Next Article in Journal
Recent Advances in Cold Atmospheric Pressure Plasma for E. coli Decontamination in Food: A Review
Previous Article in Journal
A Theoretical Study of the Ionization States and Electrical Conductivity of Tantalum Plasma
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Strong, Weak and Merging Lines in Atomic Spectra

by
Jean-Christophe Pain
1,2
1
CEA, DAM, DIF, F-91297 Arpajon, France
2
Université Paris-Saclay, CEA, Laboratoire Matière en Conditions Extrêmes, F-91680 Bruyères-le-Châtel, France
Plasma 2025, 8(2), 17; https://doi.org/10.3390/plasma8020017
Submission received: 25 March 2025 / Revised: 15 April 2025 / Accepted: 24 April 2025 / Published: 29 April 2025

Abstract

:
We present analytical estimates for the maximum line strength in a transition array, as well as for the numbers of strong and weak lines. For that purpose, two main assumptions are used as concerns the line strength distribution. The first one, due to Porter and Thomas, is more suitable for J J sets, where J is the total atomic angular momentum, and the second one, based on a decreasing-exponential modeling of the line-amplitude distribution, is more relevant for an entire transition array. We also review the different approximations of overlapping and blanketing (band model), insisting on the computation and properties of the Elsasser function. We compare different approximations of the Ladenburg–Reiche function giving the equivalent width of an ensemble of lines in a frequency bin and discuss the possibility of using statistical indicators, such as the Chernoff bound or the Gini coefficient (initially introduced in economics for the measurement of income inequality), in the statistical characterization of transition arrays.

Graphical Abstract

1. Introduction

Knowing the number of lines in a transition array (ensemble of lines between two electronic configurations differing by an electric-dipole-allowed one-electron jump) is important for calculating the absorption and emission spectra of high-temperature plasmas, corresponding typically to k B T > 5 eV ( k B being the Boltzmann constant). Of course, as can be seen in Table 1 for different transition arrays, the number of lines depends on the angular-momentum coupling scheme; throughout this paper, we consider intermediate coupling. Indeed, when these lines are sufficiently numerous, it is reasonable to represent them by continuous, smooth structures, modeled statistically by simple functions, whose characteristics (mean value, variance) can be calculated exactly using Racah’s angular momentum-coupling techniques [1,2,3,4,5,6]. Although such approaches (UTA: unresolved transition arrays and SOSA: spin–orbit–split arrays) have been proven to be very powerful, it is important to be aware of more detailed properties of transition arrays, such as high-order moments of the distribution [7], value of the maximal strength, number of weak lines (i.e., which strength is smaller than a given small threshold), strong lines (i.e., which strength is higher than a high fraction of the maximal strength), etc.
Global properties of spectra, as well as regularities or trends, are important for checking atomic databases [8,9,10,11], as well as opacity tables [12,13]. They can help in detecting anomalies or opening the way to new discoveries.
In the field of quantitative analytical spectroscopy, the foundational experimental data generally comprise measurements of total emission or absorption (or transmission) coefficients as functions of photon wavelength or energy. While the resulting spectrum may appear straightforward in cases involving a single electronic transition—thus producing only one spectral line—such simplicity is the exception rather than the norm. More commonly, multiple transitions occur simultaneously, giving rise to complex spectra with overlapping lines. The accurate quantitative interpretation of extinction coefficients within these overlapping transmission spectra represents a significant analytical challenge, which has been the focus of considerable research efforts by Elsasser [14] and Benedict et al. [15], for instance. Except for the (ideal) case of an infinite series of equally spaced, identical lines, the quantitative effect of line overlap requires the use of approximation methods. In addition to the spacing between neighboring lines, another easily accessible parameter of a transmission spectrum is the area between the spectral line and its baseline. Investigations of the effect of line overlap on this particular parameter have been limited, but for the special case of a symmetric doublet of overlapping Lorentzian lines, Bykov and Hien found that the interference contribution to spectral area can be approximated using zero- and first-order Bessel functions of the second kind with an imaginary argument [16].
In Section 2, we propose different methods to predict the maximum line strength, based on two different assumptions for the distribution of the amplitudes: the Porter–Thomas law, valid for J J sets, and a decreasing exponential (or Poisson distribution) over a whole transition array. In Section 3, different estimates for the number of strong and weak lines are obtained, in the context of the two hypotheses mentioned above. In Section 4, we review the main historical models of line opacity, both in the non-overlapping and overlapping cases. In the same section, the equivalent width is discussed, together with the Ladenburg–Reiche function. Different approximations of the latter are compared. Promising statistical tools are briefly mentioned in the two appendices, in view of their potential application to the modeling of complex atomic spectra.

