombination. The possible suppression of low temperature
dielectronic recombination (LTDR) and high temperature dielectronic recombination (HTDR) due to
the radiation field and high densities is discussed. Finally we discuss a few astrophysical examples to
highlight photoionization and recombination processes.
1. Introduction
One of the most important ways we learn about the Universe is through spectroscopy.
From spectroscopy we can typically deduce important stellar parameters such as a star’s
effective temperature,1 surface gravity (=GM∗ /R∗ 2 ) and abundances. These in turn provide
Citation: Hillier, D.J. Photoionization
insights into stellar evolution, galactic evolution, and the evolution of the Universe. To
and Electron–Ion Recombination in
perform analyses of stellar data requires atomic data, although the amount and type of
Astrophysical Plasmas. Atoms 2023,
atomic data needed varies greatly with the application. In the most extreme cases, in
11, 54. https://doi.org/10.3390/
which local thermodynamic equilibrium does not hold (discussed below), we require, for
atoms11030054
example, photoionization cross-sections, oscillator strengths for bound–bound transitions,
Academic Editors: Sultana N. Nahar line-broadening data, collisional cross-sections, autoionization rates, chemical reaction rates,
and Guillermo Hinojosa and charge exchange cross-sections. For some simple species, such as hydrogen, we have
Received: 24 January 2023
excellent atomic data while for other important species, such as Fe group elements, crucial
Revised: 3 March 2023 atomic data is lacking. Unfortunately, for many ionization stages even basic information,
Accepted: 3 March 2023 such as accurate energy levels, is also missing.2 However, invaluable work by the NIST and
Published: 9 March 2023 Imperial college atomic spectroscopy groups is helping to rectify this situation for some
important astrophysical ions (e.g., [2,3]).
In this review we discuss the importance of photoionization cross-sections for astro-
nomical applications, with an emphasis on massive stars. Such a discussion will necessarily
Copyright: © 2023 by the authors.
consider recombination processes—the inverse of photoionization processes. Before doing
Licensee MDPI, Basel, Switzerland.
so it is necessary to define some important physical and astronomical terms.
This article is an open access article
A star (or any other astrophysical object emitting radiation) is not in thermal equilib-
distributed under the terms and
rium. However, in some cases, and at some locations, it is well justified to assume that
conditions of the Creative Commons
Attribution (CC BY) license (https://
the plasma in the star is in “local” thermal equilibrium. Below the atmosphere (the thin
creativecommons.org/licenses/by/
layer that emits the radiation we observe) the material in most stars can be considered
4.0/). to be in “local” thermal equilibrium (LTE). In such cases, the state of the gas (e.g., the
thermodynamical properties, the ionization state, and the populations of atomic levels) are
set by the density and electron temperature via thermodynamic arguments. The ionization
state of the gas and the level populations are determined, for example, by the Saha and
Boltzmann equations (e.g., [4]). Moreover, there is only one temperature—the electron tem-
perature, the ion temperature, the excitation temperature, and the radiation temperature
are all identical. The temperature of the gas varies with location (it must, since radiation is
propagating outwards) but the scale on which it varies does not affect the thermodynamic
state of the gas. Unfortunately, much of the radiation we observe comes from gas that is
NOT in LTE—typically referred to as nLTE or non-LTE.
At the stellar surface the assumption of LTE becomes less valid. This is not surprising—
at the stellar surface radiation is escaping from the star and there is no incident radiation (at
least for single stars), and hence the radiation density must drop from its blackbody value
by a factor of (at least) ∼2 (since there is no incident radiation—see (see [4], p. 120). Further,
because radiation can now travel significant distances, regions of different temperatures
are directly coupled, potentially making the radiation field at a given location strongly
non-Planckian.
Fortunately, in some cases the densities are high enough that collisional processes can
still strongly couple the level populations and the ionization state of the gas to the local
electron temperature, allowing us to use the Saha and Boltzmann equations to compute
level populations. In such cases the electron temperature (which will be the same as the ion
temperature) determines the state of the gas and it is this temperature that we normally
state. In general, however, the electron and radiation temperatures will be different.3
The Boltzmann equation, which relates the populations of two levels within the same
ionization state, is
gu n∗l
n∗u = exp(− Elu /kT ) (1)
gl
where n∗l and n∗u are the LTE population densities of the lower and upper states, respectively,
gl and gu are the level degeneracies, and Elu is the difference energy between the two levels
(e.g., [4]).
The Saha equation, which relates the ground state populations of two consecutive
ionization equations, is
∗ ∗ g1,i C I
n1,i = n1,i +1 Ne exp (−ψi /kT ) (2)
g1,i+1 T 3/2
where T is the electron temperature in Kelvin, C I = 2.07 × 10−16 (cgs units), Ne is the
electron density, ψi is the ionization energy, and the subscript i (i + 1) is used to denote the
ionization stage.
In many stars, and especially those with lower density gas (nebulae, stellar winds,
supernovae) the departures from LTE are significant and MUST be allowed for. Gaseous
nebulae, which typically have densities less than 106 atoms cm−3 , provide an excellent
example in which the departures from LTE are extreme. The radiation field that ionizes the
nebula typically emanates from a hot star (Teff & 25,000 K) and is strongly diluted since
the nebula is typically very distant from the star (Figure 1). The equilibrium temperature
of the gas is typically around 10,000 K (which is primarily determined by the chemical
composition of the nebula) and is insensitive to the effective temperature of the star. The
nebula, longward (i.e., at larger wavelengths) of the H I Lyman jump at 912 Å, is transparent
to most radiation. Because of the non-Planckian radiation field and the low densities,
collisions with electrons cannot drive the gas into LTE.
Atoms 2023, 11, 54 3 of 27
Figure 1. A composite image of the Helix nebula obtained using the HST and the Cerro Tololo Inter-
American Observatory in Chile. Hα+[N II] (λλ6548.1, 6583.4 Å), emission is shown in red, an average
of Hα+[N II] and [O III] (λλ4958.9, 5006.8 Å) is shown in green, and forbidden [O III] emission is
shown in blue [5]. The nebula lies at a distance of about 200 pc, and the size of the semi-major axis
is approximately 5.5 arc minutes or 66,000 AU [5]. The central star has an effective temperature
of 104,000 K, L ≈ 80 L , and R ≈ 0.028 R [6]. Using the semi-major axis as a representative size
scale, the dilution factor ( 0.25(r/R∗ )2 ) is of order 10−18 . Detailed insights into the structure and
morphology of the Helix nebula can be found in many works (e.g., [5,7]). Image credit: NASA, ESA,
C.R. O’Dell (Vanderbilt University), and M. Meixner, P. McCullough, and G. Bacon (Space Telescope
Science Institute).
When LTE no longer holds (i.e., nLTE) we are forced to solve for the ionization state
of the gas and the level populations from first principles. That is, we need to consider all
the processes, and inverse processes, that populate a given level. These processes include
photoionization and recombination, bound–bound emission and absorption, collisional
excitation and de-excitation, collisional ionization and collisional recombination, dielec-
tronic recombination and autoionization, charge exchange reactions and, in “cooler gas”
(T / 6000 K), chemical reactions, and dust chemistry.4 Determining the state of the gas is a
complicated problem. Many of the processes above depend on the radiation field which
in turn depends on the level populations. Further, the radiation field couples the gas to
regions of different temperatures. This is a highly non-linear problem and can only be
solved by iterative techniques. In many cases we can consider the population numbers to
be static (and the equations are referred to as the equations of statistical equilibrium) but
in other cases (e.g., supernovae) we may need to allow for time dependence (and solve
the kinetic equations). Another major issue is correcting for plasma effects that limit the
number of levels in atoms and ions (i.e., crudely, an atom/ion cannot be larger than the
inter-atom spacing). One approach is probabilistic and was developed in a series of papers
by Hummer, Mihalas, and Dappen [12–14], and is used, for example, in the nLTE radiative
transfer codes TLUSTY[15] and CMFGEN [16]. In this approach, levels are assigned an
occupation probability that varies smoothly with the level energy, density and temperature,
and that leads to a finite partition function.
Thus, a vast amount of atomic data is needed. Sadly, despite heroic efforts by Bob Ku-
rucz [17,18], members of the Opacity Project [19] and Iron Project [20], Sultana Nahar [21],
and many others, much of the needed data is still missing. Extensive photoionization
data is available, for example, through TOPbase [22], TIPbase [23], and NORAD [21]. As
discussed later in this article, details can matter, and it is not always obvious, a priori,
which data are essential for accurate analyses. As this paper is concerned with photoioniza-
Atoms 2023, 11, 54 4 of 27
tion/recombination, this paper does not generally elaborate on other important processes.
Information on these additional processes can be found in, e.g., [24–27].
Below we discuss photoionization and recombination as relevant to astrophysical
applications. Much of the following discussion will be based on my own experiences
in developing CMFGEN, a nLTE radiative transfer code originally designed to model hot
massive stars ( M > 20 M , Teff ≥ 20, 000 K) and their stellar winds [16,28]. The winds in
these stars are driven by radiation pressure acting on bound–bound transitions belonging,
for example, to C, N, O, Ar, and Fe (e.g., [29–31]). Since its initial development the code
has undergone considerable revisions and improvements. It has been successfully used
to model O stars (e.g., [32–36]), Wolf–Rayet (WR) stars [37,38], luminous blue variables
(e.g., [39–41]), B stars (e.g., [42]), the central stars of planetary nebulae (e.g., [43]), and A
stars. Over the last decade CMFGEN was adapted to treat time-dependent radiation transfer,
and to solve the time-dependent kinetic equations [44], and it has been used to model
spectra resulting from a variety of SN explosions (e.g., [45–48]).
The review is organized as follows: In Section 2 we briefly discuss the importance
of photoionization processes for stellar interiors and introduce the Rosseland mean opac-
ity. We then consider photoionization processes in Section 3 with an emphasis on in-
ner shell ionizations in Section 3.1. Recombination processes are discussed in Section 4
while suppression of dielectronic recombination by collisions and the radiation field is
discussed in Section 5.1 and Section 5.2 respectively. Specific examples of where photoion-
ization/recombination processes are important are then discussed—direct recombination
(Section 6.1), the Sun (Section 6.2), O stars, WR stars, luminous blue variables (LBVs),
(Section 6.3), Of and WN stars (Section 6.3.1), carbon lines in WC stars (Section 6.3.2), C II
in a [WC] star (Section 6.4), and supernovae (Section 6.5).
2. Stellar Interiors
In stellar interiors energy is transported by radiation and by convection, and in
degenerate stars by conduction. As LTE holds, photoionization processes have no direct
influence on the level populations (since the populations are determined by the Saha and
Boltzmann equations), but they do help to determine the temperature of the gas and they
do help to set the continuous radiation field that we observe. Due to the small mean-free-
path of photons, radiation transport is diffusive and in this regime a single quantity, the
Rosseland mean opacity, is required to describe the transport of radiative energy. At depth
in the star the radiation diffuses and the specific intensity (Iν ) at frequency ν is given by
µ dBν
Iν = Bν − (3)
χν dr
2hν3 1
Bν = 2
, (4)
c exp(hν/kT ) − 1
χν is the opacity, and µ is the angle between the radius and the direction of radiation
propagation. The radiative flux is simply given by
−1 dBν
Fν = . (5)
3χν dr
The negative sign in the expression for the flux arises because T, and hence Bν , decrease
with increasing r. Integrating over all frequencies yields a total radiative flux, F, given by
T 3 dT
F = −16σ (6)
3χ R dr
Atoms 2023, 11, 54 5 of 27
where σ is the Stefan–Boltzmann constant and χ R is the Rosseland mean opacity as defined by
R∞ 1
1 0 χ dBν /dT dν
= R∞ ν . (7)
0 dBν /dT dν
χR
3. Photoionization
The photoionization rate from a level l (in an arbitrary ion of arbitrary charge) can be
written as Z ∞
dnl 4π
= −nl σν Jν dν (8)
dt PR νo hν
where nl is the population density of level l, σν is the frequency dependent photoionization
cross-section (units are cm2 ), νo the threshold frequency for ionization, h is Planck’s constant,
and Jν is the mean intensity (erg cm−2 s−1 Hz−1 ). Jν is defined by
1
I
Jν = Iν dΩ (9)
4π
where dΩ is an increment in solid angle. If the radiation field is Planckian and isotropic,
Jν = Iν = Bν . (10)
Typically in model atmosphere codes the rates (integrals) are evaluated using numeri-
cal quadrature. Thus
dnl 4π
= −nl ∑ wi σi Ji (11)
dt PR i
hνi
where wi is the quadrature weight at frequency νi .
In a simple species such as hydrogen, the photoionization process is simply
H + hν → H+ + e− .
In more electron-rich species the process is more complicated since there are multiple
photoionization routes. For example, there are two direct photoionization routes from
C III(2s 2p):
C III(2s 2p1 Po ) + hν → C IV (2s2 S) + e− &
o
C III (2s 2p1 P ) + hν → C IV (2p2 Po ) + e− .
The first process occurs provided the photon energy exceeds the ionization energy of the
C III(2s 2p 1 Po ) state5 . The second process occurs when the photon energy exceeds the
sum of the ionization energy and the difference in energy between the 2s and 2p states in
C IV. Of course, photons of sufficient energy may also ionize C2+ by ejecting an inner (1s)
electron—a process of great importance when X-rays are present.
There may also be multiple indirect photoionization routes such as:
Figure 2. Illustration of the photoionization cross-section (in megabarn, with 1 Mb = 10−18 cm2 )
of C III 2p2 1 D. Ionization to the C IV ground state occurs via autoionizing levels such as C III 2p
4d 1 Fo . Shortward of ≈325 Å, photons have sufficient energy to ionize directly to C IV 2p 2 Po . The
data were convolved with a Gaussian profile with a full width at half maximum of ∼ 600 km s−1 (i.e.,
σ = 250.0 km s−1 ). The photoionization cross-sections for C III were computed by P. J. Storey (private
communication). In other photoionization cross-sections the resonances are often much narrower
than those shown here.
In LTE the final state arising from the photoionization process is irrelevant—only
the total opacity matters. In general, in nLTE, the final state matters, since each process
contributes to the population of a different state whose population needs to be determined
from first principles. In practice this is generally not a crucial concern for most spectral
modeling since the rates for processes connecting states within an ion are generally much
larger than the photoionization and recombination rates. However, there are cases where
the final-state-dependent cross-sections are important.
Atoms 2023, 11, 54 7 of 27
Figure 3. The C IV λλ1548.2, 1550.8 P Cygni profile in the O4 I(n)fp star Zeta Puppis. The x-axis
was computed using v = c(λ/λo − 1.0) where λo = 1548.187. The spectrum has been normalized
by the continuum spectrum (Fc )—i.e., a smooth curve drawn through spectral regions showing no
evidence for bound–bound absorption or emission. Strong blue shifted absorption is seen, indicating
an outflowing stellar wind with a terminal velocity (V∞ ) in excess of 2600 km s−1 . The redshifted
emission primarily arises from continuum photons that were emitted in other directions, absorbed by
C IV, and subsequently scattered to the observer. The two narrow absorptions near 0 and 500 km s−1
are due to absorption by C IV in the interstellar medium.
4. Recombination
The recombination rate is given by
∗ Z ∞ 2hν3
dnl nl 4π
= nK σν + Jν exp(−hν/kT )dν (12)
dt RR nK νo hν c2
(e.g., [4]) where the subscript K refers to the recombining ion and the LTE population is
computed using the actual electron density. The quantity (nl /nK )∗ (for a given level) is
only a function of the electron density and temperature. When the gas is in LTE, and when
Jν = Bν , the photoionization and recombination rates (absolute values) are identical.
In my work I treat recombination as the reverse process of photoionization and hence
in CMFGEN recombination rates are computed using the photoionization cross-sections.
As noted earlier, rates are evaluated using numerical quadrature, and identical weights
are used for both the forward and reverse process. At high densities it is desirable to
treat both processes identically since small differences can cause erroneous populations
to be determined when solving the kinetic equations. At depth, where LTE conditions
apply, it is important that they identically cancel. Generally the weights are evaluated
using the trapezoidal rule—more accurate quadrature schemes are generally not feasible
because of the complex frequency dependence (and depth dependence) of the radiation
field, and because the same quadrature scheme must be used to compute the rates for both
photoionization and recombination. Care must be taken near bound-free edges, since the
integrand in the recombination rate can vary rapidly with frequency—especially true for
highly ionized states at low temperatures since the recombination rate at frequency ν scales
as exp[−h(ν − νo )/kT ].
For low densities, such as those found in H II regions, planetary nebulae, and many
collisionally ionized plasmas, recombination rates are often evaluated separately, and
treated as a distinct process. At “low” densities most transitions are optically thin, and
recombination into high states simply cascade into the ground state and metastable levels.
In a H II region, for example, the ionization of H is maintained through photoionizations
from the ground state and photoionizations from excited states can be ignored. However
transitions to the ground state can be optically thick. Consequently two limiting cases
are considered when computing H line strengths—Case A, in which all transitions are
assumed to be optically thin, and Case B, in which only the Lyman transitions are optically
thick (e.g., [25]). Under the optically thick assumption the rate of decays in a transition is
assumed to be exactly balanced by the rate of radiative excitations in the transition.
Atoms 2023, 11, 54 9 of 27
2.07 × 10−16 gu
n∗u = Ne NCIV exp (−ψl /kT ) (13)
T 3/2 gCIV
(e.g., [4]). In the above formula gu is the statistical weight for the 2p 4d 1 Fo state, gCIV
is the statistical weight of the C IV ground state (2s 2 S), Ne is the electron density, NCIV
is the ground state population of C IV, and ψl is the energy of the 2p 4d1 Fo state above
the ground state of C IV. At 104 K that single transition leads to a LTDR recombination
coefficient (defined as the (LTDR rate)/Ne /NCIV ) of 3.1 × 10−12 cm3 s−1 (P. J. Storey, pri-
vate communication) which is essentially identical to the direct recombination rate of
3.2 × 10−12 cm3 s−1 [83].
Thus, we see the following:
1. The LTDR rate is very sensitive to ψ when ψl /kT is of order unity or larger.
2. When ψl /kT << 1, the LTDR recombination rate scales as T −3/2 and thus increases
more quickly with decreasing temperature than the radiative recombination rate,
which typically scales as T −α with α ∼ 0.7. (see, e.g., [83]).
3. The LTDR process will be most important for those states with a large Einstein A
coefficient and for those states lying closest to, but above, the ion ground state.
4. The process is very dependent on the details of the atomic structure. In the above case,
the energy of the 2p 4d 1 Fo state is crucial for determining the LTDR rate. As the LTDR
autoionizing states lie well above the C III ground state, and can have large energy
Atoms 2023, 11, 54 10 of 27
widths, the energies of the states are not necessarily known. Theoretical calculations
can provide estimates, but will have difficulties for states that lie ”very close” to the
ionization limit since a small error in the energy level can make a big difference in the
recombination rate, particularly at low temperatures.
Energy
2p 2Po 2p nl 1L’
2s 2S 2p 4d 1Fo
2s nl 1L
2p2 1P
2p2 1D
2s 2p 1Po
2s2 1S
Figure 4. Simplified pseudo-Grotrian diagram for C III and C IV to help illustrate LTDR and HTDR.
Five bound C III levels are shown in black and two C IV levels are shown in red. Example HTDR
autoionizing levels (e.g., 2p 50p, 2p 50d, etc.) that converge on the C IV 2p 2 Po state are shown in
blue. One of the most important LTDR autoionizing levels is shown in orange—it has been moved up
slightly in energy to separate it more clearly from the C IV ground state—the horizontal dashed red
line is used to indicate the energy of the C IV ground state. The blue line shows a HTDR transition,
which has a wavelength approximately equal to that of the C IV resonance transitions (1548, 1551 Å),
while the dashed orange line shows a LTDR transition (at ∼412 Å). Triplet levels also experience both
LTDR and HTDR.
The LTDR rate can exceed the direct recombination rate, and in many cases plays a cru-
cial role in determining nLTE level populations, and observed line strengths
(e.g., [81,82,84]).