2. Estimation of the Maximal Line Strength

2.1. Porter–Thomas Distribution

Porter and Thomas have shown that the amplitudes of the lines between all the levels of two random matrices, i.e., two Gaussian Orthogonal Ensembles (GOE), are well described by a Gaussian distribution [17,18]. In the corresponding formalism, the number of lines whose amplitude lies between a and a + d a reads as follows:
P ( a ) = L 2 π v exp a 2 2 v .
Based on the strength being the square of the amplitude, S = a 2 , we have the following:
P ( S ) d S = P ( a ) d a + P ( a ) d a = 2 P ( a ) d a
yielding a distribution of the χ 2 type [19,20]:
P ( S ) = L 2 π S S exp S 2 S ,
where L and S represent the number of lines and the mean value of the line strength in a particular ( J , J ) set, respectively. A given ( J , J ) set represents the number of lines between a collection of levels with the same total angular momentum J to levels with the same total angular momentum J . Due to the electric-dipole selection rules, we have J = J (with J = J = 0 forbidden), J = J + 1 or J = J 1 . The variance of the amplitudes v is exactly the average strength v = S (total strength of the array divided by the number of lines).
Nuclear physicists have been pioneers in analyzing the statistical properties of microscopic systems composed of a large number of particles. Initially, they studied the energy levels of nuclei and introduced the concepts of random matrices and the Gaussian Orthogonal Ensemble (GOE). The fact that the amplitudes of the lines between all the levels of two random matrices obey a Gaussian distribution was first found to describe statistics of transition strengths between chaotic states in compound nuclei and has also proven its worth for the spectral statistics of heavy open-shell atoms [21,22,23,24]. By definition, all elements of a random matrix are independent and follow two Gaussian distributions. The variance of the distribution for the diagonal elements is twice that of the off-diagonal elements. It has been demonstrated that the eigenvalues of a high-order Hamiltonian matrix follow a Gaussian distribution, and that the distribution of their nearest-neighbor spacings corresponds to the derivative of a Gaussian function with respect to the line amplitudes. Porter and Thomas established that the transition amplitudes between all levels of two random matrices follow a Gaussian distribution. In both atomic and nuclear physics, it is tempting to consider that, within a given configuration, all energy levels corresponding to the same value of total angular momentum are eigenvalues of a random matrix. This is plausible since the total angular momentum J is a well-defined quantum number, as the Hamiltonian operator commutes with J . As pointed out by Bauche [25], other exact quantum numbers are irrelevant in this context. Indeed, parity has a fixed value within each configuration, and the ( J , M J ) states are degenerate with respect to M J (projection of J along the z-axis) in the absence of external fields. The exponential term dominates in P ( S ) , except close to the origin (i.e., for very weak lines). This is due to the fact that Random Matrix Theory contains approximate symmetries, which are not sufficient to describe the vicinity of Russell–Saunders and j j couplings [19,26]. In the model proposed by Wilson et al. [27], diagonal terms are calculated in a pure coupling using Cowan’s code [28], and off-diagonal elements are populated statistically beyond the GOE according to a bi-Gaussian distribution function where elements are correlated. They observed a huge number of off-diagonal elements of a small amplitude.
Starting from the Porter–Thomas assumption that the electric-dipolar line amplitude between two ( J , J ) sets (considered as Gaussian Orthogonal Ensembles) follows a Gaussian distribution, we can resort to the extreme-value theorem of Kendall and Stuart [29] to estimate the value of S max such that, on average, strength values larger than S max are greater than 1 / 2 . This constraint reads as follows:
L 2 π v S max 1 S exp S 2 v d S = L erfc S max 2 v = 1 2 ,
where erfc = 1 erf ( z ) , with the usual error function
erf ( x ) = 2 π 0 x e t 2 d t .
Using the following asymptotic form of the complementary error function for large values of z:
erfc ( z ) = 2 π z exp t 2 d t 1 π z exp z 2 ,
and setting u = S max / ( 2 v ) , Equation (3) can be rewritten as follows:
2 L π = u exp u 2 ,
or, in other words,
2 u 2 exp 2 u 2 = 8 L 2 π ,
whose solution can be expressed using the Lambert W function 2 u 2 = W 8 L 2 π [30,31,32] (see Figure 1). Finally,
S max = v W 8 L 2 π .
The Lambert W has the following expansion [33,34,35]:
W ( x ) = ln x ln ln x + k = 0 m = 1 c k m ( ln ln x ) m ( ln x ) k + m ,
where
c k m = ( 1 ) k m ! S k + m ( k + 1 ) ,
S k + m ( k + 1 ) is the Stirling number of the first kind [36,37], defined as follows:
x ( x 1 ) ( x n + 1 ) = i = 0 n S i ( n ) x j
and for which Karanicoloff proposed an interesting expression [38]. We suggest using Equation (5) truncated at the fourth order:
W ( x ) = ln x ln ln x + ln ln x ln x + O ln ln x ln x 2 ,
since it provides a good compromise between simplicity and accuracy [39]. S max can then be approximated as follows:
S max = v ln 8 L 2 π ln ln 8 L 2 π + ln ln 8 L 2 π ln 8 L 2 π .
Bauche and Bauche–Arnoult obtained a different expression, making the additional assumption u 2 ln ( L ) after taking the logarithm of both sides of Equation (4):
ln ( 2 L ) 1 2 ln ( π ) + ln u + u 2 ,
and obtained the following [19]:
S max = v 2 ln L ln ( ln L ) ln π 4 .
The relative difference between Formulas (6) and (8) for S max / S as a function of the total number of lines L is represented in Figure 2. It is always very small (below 2%) in the considered range of variation in the number of lines L.

2.2. Decreasing Exponential Distribution for the Line Amplitudes

The assumption made by Porter and Thomas is relevant inside ( J , J ) sets of lines of a transition array [20,25,40]. Bauche, Bauche-Arnoult, and Wyart studied the distribution of the intensities of the bound–bound transitions in an atomic transition array [25]. They found that the distribution of the line amplitudes is essentially a decreasing exponential function. However, there always appears an excess of weak lines. This excess is related to coupling effects, when the physical coupling is close to an extremal (pure) coupling, and to scars of symmetries, in the opposite case. The occurrence of several types of scars makes the whole intensity array very complex. If we are interested in the maximal strength over a whole transition array of lines, we can consider that the sum of several Gaussian distributions leads to a decreasing exponential distribution for the line amplitudes [41,42,43,44]:
P ( a ) = L λ 2 exp λ a ,
where λ is defined by v = 2 / λ 2 = S = S tot / L , S tot , representing the total strength of the transition array:
S tot = J J J c , J c S J J ,
where c and c are the initial and final configurations of the transition array. A level should be denoted γ J , γ representing the ensemble of quantum numbers (possibly) required to identify a level of total angular momentum J in a unique way. However, for simplicity, we label a level by a value of the total atomic angular momentum J. Using relation (1), we obtain the following:
P ( S ) = L λ 2 S exp λ S .
The theorem of extreme values [45,46,47] enables one to define the maximal strength of a transition array as follows:
S max P ( S ) d S = L exp λ S max = 1 2 ,
leading to the simple form for S max :
S max = v 2 ln ( 2 L ) 2 .
It is worth mentioning that Bauche-Arnoult et al. proposed an alternative approach [20]. Since it is easier to calculate integrals with finite bounds for an exponential function than for a Gaussian, the authors suggested to represent, at least locally, the Gaussian by an exponential. They performed “shaping” of the exponential by the right-half of a Gaussian distribution as follows:
2 π exp x 2 d x = η exp ( λ η ) d η ,
thus defining the relation between the x and y variables. Since both sides of the latter equation are normalized to unity when integrated from 0 to , they can be used to relate x max and y max by integrating both sides from x max (resp. y max ) to infinity. This provides the following equation:
erfc ( x max ) = exp ( η y max ) .
Thus, one has the following:
S max = 1 η 2 ln erfc 1 2 2 ln ( L ) ln ( ln L ) ln π 4 1 / 2 .
Of course, the result depends on the value of η (Bauche et al. suggest to take η = 1 ). Within that approach, it is possible to combine both assumptions, first the Gaussian one (since y max is deduced from x max , stemming from the Gaussian assumption), and the exponential one (for the shape of the distribution). This might be interesting since both distributions can be viewed as limiting cases, the truth likely lying in between. However, the fact that the results depend on the (arbitrary) parameter λ is clearly a weakness of the model.
The values of expressions (6) and (9) for S max / S as a function of the total number of lines L are compared in Figure 3. The maximal strength estimated using Formulas (6), (8) and (9) is compared to the exact value for two different transition arrays: Pd VII 4d4 − 4d3 5p and Cd V 4d7 5s − 4d7 5p in Table 2. (The Slater and spin-orbit integrals used for the diagonalization of Pd VII 4d4 − 4d3 5p are given in Table 3). We can see that Formula (9) gives the best agreement with the exact value, which is due to the fact that the decreasing exponential is more relevant than the Gaussian to describe the distribution of line amplitudes inside a whole transition array. The two transition arrays, computed using Cowan’s code [28], are displayed in Figure 4 and Figure 5, respectively.