The LTDR process is complicated by states that are forbidden to autoionize in LS
coupling, such as the 2p 4d 3 Do state in C III or the quartet states in C II. In such cases the
populations of these levels must be determined by solving the rate equations. These levels
will be collisionally and radiatively coupled to states that can autoionize and, because of
departures from LS coupling, they can also have non-zero autoionization rates which are
larger than the radiative decay routes from the state. Thus, these levels can be an important
additional recombination channel.
In CMFGEN we handle the quartet states in C II as part of our atomic models while the
doublet autoionizing states are assumed to be in LTE with respect to the ground state of
C III and are not directly treated. Recombination through the quartet states is treated via
the line transitions connecting them to lower levels, while transitions for the autoionizing
states are treated via the photoionization cross-sections. Generally we assume the states
within a term are populated according to their statistical weights, although this will not
be valid for some levels since the autoionizing rates can depend strongly on their total
angular momentum. For example, the autoionizing probabilities for the C II 2s 2p(3 Po )
4s 4 Po j = 1/2, 3/2, and 5/2 states are 5.3 × 109 , 1.3 × 1010 , and <3 s−1 . These were obtained
from the full width at half maximum tabulated by [84]. One issue, potentially important at
high densities, is that we do not have accurate collisional cross-sections for the states not
permitted to autoionize in LS coupling.
For C III we typically assume, following [82], that all low lying states can autoionize.
LTDR is easily taken into account via the photoionization cross-sections; however, the
assumption is only necessarily valid for those states in which the autoionization rates
(greatly) exceed other processes populating/depopulating the autoionizing state.
A potential problem in nLTE calculations is that, due to difficulties of current atomic
codes to compute accurate energies, the resonances in the photoionization cross-sections
are offset from their true positions. Such offsets are probably unimportant when computing
Atoms 2023, 11, 54 11 of 27
the Rosseland mean opacity, but can be important for spectral studies. First, an inaccurate
energy will influence the location of observable resonances in stellar spectra. Second, it can
have an effect at “low” temperatures due to the scaling of the LTDR rate with temperature
(exp(−ψ/kT )/T 3/2 ). Third, complicated nLTE effects could arise. For example, a wrong
resonance wavelength can potentially cause issues if a strong resonance coincides (or now
does not coincide) with another bound–bound transition (since a strong resonance can
affect the radiation field in the transition and vice versa).
The direct inclusion of resonances in photoionization cross-sections also has other
potential issues. First, the resonances vary much more rapidly than the background cross-
section and hence a very fine frequency grid needs to be used—this is particularly true for
narrow resonances. For computational expedience, we typically sample the continuum
cross-sections in CMFGEN every 500 km s−1 (but finer near level edges). To avoid aliasing10
we smooth the cross-sections. In early versions of the atomic data the cross-sections were
smoothed to a resolution of 3000 km s−1 but in new data sets we no longer store the
smoothed cross-sections. Instead, newer cross-sections can be smoothed to the desired
resolution, set by a control parameter, when they are read in.
Second, a narrow resonance can mean that the autoionization lifetime of the upper
level may be comparable to, or even larger than, radiative transitions from the same level.
As a consequence the upper level may not be in LTE with respect to the ion and hence
the photoionization cross-section should not be used to compute the recombination rate.
When identified, such a resonance should be clipped out and the upper levels treated as a
bound state.
Third, photoionization cross-sections are usually computed in LS coupling. This
means, for example, that the multiplet structure of the resonances is not treated—a problem
more crucial when the resonances are “narrow”.
Figure 5. Comparison of the recombination coefficients for normal radiative recombination, for
LTDR and for HTDR for recombination from C3+ to C2+ . Only at low temperatures (.1000 K) does
radiative recombination dominate. Above this temperature LTDR dominates until a temperature of
∼17,000 K, at which time HTDR becomes the dominant recombination mechanism.
φ(ν) is the line absorption/emission profile (which, in general, is determined by the finite
lifetimes of the levels involved in the transitions, thermal motions of the atoms, and the
interaction of the radiating atom/ion with its neighbors (see [4], Chapter 9)) and Slu is the
line source function given by
2hν nu /gu
Slu = 2 . (16)
c nl /gl − nu /gu
Atoms 2023, 11, 54 13 of 27
In LTE nu = ( gu /gl )nl exp(−hν/kT ) and hence Slu simplifies to the Planck function. As
is readily apparent the net rate does not directly depend on the optical depth—such a
dependence only occurs indirectly through the dependence of J̄ on the optical depth.
In nebula conditions J̄/Slu is typically << 1 (since the nebula is very distant from the
star the radiation field is greatly diluted) and the contribution to the recombination rate by
this single transition is simply nu Aul . When the rates are summed over all resonances you
recover the LTDR/HDTR recombination rate. However, such a rate is typically an upper
limit since the radiation field can reduce this rate.
At depth in a stellar atmosphere J̄ ≡ Slu ≡ Bν , and thus the net LTDR and HTDR
rates are identically zero—that is, every downward transition in the stabilizing transition
is balanced by an upward transition. However, above the atmosphere the temperature
of the radiation field and the electrons are not the same. Typically J̄ will fall below Slu ;
however, in a wind J̄ can be greater than Slu in some transitions. In Figure 6 we show the
mean intensity (in the comoving frame) and the blackbody mean intensity at a temperature
of 1.6 × 104 K and a density of ∼1011 electrons cm−3 —roughly 50% of the emission in the
line referred to as C III λ2297 originates above that density. From that figure we see that
the radiation field at the wavelength of the C IV resonance transition, and at/near the
stabilizing transition, is close to a blackbody at the local electron temperature. Thus, the
radiation can act to suppress HTDR.
In Figure 7, we show the recombination and photoionization rates for n = 26 through
30 singlet states of C III (treated as a single level) for a test calculation in which we included
HTDR transitions for levels up to n = 30, and with no suppression of the recombination rate
with the angular orbital quantum number. The resulting model spectrum is almost identical
to the spectrum computed without HTDR—a consequence of the low temperatures in the
C III line-formation region and the suppression of HTDR via the radiation field in the
stabilizing transition. This result is model dependent—in practice the importance of HTDR
needs to be examined on a case-by-case basis.
Figure 6. Comparison of the mean intensity (in the comoving frame, red curve) and the blackbody
mean intensity (green dashed curve) at a temperature of 1.6 × 104 K and a density of ∼1011 electrons
cm−3 —roughly 50% of the emission in C III λ2297 originates above that density. At this density
the C IV 2s and 2p states are collisionally coupled and as a result the radiation field near the wave-
lengths of the C IV transitions at 1548, 1551 Å is also close to that of a blackbody. The blue curve is
similar to the red curve except that we used a Voigt profile for the line absorption/emission profile
(see Equation (15)), rather than the simple Doppler profile (see [4]) which is generally used when
computing the atmospheric structure and level populations. The use of a Voigt profile is crucial for
explaining the observed profile of the C IV resonance doublet at 1548, 1551 Å in WC stars [91].
The influence of the radiation field is not an issue if the resonance is included as part
of the photoionization cross-section as the influence of the radiation field is automatically
taken into account. It is also not an issue for “autoionizing” states treated as bound levels,
since the radiation field is again taken into account. However, it is a potential issue if
Atoms 2023, 11, 54 14 of 27
the LTDR or HTDR rate is included as a separate process, and the inverse process is
not included.
Figure 7. Illustration of the photoionization and recombination rates to/from a combined level
(n = 26, . . . , 30; singlet levels only) in C III. The solid curves show the rates with HTDR included, while
the dashed curves show the rates without HTDR included. At most densities the recombination and
photoionization rates agree closely in the model with HTDR. In such regions the influence of HTDR
on the ionization structure is suppressed. Towards the outer boundary the HTDR recombination rate
converges to that of a model without HTDR due to the decrease in electron temperature.
6. Astrophysical Examples
Below we discuss some examples of where photoionization data is crucial. It is
unrealistic to discuss all cases, since photoionization data is crucial in any photoionized
plasma and is crucial for nLTE analyses. For some plasmas, in which collisional processes
dominate, photoionization is less important but the inverse process (recombination) is
still critical.
optically thin. In hot stars the departure coefficients (=n/n∗ ) for H and He II levels typically
rise above the photosphere. Bound–bound transitions are optically thick, preventing
cascades. The radiation is diluted, hence recombinations into a level typically exceed
photoionizations from that level. Eventually, however, photon escape in lines becomes
important and the departure coefficients decrease. However, one must be careful with
generalizations—at some wavelengths (particularly in the Wien regime of the blackbody
curve) the rapid fall of the electron temperature with height above the photosphere means
that the energy density in the radiation field may initially exceed the radiation energy
density predicted by the blackbody formula for the local electron temperature.
Bound–bound processes are crucial for determining line strengths. However, it is
ultimately photoionization and recombination that determines the ionization state. In some
cases, charge exchange processes are crucial [96,97]. Particularly important are charge
exchange process of neutral H with, for example, Fe2+ and O+ . The reaction
O+ + H
O + H+
is resonant, has a total rate coefficient of order 10−9 cm3 s−1 [98], and is crucial for determin-
ing the O+ /O ratio in regions where the neutral hydrogen fraction exceeds roughly 10−3 .
spectrum formation occurs in the stellar wind and nLTE spherical models that treat the
wind are essential.
Figure 8. Illustration of the influence of LTDR on the strength of several N IV emission lines. The
model in red is the full calculation, while for the blue spectrum we omitted LTDR transitions in the
calculation of the nLTE populations. A N IV transition at 1718 Å is also influenced by LTDR. For
these calculations we used smooth N IV photoionization cross-sections and the LTDR data of [81,82].
Calculations using the opacity photoionization cross-sections of [111], that were obtained from [22],
yield a spectrum very similar to that shown in red. The model has a luminosity of 3 × 105 L , a radius
(at a Rosseland optical depth of 2/3) of 2.9 R , an effective temperature of ∼78,000 K, a mass-loss rate
of Ṁ = 1.5 × 10−5 M yr−1 , and a volume filling factor (which characterizes the degree of clumping
in the wind) of 0.1.