2.3. Maximal Strength Associated with Learner’s Rule

In a seminal 1982 study, Learner measured a large number of line intensities in the atomic emission spectrum of neutral iron and uncovered a striking statistical regularity: a power-law relationship governing the density of spectral lines as a function of their intensity [48]. Specifically, he demonstrated that the logarithm of the number of lines, N n , whose intensities fall within the interval [ 2 n I 1 , 2 n + 1 I 1 ] (with n being an integer), decreases linearly with n. The reference intensity I 1 is chosen such that this linear law holds consistently over nine octaves ( 1 n 9 ), based on an analysis of approximately 1500 spectral lines in the wavelength range from 290 nm to 550 nm. The number of lines is multiplied by 10 p when the size of the interval is multiplied by two. Learner’s observations imply the following ( S 1 is the line strength corresponding to intensity I 1 in the interval 2 n 1 S 1 2 n S 1 ) [48,49,50]:
log 10 2 n 1 S 1 2 n S 1 P ( S ) d S q 0 p . n ,
with P ( S ) k . S α and p > 0 . Equation (10) can be put in the following form:
log 10 2 n 1 S 1 2 n S 1 P ( S ) d S log 10 k S 1 α 2 n 1 S 1 2 n S 1 1 α log 10 k S 1 1 α 1 2 α 1 ( 1 α ) + ( 1 α ) log 10 ( 2 ) . n .
Identifying Equation (11) with Equation (10), we obtain q 0 log 10 k S 1 1 α 2 1 α 1 ( 1 α ) and p ( 1 α ) log 10 ( 2 ) . Since p 1 2 log 10 ( 2 ) (observation made by Learner), we find α = 3 2 and P ( S ) k . S 3 / 2 , where
k ( 1 α ) 10 q 0 S 1 1 α 1 2 α 1 10 q 0 S 1 2 ( 2 1 ) .
The extreme-value theorem yields the following:
S max P ( S ) d S = 2 k S max = 1 2 ,
i.e.,
S max 16 k 2 4 . 10 2 q 0 S 1 ( 2 1 ) 2 .

3. On the Number of Strong and Weak Lines

Estimating the number of weak and strong lines can be useful in order to choose the best statistical representation for an array of lines and to derive hybrid models [51] such as the Mixed-UTA approach (see Section 4) below [52]. We can decide, following Bauche et al. [19], that strong lines are such that their strength is greater than a small value S 0 (e.g., a small fraction of the maximal strength S max determined in the previous section).

3.1. Porter–Thomas Law

In the first approach (Porter–Thomas, Gaussian distribution of the amplitudes, ( J , J ) set), the number of weak lines reads, using Equation (2):
N weak = 0 S 0 L 2 π S S exp S 2 S d S
or equivalently,
N weak = 0 S 0 L 2 π v S exp S 2 v d S
giving the following based on the change in variables S = x 2 :
N weak = L 2 π v 0 S 0 exp x 2 2 v d x
and thus, setting t = x / 2 v = t / 2 S :
N weak = L erf S 0 2 S ,
or equivalently:
N strong = L 1 erf S 0 2 S .
We suggest estimating the erf function through the following expansion:
erf ( x ) = 2 π exp x 2 x + 2 3 x 3 + 4 15 x 5 + o ( x 6 e x 2 ) .
The latter formula is very accurate for the values of the argument encountered in the present work. It may be useful to have an expression in the form of a polynomial multiplied by a Gaussian (which is related to the moments of the Gaussian) for integration purposes, for instance. An even better compromise between simplicity and accuracy can be achieved via the following approximation [53]:
erf ( x ) 1 exp 1.9 x 1.3 ,
with a difference below 2.2 × 10−2. This gives the following:
N weak = L 1 exp 1.9 S 0 2 v 0.65 ,
or equivalently:
N strong = L exp 1.9 S 0 2 v 0.65 .
The approximation improves to yield a difference of less than 10−2 for x 1 . Other useful methods for approximating the error function are recalled in Appendix A.

3.2. Decreasing Exponential Distribution of the Amplitudes

In the second approach (whole transition array, decreasing exponential of the amplitudes), one obtains the following:
N strong = L N weak = L exp λ S 0 .
We have seen that v = 2 / λ 2 . On the other hand, v = S tot / L , where S tot is the total strength of the transition array and L its number of lines. This yields the following:
λ = 2 L S tot
and thus, one obtains the following:
N strong = L N weak = L exp 2 L S 0 S tot = L exp 2 S 0 S .
As can be checked in Table 4, the approximate Formulas (13) and (17) do not give accurate estimations of the numbers of lines whose strength is greater than a specific value. However, they both provide a general trend. What is more surprising in that case is that Formula (13) is closer to the exact values, although it is expected (Porter–Thomas) to be more reliable for J J sets than for a whole transition array. Actually, the disagreement can be easily understood here. Indeed, as can be see in Figure 4, the strongest line is “alone”, in the sense that its strength is much larger than the strength of the other lines. It is easily understandable, thus, that basing the estimator of the weak/strong lines on this singularity is not the most appropriate choice. It would be, in that case, much more relevant to consider the strong lines as lines having a strength larger than a fraction of the total strength of the array for instance, or having a strength larger than the average strength, etc.

3.3. Imposing a Number of Lines and Searching for the Threshold

In the computation of hot-plasma atomic spectra, one may be interested in imposing that the number of lines does not exceed a certain value. Thus, it may seem natural to keep only the strongest lines N strong . Equation (17) gives the following in the case of decreasing exponential variation in the amplitudes:
S 0 = S 2 ln L N strong 2 .
However, in the case of the Porter–Thomas distribution, as can be seen from Equation (12), one needs to invert the erf function [54,55,56]. This can be described using a series expansion:
erf 1 ( z ) = k = 0 b k 2 k + 1 π 2 z 2 k + 1
where
c k = m = 0 k 1 b m b k 1 m ( m + 1 ) ( 2 m + 1 )
and
b 0 , b 1 , b 2 , b 3 , = 1 , 1 , 7 6 , 127 90 , .
We obtain the following expansion:
erf 1 ( z ) = 1 2 π z + π 12 z 3 + 7 π 2 480 z 5 + 127 π 3 40320 z 7 + 4369 π 4 5806080 z 9 + 34807 π 5 182476800 z 11 + .
Using the simple and efficient formula (15), one obtains the following:
S 0 = 2 S 1 1.9 ln L N strong 20 13 .
An approximation (14) of the erf function (more precisely, of 1 erf ( x ) ) is illustrated in Figure 6. A comparison between Formulas (18) and (19) is displayed in Figure 7 for the quantity S 0 / S .