The electron temperature structure and wind velocity of a model for the LMC WC4
star, BAT99-9, is shown in Figure 9. The non-Planckian nature of the radiation field at two
depths in the wind is illustrated in Figure 10.
Figure 9. Illustration of the electron temperature (blue) and velocity structure (red) in the model for
the WC4-type star, BAT99-9. The model has log L/L = 5.48, an effective temperature of 84,000 K,
and a mass-loss rate of 1.4 × 10−5 M yr−1 .
Figure 10. An illustration of the radiation field in a WC4 star at a depth where V (r ) = 0.67 × V∞
( Ne = 1.7 × 1013 cm−3 ) (top panel) and at a depth V (r ) = 0.934 × V∞ ( Ne = 1.2 × 1011 cm−3 ) (lower
panel). Shown also is the blackbody spectrum at the local electron temperatures (Te = 4.90 × 104 K
and Te = 1.61 × 104 K). In the top panel (where V (r ) = 0.67 × V∞ ) we also illustrate the blackbody
spectrum at the effective temperature (Te = 8.19 × 104 K) that has been normalized to match the
model spectrum at 5.0 µm. As readily apparent, the local radiation field is very different from that
defined by the blackbody at the local electron temperature or that defined by the effective temperature.
Consequently, and because of the low electron densities, the ionization state of the gas, and the level
populations, are far from their LTE values. Due to the strong departures from LTE accurate atomic
data is crucial for determining the state of the gas and hence for predicting the stellar spectrum.
O (O III through O VI). To understand driving at the base of the wind additional ionization
stages are needed—in some models we include Fe IV through Fe XVII.
Figure 11. Illustration of the stratified ionization structure in the WC4 star, BAT99-9. On the top axis
we show the Rosseland optical depth (left plot) and the electron density in units of electrons per cm3
(right plot). The top scales are not linear. As we move out in the wind, the ionization decreases. This
variation is verified by the emission profile line widths—O VI lines are narrower than O III lines.
The stratified ionization structure is a consequence of several factors. First, the winds
of WR stars are not transparent. For example, the He II Lyman continuum (shortward
of 228 Å) is optically thick. Further, the transparency is a strong function of wavelength.
Second, as we move out in the wind the radiation field becomes diluted. Third, the intense
radiation field in some spectra bands can pump low lying levels. Because of this, and
because of the high densities which reduce cascades, ionizations from excited states can
play an important role in determining the ionization state of the gas.
In Figure 12 we illustrate the photoionization rate, normalized by the total recombi-
nation rate to all (included) levels. The normalization was primarily chosen to emphasize
the important process in the line formation region. In the inner regions of these dense
winds photoionizations and recombinations to each level will be in detailed balance. As we
move out in the wind the photoionization from most levels will decrease due to dilution of
the radiation, although for some levels the photoionization and recombination rates may
maintain equality if the continua are optically thick. For C III we see that three levels, in
order of importance, control the ionization—2s 2p 3 Po , 2s2 1 S, and 2s 2p 1 Po . On the other
hand, for C IV it is the n = 3 levels (3s, 3p, and 3d) that help to determine the C IV/C V
ionization ration. The ionization eventually shifts because the radiation is becoming diluted
(as 1/r2 ) and the populations of the n= 3 levels are also declining. One reason for the
difference in behavior of C III and C IV is there is often a rapid decline in the strength of
the radiation field shortward of ∼228 Å.
The presence of multiple ionization stages in the wind results in emission from mul-
tiple ionization stages. For example, in the case of BAT99-9 we see emission from two
ionization stages of carbon (C III & C IV)15 and four ionization stages of oxygen (O III
through O VI) with the characteristic line width (after allowance for blending and for the
formation mechanism) decreasing as the ionization increases. The origin of one O and two
C lines is shown in Figure 13—it shows that a given emission line originates over a range
of radii and that lower ionization features form farther out in the wind.
Atoms 2023, 11, 54 19 of 27
Figure 12. Illustration of the model photoionization rates for the WC4 star, BAT99-9. For illustration
purposes, the rates for the 3l states have been combined in the right plot.
Figure 13. Illustration of where O VI λ3811, C IV λ5471, and C III λ2297 originate in the wind.
R
ζ d ln r is proportional to the emission. High excitation lines originate in the inner denser wind,
while lower excitation lines originate in the outer wind.
Figure 14. A small section of the spectrum of the WC 11 star J0608 (blue) which was obtained
by Nidia Morrell (private communication). A model fit (in red) is shown and model lines are
identified. The broad feature near 5115 Å (the first unidentified feature) is due to a resonance in the
C II free–free (i.e., Bremsstrahlung) cross-section. It arises from the 2s 2p(3 Po )3d 2 Po − 2s 2p(3 Po )4f 2 D
transition [117–119]. Both these states are autoionizing.
In Figure 14, we also show a free–free resonance (λ∼ 5115 Å) previously identified in
CPD–56o 8032 [117]. The observations were obtained with a resolution of ≈ 7 km s−1 and
hence the line is resolved. In CMFGEN the resonance is treated as a free–free resonance since
both levels involved in the transition are autoionizing (with A∼1011 s−1 ; [117]). In the case
of this free–free resonance it was trivial to omit it from a “continuum” calculation. The latter
is needed so we can rectify the spectrum (i.e., normalize the continuum to unity). However,
this is not the case for bound-free resonances that appear in the photoionization cross-
sections. Such resonances can appear in the computed continuum, distorting an otherwise
smooth spectrum. These resonances riddle the UV continuum spectrum. However, in
practice they are difficult to discern because of the rich forest of bound–bound transitions
which mask the continuum spectrum.
6.5. Supernovae
Supernovae are fascinating objects. They represent the end points of evolution for
many stars and are an important source of metals (astronomical jargon for all elements more
massive than He) in the Universe. Broadly speaking there are two classes of supernovae
—those arising from the core collapse of a massive star (e.g., [120]) and those arising from
the thermonuclear detonation of a white dwarf (WD) star (a compact object of stellar origin
with a mass less than ∼1.4 M that is supported by electron-degeneracy pressure). The
latter class is designated as a Type Ia SN and, while we know that it involves a WD, we do
not know in what type of binary system the explosion occurs.
An extensive discussion of the possible progenitors of Type Ia SNe is given by [121].
Type Ia SN could arise when the WD accretes hydrogen-rich material from a “normal” star
(e.g., a red supergiant or a main sequence star). As a WD accretes mass its radius shrinks
(assuming it does not eject the accreted mass via a surface explosion in an event called a
nova, which is believed to occur in many systems). As it approaches the Chandrasekhar
mass of 1.414 M (the upper mass limit or a WD star) it will undergo a thermonuclear
explosion. Another possibility is that the WD star accretes He rich material from a WD
companion. This material undergoes a surface thermonuclear explosion which triggers
an inward propagating shock that triggers the detonation of the accreting WD. A third
possibility involves collisions and mergers of two WD stars. In the first scenario the
exploding WD has a mass of 1.4 M , while in the other two cases the mass of the exploding
WD is (typically) less than 1.4 M . The different scenarios predict different chemical
compositions for the ejecta and thus determining the chemical composition of the ejecta
offers a potential means of determining the nature of exploding WD.
Atoms 2023, 11, 54 21 of 27
At late times (say 200 days) the spectra of Type Ia supernovae ejecta are dominated
by emission lines of Fe, although lines due to Ni, Ca, and S are also present. One issue
with current models of the ejecta is that they fail to yield an iron spectrum in agreement
with observation (e.g., [45,122]). Basically, the Fe II lines are too weak relative to Fe III
(Figure 15) and this limits our ability to interpret ejecta observations. Is the issue related
to a problem in the ejecta explosion models, is it due to the ejecta being clumped (which
enhances recombination and hence lowers the ionization), is it due to issues with the iron
atomic data, is it due to problems treating the thermalization of high-energy electrons [122],
or are we missing additional physics? Unfortunately in these systems the iron atomic data
is of crucial importance, since Fe II/ Fe III is of order unity and the lines of both species
probably form in the same region. Thus, a factor of 2 error in the Fe II recombination
rate will make a factor of 2 error in the Fe II/ Fe III ionization fraction and will change
the relative line strengths (which are produced via collisional excitations) by a factor of 2.
Fortunately, for Fe II and Fe III, HTDR is unimportant at the relevant temperatures, as can
be gleaned from the rates provided by [123]. Nebular spectra of Type Ia SNe 2
Figure
Figure1. Comparison
15. Illustration
of SUB1of anda CHAN
small models
section of the spectrum
to SN2011fe. Flux valuesofforaSN2011fe
Ia SN showing the influence
have been scaled of7clumping
by 4.225 × 10 to account for models at 1
Then, all flux values were scaled by the same value that is of order unity. Data from SN2011fe are shown in brown and are 204.5 d post B-band maximum
which enhances recombination and hence reduces the ionization state of the gas. In the unclumped
two most striking discrepancies between the models and observations are as follows: (i) the [Fe III] lines are too strong relative to the [Fe II] lines (espe
inmodel,
the unclumped f = 1, Fe
with models), (ii) lines
and III near 5000ofÅthe
the overprediction are[Catoo strong.
II] doublet Increasing
near the clumping
7300 Å (especially The data were ftaken
(i.e., decreasing
in model CHAN). ) from the
Supernova
lowers theArchive and collectedthus
ionization, from improving
Yaron & Gal-Yamthe(2012) and Mazzali
agreement. et al. (2015).the emission feature near 7400 Å (a
However,
blend of Fe II, Ni II, and Ca II lines) is worse. The shape of the feature near 5000 Å is also sensitive to
we discuss the effects on iron features, nickel features, and IMEs, clumping in CMFGEN differs from that of a concentric shell-
the adopted
respectively. Fe IIwe
Finally, photoionization
summarize our work cross-sections
in Section 4. and updated calculations
structure, which iscould also improve
susceptible the fit. effects ac
to radiative-transfer
a shell and requires a large number of depth points to resolv
7. Conclusions shells.
2 TECHNIQUE The clumping factor scales many variables, such as dens
Accurate photoionization cross-sections are essential for many areas of astrophysics.