4. Merging Lines Together: Band Models

The modeling of several lines, and in particular, their potential overlap has been a subject of investigation since the early times of spectroscopy. The models presented in this section have been widely used in the past, but mostly in the context of molecular spectroscopy, in particular by the National Aeronautics and Space Administration; for instance, for the determination of atmospheric temperature profiles from planetary limb radiance profiles [57], or the effects of transmission models in the rotational water-vapor band on radiance calculations and constituent inferences [58]. They have been applied to hot plasma as well, but to a lesser extent. They are interesting because they provide insight about the coalescence properties of spectra, which are directly related to the so-called “porosity” of transition arrays. This is a cornerstone of opacity theory, because the modeling of radiation transport requires, at least in the diffusion approximation, the knowledge of the Rosseland mean opacity:
κ R = 0 i n f t y W R ( u ) κ ( u ) d u 1 ,
where κ ( u ) is the spectral opacity and
W R ( u ) = 15 4 π 4 u 4 e u ( 1 e u ) 2
is the Rosseland weighting function, proportional to the derivative of the Planck distribution with respect to the temperature. The Rosseland mean is thus a weighted harmonic mean of the spectral (monochromatic) opacity. Being able to quantify this “amount of merging” may be very useful in order to define coalescence criteria and thus correct the predictions of the statistical methods (UTA, …) using a coefficient accounting for the porosity. Such models are also useful to define a resulting average line full width at half maximum. However, these models are often dependent on strong assumptions. For instance, the Elsasser model [14] (see Section 4.2.1 below) assumes Lorentz line shape, equal spacing, and equal intensity for the lines.
The common assumptions of band models are as follows. We consider an infinite array of absorption lines of uniform statistical properties. Moreover, an interval of this array (containing several lines) is taken to have properties similar to those of an interval of the real interval under consideration. Finally, each interval is flawed by statistically similar intervals. This will not necessarily be true for a real band. In the following, we discuss two different cases: the non-overlapping model, in which all lines are modeled by a single profile, and the overlapping model, in which each line has its own profile.

4.1. Non-Overlapping Model

4.1.1. Lorentzian Shape

Let us define the absorptance as follows:
A ¯ = 1 d ( 1 e ρ κ ν ) d ν ,
where d is the average spacing between lines, ρ is the mass density and is the thickness of the sample ( ρ is therefore the areal mass density). Assuming a Lorentzian shape for the opacity gives the following:
κ ν = S α π 1 α 2 + ν 2 ,
and we obtain the following for the absorptance:
A ¯ = 2 π y u e u J 0 ( i u ) i J 1 ( i u ) = 2 π y L ( u ) ,
where L ( u ) is the Ladenburg–Reiche function [59,60,61] ( J 0 and J 1 are Bessel functions of the first kind of orders 0 and 1, respectively [62]). We have introduced the following reduced variables:
x = ν d , y = α d and u = S ρ 2 π α .
For strong lines u 1 , L ( u ) 2 u / π , and for weak lines, L ( u ) u .

4.1.2. Gaussian (Doppler) Shape

In the case of the Gaussian assumption, we obtain the following for the opacity:
k ν = S π Δ ν D exp ν 2 Δ ν D 2
with
x = ν Δ ν D and w = S ρ π Δ ν D ,
yielding the absorptance:
A ¯ = Δ ν D d 1 exp w e x 2 d x .
For large values of w, we have the following:
A ¯ = 2 Δ ν D d ( log w ) 1 / 2 + 0.2886 ( log w ) 1 / 2 0.1335 ( log w ) 3 / 2 + 0.0070 ( log w ) 5 / 2 . . . .
Although these two non-overlapping models (together with the Lorentzian and Gaussian assumptions) are both based on a strong approximation, they were of great use to physicists in the early stages of opacity calculations [63].

4.2. Overlapping Models

The case of overlapping lines is of course more realistic and frequent [64] in the framework of the calculation of hot plasma absorption and emission coefficients.