(scaled by 1/f) and the emissivities and opacities (calculated u
2.1
The Ejecta models quality varies greatly with the application.
required The biggest
the populations needs
in the clumps are scaled
and then at “in- by f). This m
termediate”
This research uses densities where
two hydrodynamic nLTE
models is relevant,
of Wilk et al. (2018), clumping
and where formalismdensity
complex leaves theand massradiation
column density unchan
DDC10
processes(CHAN) and SCH5p5
directly (SUB1).
affect levelDDC10 is a MCh (1.40
populations. M"is
This case forhasmany
) the CMFGEN incorporated this clumping
astrophysical method since Hilli
phenom-
model from Blondin et al. (2013). Model SCH5p5 is a sub-MCh Miller (1999) first applied it to massive stars.
ena associated with, for example, stellar winds, accretion At adisks, andof supernovae
time-step ejecta. (∼ + 200 d
216.5 d post-explosion
(1.04 M" ) from Blondin et al. (2017), but we have scaled the density
56
by 0.98 (see Wilk et al. 2018) in order to have the same Ni. Both maximum), we re-solved the relativistic radiative-transfer equ
Funding:
CHAN Partial
and SUB1 havesupport
the same 56for the
Ni at work
0.62 was provided
M" . However, CHAN by NASA for our theory
models SUB1
grant and CHAN using a clumping
80NSSC20K0524 and factor (
has a stable
STScI GrantironNocore, whereas SUB1 does not. STScI
HST-AR-16131.001-A. We useisCMFGEN
operated by 0.33,
the0.25, and 0.10 (values
Association partially motivated
of Universities for Re-by modellin
tosearch
solve intheAstronomy,
spherically symmetric, time-dependent, relativistic radiation-driven winds in massive stars; see, e.g. Hillier & M
Inc., under NASA contract NAS 5-26555.
radiative transfer equation allowing for non-LTE processes. We 1999; Najarro 2001; Bouret et al. 2003; Puls et al. 2003).
take
Data these two models atStatement:
Availability 216.5 d post-explosion
CMFGENfrom , and Wilk
theetatomic
al. range used
data of values
byisCMFGEN
also consistent
, arewith what hasatbeen inferre
available
(2018). Table 1 lists the initial (accessed
masses for each model as well as2023).
the clumps in planetary nebulae (Boffi & Stanghellini 1994; Stasi
www.pitt.edu/~hillier on 1 February This site also contains some older O star
mass abundance of calcium, iron, cobalt, 58 Ni plus 60 Ni, and 56 Ni. et al. 2004; Ruiz et al. 2011). Since the previous time-step
models. Data from supernova calculations can be downloaded from Zenodo or requested
not computed using clumping in the models, we from thescaled the in
We see CHAN has more than twice the amount of stable nickel –
appropriate
58 60author. A grid of spectra are available from the
M( Ni) + M( Ni) – than SUB1 as well as almost a factor of 2 Pollux data base [124]. A large grid of
input populations by 1/f between successive ionizations. This si
more
CMFGEN spectra models has been constructed and is being made
calcium. scaling available
is adequate[125,126].
since time-dependent
Hillier willeffects
alsohave a negli
provide CMFGEN models upon request. influence on the spectrum.
Acknowledgments: The author would like to thank P. J. Storey for extensive discussions on atomic
data, and for supplying atomic data on C II and C III that was directly used in this review. He would
also like to thank the numerous workers who have undertaken extensive atomic data calculations
and made their work freely available on the internet. A special thanks to Nidia Morrell who obtained
the high resolution spectrum of J0608. The invaluable comments made by the referees are also greatly
appreciated. This paper has made use of NASA’s Astrophysics Data System Bibliographic Services.
Conflicts of Interest: The authors declare no conflict of interest.
Abbreviations
The following abbreviations are used in this manuscript:
Notes
1 The effective temperature of the star is defined by the relation L = 4πR2∗ σSB Teff 4 where L is the stellar luminosity (energy emitted
per second), σSB the Stefan–Boltzmann constant, and R∗ is the radius of the star.
2 The meaning of “accurate” is highly context dependent. Atomic data calculations can, in some cases, give energy levels accurate
to 1% and for some purposes this is sufficient. However, for spectral modeling such energy levels cannot be used to compute
transition wavelengths—a 1% shift (which will be potentially larger if the levels are close in energy) will move a line far from its
correct location, influencing spectral synthesis calculations. Moreover, in non-LTE a wrong wavelength will influence how a line
interacts with neighboring transitions. In O supergiants two weak Fe IV lines, that overlap with He I λ304, influence the strength
of He I singlet transitions in the optical and accurate wavelengths (and oscillator strengths) of these Fe IV lines are crucial for
understanding the He I singlet transitions [1].
3 Strictly speaking, a radiation temperature is only well defined if the radiation field is Planckian. However, astronomers often use
color temperatures, defined by fitting a scaled blackbody ratio to the flux at two wavelengths, to characterize the nature of the
radiation field in some pass band. In nLTE, astronomers may also use the excitation temperature to characterize the excitation or
ionization state of a gas. In general these will not be the same as the local electron temperature, and will vary with level and
ionization stage (though possibly in a systematic way).
4 The Sun’s atmosphere is cool and dense enough for molecules to form and 50 molecular species have been identified [8]. In
the solar spectrum, spectral features due, for example, to CO, SiO, H2 , OH, CH, C2 , and CN, have been identified [8,9]. Dust
formation in red giants and supergiants is common but not well understood [10], and in some cases may be associated with
non-equilibrium chemistry [11].
5 Throughout the article we neglect full shells when providing the electron configuration. We use the principal quantum number
(n), orbital angular momentum number (l = 0, . . . , n − 1 = s, p, d, f, g . . . ), and spin (±1/2) to describe the state of an electron.
Thus, 2p indicates an electron with n = 2 and l = 1. LS-coupling (in which the orbital angular momenta are coupled and the
electron spins are coupled) is used to provide the term designation. A term designation has the format 2S+1 Lx where S is the sum
of the (valence) electron spins, L is the total orbital angular momentum, and “o” is used to indicate that the arithmetic sum of the
electron orbital angular momenta is odd (o) or even (in which case e is omitted by convention). An excellent primer on atomic
spectroscopy is provided by [63] .
6 WR stars are a class of massive stars that evolved from O stars (stars with initial masses & 15 M ). They are experiencing mass
loss via a stellar wind (induced by radiation pressure acting through bound–bound transitions) with a mass-loss rate typically in
excess of 10−5 M yr−1 and a terminal wind speed of ∼1000 to 3500 km s−1 [64,65]. In many WR stars the wind is sufficiently
Atoms 2023, 11, 54 23 of 27
dense that the entire spectrum we observe originates in the wind—the hydrostatic core of the star is not seen. There are two main
WR classes: WN stars exhibit N and He (and sometimes H) emission lines, and exhibit enhanced N and He at the stellar surface
due to the CNO cycle (the main H-fusion chain in massive stars). WC stars exhibit emission lines of He, C, and O, with a C
abundance comparable to that of He (e.g., [66,67]). They have lost all of their hydrogen envelope, with the enhanced C abundance
arising from the triple alpha process (3 4 He→12 C).
7 A P Cygni profile is formed when continuum radiation is absorbed and scattered by outflowing material. Outflowing gas along
the line of sight absorbs continuum radiation and scatters it out of the line of sight, producing blue-shifted absorption. Radiation
absorbed in other directions can be scattered into the line of sight and, for a spherically symmetric expanding gas, the combination
with the blue-shifted absorption will give rise to red-shifted emission.
8 The ejecta of Type Ia SNe are composed primarily of intermediate mass elements (Ca, Si) and iron group (Fe, Ni, Co) elements. In
such ejecta we may need to treat Auger ionization more rigorously since it could potentially affect the ionization state of the gas
and the thermalization of non-thermal electrons. The non-thermal electrons are initially produced via Compton scattering of
gamma-ray photons produced from decay of radioactive 56 Ni and 56 Co. In this case inner shell ionization will most likely occur
via non-thermal electrons. However the subsequent Auger ionization and fluorescence are independent of how the K-shell hole
was created.
9 From AUTOSTRUCTURE calculations made by a collaboration of researchers at Auburn University, Rollins College, the University of
Strathclyde, and other universities. Tables produced by N. R. Badnell and are available at Atomic Data from AUTOSTRUCTURE.
10 A signal processing term that refers to the distortion of data due to sampling which is too coarse. In the present case a narrow but
strong resonance could be missed in the photoionization cross-section when the frequency sampling is too coarse. Alternatively,
its influence could be artificially enhanced if it is not fully resolved.
11 The Sun is simultaneously oscillating in thousands of different vibration modes. The frequency and strength of these modes
depends on the internal structure of the Sun (e.g., the depth of the convection, the sound speed).
12 At the Eddington limit the force arising from the scattering of radiation by free electrons matches the gravitational force.
13 The strength of most emission lines in WR stars is proportional to the density squared. Thus, a clumped wind can yield the same
line strengths for a lower mass-loss rate (i.e., for a lower average density). On the other hand electron scattering line wings arise
from Thomson scattering of line photons by free electrons and hence scale with density. Thus, the strength of electron scattering
wings relative to their neighboring emission line can act as a global diagnostic of clumping. In WR stars the wings are offset to
the red from their originating transition because of the large outflow velocities.
14 In this process a strong transition (typically in the UV) absorbs continuum photons, a process whose efficiency is enhanced by the
velocity field which allows the UV transitions to intercept more continuum radiation. In many cases the absorbed photons will
typically be re-emitted in the same transition. However, in some cases the upper levels have an alternate decay route—decay via
this transition can then lead to emission in this bound–bound transition. This is also known as the Swings mechanism [109].
15 C II emission is also predicted but this is masked by blending with other lines.
16 The 11 appended to WC denotes the ionization class of the star—in this case, a spectrum dominated by C II with little evidence
for C III.
References
1. Najarro, F.; Hillier, D.J.; Puls, J.; Lanz, T.; Martins, F. On the sensitivity of He I singlet lines to the Fe IV model atom in O stars.
Astron. Astrophys. 2006, 456, 659–664. [CrossRef]
2. Nave, G.; Johansson, S. The Spectrum of Fe II. Astrophys. J. 2013, 204, 1. [CrossRef]
3. Clear, C.P.; Pickering, J.C.; Nave, G.; Uylings, P.; Raassen, T. Wavelengths and Energy Levels of Singly Ionized Nickel (Ni II)
Measured Using Fourier Transform Spectroscopy. Astrophys. J. 2022, 261, 35. [CrossRef]
4. Mihalas, D. Stellar Atmospheres, 2nd ed.; W. H. Freeman and Company: San Francisco, CA, USA, 1978.
5. O’Dell, C.R.; McCullough, P.R.; Meixner, M. Unraveling the Helix Nebula: Its Structure and Knots. Astrophys. J. 2004,
128, 2339–2356. [CrossRef]
6. Benedict, G.F.; McArthur, B.E.; Napiwotzki, R.; Harrison, T.E.; Harris, H.C.; Nelan, E.; Bond, H.E.; Patterson, R.J.; Ciardullo, R.
Astrometry with the Hubble Space Telescope: Trigonometric Parallaxes of Planetary Nebula Nuclei NGC 6853, NGC 7293, Abell
31, and DeHt 5. Astrophys. J. 2009, 138, 1969–1984. [CrossRef]
7. Meaburn, J.; López, J.A.; Richer, M.G. Optical line profiles of the Helix planetary nebula (NGC 7293) to large radii. Mon. Not. R.
Astron. Soc. 2008, 384, 497–503. [CrossRef]
8. Jørgensen, U.G. Molecules in Stellar and Star-Like Atmospheres. In Stellar Atmosphere Modeling; Astronomical Society of the
Pacific Conference Series; Hubeny, I., Mihalas, D., Werner, K., Eds.; Astronomical Society of the Pacific: San Francisco, CA, USA,
2003 ; Volume 288, p. 303.
9. Grevesse, N.; Sauval, A.J. Molecules in the Sun and Molecular Data. In IAU Colloq. 146: Molecules in the Stellar Environment;
Jorgensen, U.G., Ed.; Springer: Berlin/Heidelberg, Germany, 1994; Volume 428, p. 196. [CrossRef]
Atoms 2023, 11, 54 24 of 27
10. Cherchneff, I.; Sarangi, A. New Insights on What, Where, and How Dust Forms in Evolved Stars. In The B[e] Phenomenon: Forty
Years of Studies; Astronomical Society of the Pacific Conference Series; Miroshnichenko, A., Zharikov, S., Korčáková, D., Wolf, M.,
Eds.; Astronomical Society of the Pacific: San Francisco, CA, USA, 2017; Volume 508, p. 57.
11. Gobrecht, D.; Cherchneff, I.; Sarangi, A.; Plane, J.M.C.; Bromley, S.T. Dust formation in the oxygen-rich AGB star IK Tauri. Astron.
Astrophys. 2016, 585, A6. [CrossRef]
12. Hummer, D.G.; Mihalas, D. The equation of state for stellar envelopes. I—An occupation probability formalism for the truncation
of internal partition functions. Astrophys. J. 1988, 331, 794–814. [CrossRef]
13. Mihalas, D.; Dappen, W.; Hummer, D.G. The equation of state for stellar envelopes. II—Algorithm and selected results. Astrophys.
J. 1988, 331, 815–825. [CrossRef]
14. Daeppen, W.; Mihalas, D.; Hummer, D.G.; Mihalas, B.W. The equation of state for stellar envelopes. III—Thermodynamic
quantities. Astrophys. J. 1988, 332, 261–270. [CrossRef]
15. Hubeny, I.; Lanz, T. Non-LTE line-blanketed model atmospheres of hot stars. 1: Hybrid complete linearization/accelerated
lambda iteration method. Astrophys. J. 1995, 439, 875–904. [CrossRef]
16. Hillier, D.J.; Miller, D.L. The Treatment of Non-LTE Line Blanketing in Spherically Expanding Outflows. Astrophys. J. 1998,
496, 407–427. [CrossRef]
17. Kurucz, R.; Bell, B. Atomic Line Data. In Atomic Line Data (R.L. Kurucz B. Bell) Kurucz CD-ROM No. 23.; Smithsonian Astrophysical
Observatory: Cambridge, MA, USA, 1995; Volume 23.
18. Kurucz, R.L. Including All the Lines. Am. Inst. Phys. Conf. 2009, 1171, 43–51. [CrossRef]
19. Seaton, M.J. Atomic data for opacity calculations. I—General description. J. Phys. B At. Mol. Phys. 1987, 20, 6363–6378. [CrossRef]
20. Hummer, D.G.; Berrington, K.A.; Eissner, W.; Pradhan, A.K.; Saraph, H.E.; Tully, J.A. Atomic data from the IRON Project. 1:
Goals and methods. Astron. Astrophys. 1993, 279, 298–309.
21. Nahar, S. Database NORAD-Atomic-Data for Atomic Processes in Plasma. Atoms 2020, 8, 68. [CrossRef]
22. Cunto, W.; Mendoza, C.; Ochsenbein, F.; Zeippen, C.J. Topbase at the CDS. Astron. Astrophys. 1993, 275, L5.
23. Mendoza, C. TOPbase/TIPbase. In Atomic and Molecular Data and Their Applications, ICAMDATA; American Institute of Physics
Conference Series; Berrington, K.A., Bell, K.L., Eds.; American Institute of Physics: Melville, NY, USA, 2000; Volume 543,
pp. 313–315. [CrossRef]
24. Rybicki, G.B.; Lightman, A.P. Radiative Processes in Astrophysics; John Wiley & Sons: Hoboken, NJ, USA, 1979.
25. Osterbrock, D.E.; Ferland, G.J. Astrophysics of Gaseous Nebulae and Active Galactic Nuclei; University Science Books: Mill Valley, CA,
USA, 2006.
26. Pradhan, A.K.; Nahar, S.N. Atomic Astrophysics and Spectroscopy; Cambridge University Press: New York, NY, USA, 2015.
27. Ralchenko, Y. Modern Methods in Collisional-Radiative Modeling of Plasmas; Springer International Publishing: Cham, Switzerland, 2016.
28. Hillier, D.J. An iterative method for the solution of the statistical and radiative equilibrium equations in expanding atmospheres.
Astron. Astrophys. 1990, 231, 116–124.
29. Castor, J.I.; Abbott, D.C.; Klein, R.I. Radiation-driven winds in Of stars. Astrophys. J. 1975, 195, 157–174. [CrossRef]
30. Pauldrach, A.; Puls, J.; Kudritzki, R.P. Radiation-driven winds of hot luminous stars—Improvements of the theory and first
results. Astron. Astrophys. 1986, 164, 86–100.
31. Sundqvist, J.O.; Björklund, R.; Puls, J.; Najarro, F. New predictions for radiation-driven, steady-state mass-loss and wind-
momentum from hot, massive stars. I. Method and first results. Astron. Astrophys. 2019, 632, A126. [CrossRef]
32. Martins, F.; Schaerer, D.; Hillier, D.J. On the effective temperature scale of O stars. Astron. Astrophys. 2002, 382, 999–1004.
[CrossRef]
33. Crowther, P.A.; Hillier, D.J.; Evans, C.J.; Fullerton, A.W.; De Marco, O.; Willis, A.J. Revised Stellar Temperatures for Magellanic
Cloud O Supergiants from Far Ultraviolet Spectroscopic Explorer and Very Large Telescope UV-Visual Echelle Spectrograph
Spectroscopy. Astrophys. J. 2002, 579, 774–799. [CrossRef]
34. Bouret, J.C.; Lanz, T.; Hillier, D.J.; Heap, S.R.; Hubeny, I.; Lennon, D.J.; Smith, L.J.; Evans, C.J. Quantitative Spectroscopy of O
Stars at Low Metallicity: O Dwarfs in NGC 346. Astrophys. J. 2003, 595, 1182–1205. [CrossRef]
35. Hillier, D.J.; Lanz, T.; Heap, S.R.; Hubeny, I.; Smith, L.J.; Evans, C.J.; Lennon, D.J.; Bouret, J.C. A Tale of Two Stars: The Extreme
O7 Iaf+ Supergiant AV 83 and the OC7.5 III((f)) star AV 69. Astrophys. J. 2003, 588, 1039–1063. [CrossRef]
36. Bouret, J.C.; Hillier, D.J.; Lanz, T.; Fullerton, A.W. Properties of Galactic early-type O-supergiants. A combined FUV-UV and
optical analysis. Astron. Astrophys. 2012, 544, A67. [CrossRef]
37. Hillier, D.J.; Aadland, E.; Massey, P.; Morrell, N. BAT99-9—a WC4 Wolf-Rayet star with nitrogen emission: Evidence for binary
evolution? Mon. Not. R. Astron. Soc. 2021, 503, 2726–2732. [CrossRef]
38. Aadland, E.; Massey, P.; Hillier, D.J.; Morrell, N.I.; Neugent, K.F.; Eldridge, J.J. WO-type Wolf-Rayet Stars: The Last Hurrah of
Massive Star Evolution. Astrophys. J. 2022, 931, 157. [CrossRef]
39. Najarro, F. Spectroscopy of P Cygni. In P Cygni 2000: 400 Years of Progress; Astronomical Society of the Pacific Conference Series;
de Groot, M., Sterken, C., Eds.; Astronomical Society of the Pacific: San Francisco, CA, USA, 2001; Volume 233, p. 133.
40. Groh, J.H.; Hillier, D.J.; Damineli, A.; Whitelock, P.A.; Marang, F.; Rossi, C. On the Nature of the Prototype Luminous Blue
Variable Ag Carinae. I. Fundamental Parameters During Visual Minimum Phases and Changes in the Bolometric Luminosity
During the S-Dor Cycle. Astrophys. J. 2009, 698, 1698–1720. [CrossRef]
Atoms 2023, 11, 54 25 of 27
41. Groh, J.H.; Hillier, D.J.; Damineli, A. On the Nature of the Prototype Luminous Blue Variable AG Carinae. II. Witnessing a
Massive Star Evolving Close to the Eddington and Bistability Limits. Astrophys. J. 2011, 736, 46. [CrossRef]
42. Puebla, R.E.; Hillier, D.J.; Zsargó, J.; Cohen, D.H.; Leutenegger, M.A. X-ray, UV and optical analysis of supergiants: e Ori. Mon.
Not. R. Astron. Soc. 2016, 456, 2907–2936. [CrossRef]
43. Herald, J.E.; Bianchi, L. Far-Ultraviolet Spectroscopic Analyses of Four Central Stars of Planetary Nebulae. Astrophys. J. 2004,
609, 378–391. [CrossRef]
44. Hillier, D.J.; Dessart, L. Time-dependent radiative transfer calculations for supernovae. Mon. Not. R. Astron. Soc. 2012,
424, 252–271. [CrossRef]
45. Wilk, K.D.; Hillier, D.J.; Dessart, L. Understanding nebular spectra of Type Ia supernovae. Mon. Not. R. Astron. Soc. 2020,
494, 2221–2235. [CrossRef]
46. Dessart, L.; Hillier, D.J. Radiative-transfer modeling of nebular-phase type II supernovae. Dependencies on progenitor and
explosion properties. Astron. Astrophys. 2020, 642, A33. [CrossRef]
47. Dessart, L.; Hillier, D.J.; Sukhbold, T.; Woosley, S.E.; Janka, H.T. The explosion of 9-29 M stars as Type II supernovae: Results
from radiative-transfer modeling at one year after explosion. Astron. Astrophys. 2021, 652, A64. [CrossRef]
48. Dessart, L.; Hillier, D.J.; Sukhbold, T.; Woosley, S.E.; Janka, H.T. Nebular phase properties of supernova Ibc from He-star
explosions. Astron. Astrophys. 2021, 656, A61. [CrossRef]
49. Hubeny, I.; Hummer, D.G.; Lanz, T. NLTE model stellar atmospheres with line blanketing near the series limits. Astron. Astrophys.
1994, 282, 151–167.
50. Christensen-Dalsgaard, J. Solar structure and evolution. Living Rev. Sol. Phys. 2021, 18, 2. [CrossRef]
51. Seaton, M.J.; Yan, Y.; Mihalas, D.; Pradhan, A.K. Opacities for stellar envelopes. Mon. Not. R. Astron. Soc. 1994, 266, 805.
[CrossRef]
52. Iglesias, C.A.; Rogers, F.J. Updated Opal Opacities. Astrophys. J. 1996, 464, 943. [CrossRef]
53. Colgan, J.; Kilcrease, D.P.; Magee, N.H.; Sherrill, M.E.; Abdallah, J., J.; Hakel, P.; Fontes, C.J.; Guzik, J.A.; Mussack, K.A. A New
Generation of Los Alamos Opacity Tables. Astrophys. J. 2016, 817, 116. [CrossRef]
54. Blancard, C.; Cossé, P.; Faussurier, G. Solar Mixture Opacity Calculations Using Detailed Configuration and Level Accounting
Treatments. Astrophys. J. 2012, 745, 10. [CrossRef]
55. Magee, N.H.; Abdallah, J.; Clark, R.E.H.; Cohen, J.S.; Collins, L.A.; Csanak, G.; Fontes, C.J.; Gauger, A.; Keady, J.J.; Kilcrease, D.P.; et al.
Atomic Structure Calculations and New Los Alamos Astrophysical Opacities. In Astrophysical Applications of Powerful New Databases;
Astronomical Society of the Pacific Conference Series; Adelman, S.J., Wiese, W.L., Eds.; Astronomical Society of the Pacific: San Francisco,
USA, 1995 ; Volume 78, p. 51.
56. Badnell, N.R.; Ballance, C.P.; Griffin, D.C.; O’Mullane, M. Dielectronic recombination of W20+ (4d10 4f8 ): Addressing the half-open
f shell. Phys. Rev. A 2012, 85, 052716. [CrossRef]
57. Metzger, B.D.; Martínez-Pinedo, G.; Darbha, S.; Quataert, E.; Arcones, A.; Kasen, D.; Thomas, R.; Nugent, P.; Panov, I.V.; Zinner,
N.T. Electromagnetic counterparts of compact object mergers powered by the radioactive decay of r-process nuclei. Mon. Not. R.
Astron. Soc. 2010, 406, 2650–2662. [CrossRef]
58. Kasen, D.; Badnell, N.R.; Barnes, J. Opacities and Spectra of the r-process Ejecta from Neutron Star Mergers. Astrophys. J. 2013,
774, 25. [CrossRef]
59. Barnes, J.; Kasen, D. Effect of a High Opacity on the Light Curves of Radioactively Powered Transients from Compact Object
Mergers. Astrophys. J. 2013, 775, 18. [CrossRef]
60. Fontes, C.J.; Fryer, C.L.; Hungerford, A.L.; Wollaeger, R.T.; Korobkin, O. A line-binned treatment of opacities for the spectra and
light curves from neutron star mergers. Mon. Not. R. Astron. Soc. 2020, 493, 4143–4171. [CrossRef]
61. Tanaka, M.; Kato, D.; Gaigalas, G.; Kawaguchi, K. Systematic opacity calculations for kilonovae. Mon. Not. R. Astron. Soc. 2020,
496, 1369–1392. [CrossRef]
62. Fontes, C.J.; Fryer, C.L.; Wollaeger, R.T.; Mumpower, M.R.; Sprouse, T.M. Actinide opacities for modelling the spectra and light
curves of kilonovae. Mon. Not. R. Astron. Soc. 2023, 519, 2862–2878. [CrossRef]
63. Martin, W.C.; Wise, W.L. Atomic Spectroscopy: An Introduction. 2016. Available online: https://www.nist.gov/system/files/
documents/2016/10/03/atspec.pdf (accessed on 1 February 2023).
64. Hillier, D. Wolf-Rayet Stars. In Encyclopedia of Astronomy and Astrophysics; Murdin, P., Ed.; Institute of Physics Publishing: Bristol,
UK, 2000; p. 1895. [CrossRef]
65. Crowther, P.A. Physical Properties of Wolf-Rayet Stars. Annu. Rev. Astron. Astrophys. 2007, 45, 177–219. [CrossRef]
66. Sander, A.; Hamann, W.R.; Todt, H. The Galactic WC stars. Stellar parameters from spectral analyses indicate a new evolutionary
sequence. Astron. Astrophys. 2012, 540, A144. [CrossRef]
67. Aadland, E.; Massey, P.; Hillier, D.J.; Morrell, N. The Physical Parameters of Four WC-type Wolf-Rayet Stars in the Large
Magellanic Cloud: Evidence of Evolution. Astrophys. J. 2022, 924, 44. [CrossRef]
68. Weisheit, J.C. X-Ray Ionization Cross-Sections and Ionization Equilibrium Equations Modified by Auger Transitions. Astrophys. J.
1974, 190, 735–740. [CrossRef]
69. Kaastra, J.S.; Mewe, R. X-ray emission from thin plasmas. I—Multiple Auger ionisation and fluorescence processes for Be to Zn.
Astron. Astrophys. 1993, 97, 443–482.
70. McGuire, E.J. K-Shell Auger Transition Rates and Fluorescence Yields for Elements Be-Ar. Phys. Rev. 1969, 185, 1–6. [CrossRef]
Atoms 2023, 11, 54 26 of 27
71. McGuire, E.J. K-Shell Auger Transition Rates and Fluorescence Yields for Elements Ar-Xe. Phys. Rev. A 1970, 2, 273–278.
[CrossRef]
72. Bambynek, W.; Crasemann, B.; Fink, R.W.; Freund, H.U.; Mark, H.; Swift, C.D.; Price, R.E.; Rao, P.V. X-Ray Fluorescence Yields, Auger,
and Coster-Kronig Transition Probabilities. Rev. Mod. Phys. 1972, 44, 716–813. [CrossRef]
73. Owocki, S.P.; Castor, J.I.; Rybicki, G.B. Time-dependent models of radiatively driven stellar winds. I—Nonlinear evolution of
instabilities for a pure absorption model. Astrophys. J. 1988, 335, 914–930. [CrossRef]
74. Feldmeier, A. Time-dependent structure and energy transfer in hot star winds. Astron. Astrophys. 1995, 299, 523.
75. Sundqvist, J.O.; Owocki, S.P.; Puls, J. 2D wind clumping in hot, massive stars from hydrodynamical line-driven instability
simulations using a pseudo-planar approach. Astron. Astrophys. 2018, 611, A17. [CrossRef]
76. Chlebowski, T. X-ray emission from O-type stars—Parameters which affect it. Astrophys. J. 1989, 342, 1091–1107. [CrossRef]
77. Berghoefer, T.W.; Schmitt, J.H.M.M.; Cassinelli, J.P. The ROSAT all-sky survey catalogue of optically bright OB-type stars. Astron.
Astrophys. 1996, 118, 481–494. [CrossRef]
78. Cassinelli, J.P.; Olson, G.L. The effects of coronal regions on the X-ray flux and ionization conditions in the winds of OB
supergiants and Of stars. Astrophys. J. 1979, 229, 304–317. [CrossRef]
79. Pauldrach, A.W.A.; Hoffmann, T.L.; Lennon, M. Radiation-driven winds of hot luminous stars. XIII. A description of NLTE line
blocking and blanketing towards realistic models for expanding atmospheres. Astron. Astrophys. 2001, 375, 161–195. [CrossRef]
80. Zsargó, J.; Hillier, D.J.; Bouret, J.C.; Lanz, T.; Leutenegger, M.A.; Cohen, D.H. On the Importance of the Interclump Medium for
Superionization: O VI Formation in the Wind of ζ Puppis. Astrophys. J. 2008, 685, L149–L152. [CrossRef]
81. Nussbaumer, H.; Storey, P.J. Dielectronic recombination at low temperatures. Astron. Astrophys. 1983, 126, 75–79.
82. Nussbaumer, H.; Storey, P.J. Dielectronic recombination at low temperatures. II Recombination coefficients for lines of C, N, O.
Astron. Astrophys. 1984, 56, 293–312.
83. Aldrovandi, S.M.V.; Pequignot, D. Radiative and Dielectronic Recombination Coefficients for Complex Ions. Astron. Astrophys.
1973, 25, 137.
84. Sochi, T.; Storey, P.J. Dielectronic recombination lines of C+ . At. Data Nucl. Data Tables 2013, 99, 633–650. [CrossRef]
85. Burgess, A. Dielectronic recombination in the corona. Ann. D’Astrophysique 1965, 28, 774. [CrossRef]
86. Burgess, A. A General Formula for the Estimation of Dielectronic Recombination Co-Efficients in Low-Density Plasmas. Astrophys.
J. 1965, 141, 1588–1590. [CrossRef]
87. Burgess, A. Delectronic Recombination and the Temperature of the Solar Corona. Astrophys. J. 1964, 139, 776–780. [CrossRef]
88. Davidson, K. Dielectronic recombination and abundances near quasars. Astrophys. J. 1975, 195, 285–291. [CrossRef]
89. Badnell, N.R.; Pindzola, M.S.; Dickson, W.J.; Summers, H.P.; Griffin, D.C.; Lang, J. Electric Field Effects on Dielectronic
Recombination in a Collisional-Radiative Model. Astrophys. J. 1993, 407, L91. [CrossRef]
90. Nikolić, D.; Gorczyca, T.W.; Korista, K.T.; Ferland, G.J.; Badnell, N.R. Suppression of Dielectronic Recombination due to Finite
Density Effects. Astrophys. J. 2013, 768, 82. [CrossRef]
91. Hillier, D.J. WC stars—Hot stars with cold winds. Astrophys. J. 1989, 347, 392–408. [CrossRef]
92. Bowen, I.S. The Excitation of the Permitted O III Nebular Lines. Publ. Astron. Soc. Pac. 1934, 46, 146–148. [CrossRef]
93. Fang, X.; Liu, X.W. Very deep spectroscopy of the bright Saturn nebula NGC 7009—II. Analysis of the rich optical recombination
spectrum. Mon. Not. R. Astron. Soc. 2013, 429, 2791–2851. [CrossRef]
94. Peimbert, M.; Peimbert, A.; Delgado-Inglada, G. Nebular Spectroscopy: A Guide on Hii Regions and Planetary Nebulae. Publ.
Astron. Soc. Pac. 2017, 129, 082001. [CrossRef]
95. Hillier, D.J. An empirical model for the Wolf-Rayet star HD 50896. Astrophys. J. 1987, 63, 965–981. [CrossRef]
96. Field, G.B.; Steigman, G. Charge Transfer and Ionization Equilibrium in the Interstellar Medium. Astrophys. J. 1971, 166, 59.
[CrossRef]
97. Williams, R.E. The ionization structure of planetary nebulae–X. The contribution of charge exchange and optically thick
condensations to [O I] radiation. Mon. Not. R. Astron. Soc. 1973, 164, 111. [CrossRef]
98. Stancil, P.C.; Schultz, D.R.; Kimura, M.; Gu, J.P.; Hirsch, G.; Buenker, R.J. Charge transfer in collisions of O+ with H and H+ with
O. Astron. Astrophys. 1999, 140, 225–234. [CrossRef]
99. Nagayama, T.; Bailey, J.E.; Loisel, G.P.; Dunham, G.S.; Rochau, G.A.; Blancard, C.; Colgan, J.; Cossé, P.; Faussurier, G.; Fontes, C.J.;
et al. Systematic Study of L -Shell Opacity at Stellar Interior Temperatures. Phys. Rev. Lett. 2019, 122, 235001. [CrossRef]
100. Bailey, J.E.; Nagayama, T.; Loisel, G.P.; Rochau, G.A.; Blancard, C.; Colgan, J.; Cosse, P.; Faussurier, G.; Fontes, C.J.; Gilleron, F.; et
al. A higher-than-predicted measurement of iron opacity at solar interior temperatures. Nature 2015, 517, 56–59. [CrossRef]
101. Eversberg, T.; Lepine, S.; Moffat, A.F.J. Outmoving Clumps in the Wind of the Hot O Supergiant zeta Puppis. Astrophys. J. 1998,
494, 799–805. [CrossRef]
102. Lépine, S.; Moffat, A.F.J. Direct Spectroscopic Observations of Clumping in O-Star Winds. Astrophys. J. 2008, 136, 548–553.
[CrossRef]
103. Hillier, D.J. The effects of electron scattering and wind clumping for early emission line stars. Astron. Astrophys. 1991,
247, 455–468.
104. Hillier, D.J.; Miller, D.L. Constraints on the Evolution of Massive Stars through Spectral Analysis. I. The WC5 Star HD 165763.
Astrophys. J. 1999, 519, 354–371. [CrossRef]
Atoms 2023, 11, 54 27 of 27
105. Conti, P.S.; Alschuler, W.R. Spectroscopic Studies of O-Type Stars. I. Classification and Absolute Magnitudes. Astrophys. J. 1971,
170, 325. [CrossRef]
106. Walborn, N.R. Some Spectroscopic Characteristics of the OB Stars: An Investigation of the Space Distribution of Certain OB Stars
and the Reference Frame of the Classification. Astrophys. J. 1971, 23, 257. [CrossRef]
107. Mihalas, D.; Hummer, D.G.; Conti, P.S. On the N III λλ4640, 4097 lines is Of stars. Astrophys. J. 1972, 175, L99–L104. [CrossRef]
108. Rivero González, J.G.; Puls, J.; Najarro, F. Nitrogen line spectroscopy of O-stars. I. Nitrogen III emission line formation revisited.
Astron. Astrophys. 2011, 536, A58. [CrossRef]
109. Swings, P. Anomalies in the earliest spectral types. Ann. D’Astrophysique 1948, 11, 228–246.
110. Hillier, D.J. The formation of nitrogen and carbon emission lines in HD 50896 (WN5). Astrophys. J. 1988, 327, 822–839. [CrossRef]
111. Tully, J.A.; Seaton, M.J.; Berrington, K.A. Atomic data for opacity calculations. XIV—The beryllium sequence. J. Phys. B At. Mol.
Phys. 1990, 23, 3811–3837. [CrossRef]
112. Margon, B.; Manea, C.; Williams, R.; Bond, H.E.; Prochaska, J.X.; Szymański, M.K.; Morrell, N. Discovery of a Rare Late-type,
Low-mass Wolf-Rayet Star in the LMC. Astrophys. J. 2020, 888, 54. [CrossRef]
113. Williams, R.; Manea, C.; Margon, B.; Morrell, N. Line Identification and Excitation of Autoionizing States in a Late-type, Low-mass
Wolf-Rayet Star. Astrophys. J. 2021, 906, 31. [CrossRef]
114. Leuenhagen, U.; Hamann, W.R.; Jeffery, C.S. Spectral analyses of late-type WC central stars of planetary nebulae. Astron.
Astrophys. 1996, 312, 167–185.
115. De Marco, O.; Barlow, M.J.; Storey, P.J. The WC10 central stars CPD-56 deg8032 and He 2-113—I. Distances and nebular
parameters. Mon. Not. R. Astron. Soc. 1997, 292, 86–104. [CrossRef]
116. De Marco, O.; Crowther, P.A. The WC10 central stars CPD-56 deg8032 and He2-113—II. Model analysis and comparison with
nebular properties. Mon. Not. R. Astron. Soc. 1998, 296, 419–429. [CrossRef]
117. De Marco, O.; Storey, P.J.; Barlow, M.J. The WC10 central stars CPD-56 deg8032 and He2-113—III. Wind electron temperatures
and abundances. Mon. Not. R. Astron. Soc. 1998, 297, 999–1014. [CrossRef]
118. Barlow, M.J.; Storey, P.J. The Wind Temperature and C/He and O/He Ratios of the WC10 Central Star CPD560 8032. In Planetary
Nebulae; IAU Symposium; Weinberger, R., Acker, A., Eds.; Kluwer Academic: Dordrecht, Germany, 1993 ; Volume 155, p. 92.
119. Storey, P.J.; Sochi, T. Electron temperatures and free-electron energy distributions of nebulae from C II dielectronic recombination
lines. Mon. Not. R. Astron. Soc. 2013, 430, 599–610. [CrossRef]
120. Foglizzo, T. Explosion Physics of Core-Collapse Supernovae. In Handbook of Supernovae; Alsabti, A.W., Murdin, P., Eds.; Springer
International Publishing: Berlin/Heidelberg, Germany, 2017; p. 1053. [CrossRef]
121. Hoeflich, P. Explosion Physics of Thermonuclear Supernovae and Their Signatures. In Handbook of Supernovae; Alsabti, A.W.,
Murdin, P., Eds.; Springer International Publishing: Berlin/Heidelberg, Germany, 2017; p. 1151. [CrossRef]
122. Shingles, L.J.; Flörs, A.; Sim, S.A.; Collins, C.E.; Röpke, F.K.; Seitenzahl, I.R.; Shen, K.J. Modelling the ionization state of Type Ia
supernovae in the nebular phase. Mon. Not. R. Astron. Soc. 2022, 512, 6150–6163. [CrossRef]
123. Shull, J.M.; van Steenberg, M. The ionization equilibrium of astrophysically abundant elements. Astrophys. J. 1982, 48, 95–107.
[CrossRef]
124. Palacios, A.; Gebran, M.; Josselin, E.; Martins, F.; Plez, B.; Belmas, M.; Lèbre, A. POLLUX: A database of synthetic stellar spectra.
Astron. Astrophys. 2010, 516, A13. [CrossRef]
125. Zsargó, J.; Rosa Fierro-Santillán, C.; Klapp, J.; Arrieta, A.; Arias, L.; Mendoza Valencia, J.; Sigalotti, L.D.G. Creating and using
large grids of pre-calculated model atmospheres for rapid analysis of stellar spectra. arXiv 2021. [CrossRef]
126. Zsargó, J.; Fierro-Santillán, C.R.; Klapp, J.; Arrieta, A.; Arias, L.; Valencia, J.M.; Sigalotti, L.D.G.; Hareter, M.; Puebla, R.E. Creating
and using large grids of precalculated model atmospheres for a rapid analysis of stellar spectra. Astron. Astrophys. 2020, 643, A88.
[CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.