4.2.1. The Elsasser Model

Within the Elsasser model [14], the absorption coefficient is modeled by a sum of Lorentzian lines:
κ ν = n = α π S ( ν n d ) 2 + α 2
yielding the absorptance coefficient:
A ¯ = 1 1 2 1 2 exp 2 π u y sinh ( 2 π y ) cos ( 2 π y ) cos ( 2 π x ) d x = 1 E ( y , u ) ,
where E ( y , u ) is the so-called Elsasser function [14,65,66]:
E ( y , u ) = 1 π 0 π exp u z sinh ( z ) cosh ( z ) cos ( y ) d y .
The Elsasser function can be put in the following form [67]:
= 1 0 u z exp [ cotanh ( z ) . a ] I 0 a sinh z d a ,
where I 0 is the modified Bessel function of the first kind of order 0. The modified Bessel function of the first kind I n ( z ) can be defined by the contour integral:
I n ( z ) = 1 2 π i e ( z / 2 ) ( t + 1 / t ) t n 1 d t ,
where the contour encloses the origin and is crossed in a counterclockwise direction [68]. I n ( x ) is related to J n ( x ) as follows:
I n ( x ) = i n J n ( i x ) = e n π i / 2 J n ( x e i π / 2 ) .
When y , sinh ( 2 π y ) cosh ( 2 π y ) , this yields the following:
E ( y , u ) e 2 π y u .
In the strong-line limit ( u 1 ), neglecting y 2 in the denominator of the Lorentz profile, one has sinh ( 2 π y ) 2 π y and cosh ( 2 π y ) , leading to the following:
E ( y , u ) 1 erf ( π y 2 u ) .
An approximation for moderately large y values can also be derived, noticing that
2 π u y sinh ( 2 π y ) cosh ( 2 π y ) cos ( 2 π x ) 2 π y u 1 + 2 e 2 π y cos ( 2 π x ) .
The corresponding form is as follows:
E ( y , u ) I 0 ( 4 π y u e 2 π y ) e 2 π y u .
The evaluation of the Elsasser function is discussed in [69]. Other tabulations have been given by [70,71,72,73]. Figure 8, Figure 9, Figure 10, Figure 11, Figure 12 and Figure 13 show the comparison between the exact calculation and different approximations of the Elsasser function as a function of y for different values of u, and as a function of u for different values of y. The three approximations are given by Equations (20), (21) and (22), respectively. The third one represents a good compromise between simplicity and range of reasonable accuracy.

4.2.2. The Mayer–Goody Model

Taking into account the line strength distribution P ( S ) , the mean absorptance reads as follows [63,74,75,76]:
A ¯ = 1 exp 1 d 0 P ( S ) ( 1 e κ ν ρ ) d S d ν 0 P ( S ) d S ,
where different forms of P ( S ) were introduced in Section 2. Two distributions were studied by Mayer and Goody.
The first one assumes that P ( S ) = 1 σ e S σ . In this case, one has the following:
A ¯ = 1 exp 1 d ρ σ f ( ν , σ ) 1 + ρ σ f ( ν , σ ) d ν
with
f ( ν , σ ) = α π 1 ν 2 + α 2
yielding
A ¯ = 1 exp ρ σ α d α 2 + ρ σ α π 1 / 2 .
The second model sets P ( S ) = K S , yielding the following for the absorptance:
A ¯ = 1 exp 2 π K α d e u J 0 ( i u ) + 2 u e u J 0 ( i u ) + J 1 ( i u ) 1

4.3. Equivalent Width of a Lorentz Line

The equivalent width W of a line is the width of a rectangular line with 100% absorption that has the same absorption as the actual line [77,78,79]. Using the definition of the absorptance of a line, we obtain the following:
W = A ¯ = 1 d ( 1 e ρ κ ν ) d ν ,
which involves the Ladenburg–Reiche function (see the non-overlapping case discussed in Section 4.1.1) [59,60]:
L ( x ) = x e x J 0 ( i x ) i J 1 ( i x ) ,
where J n ( x ) is a Bessel function of the first kind. Since
J n ( i x ) = i n I n ( x ) ,
where I n ( x ) is a modified Bessel function of the first kind, one has L ( x ) f ( x ) , where
f ( x ) = x e x I 0 ( x ) + I 1 ( x )
and different approximations were proposed. One has, for instance [80]:
f ( x ) 2 x π 1 exp π x 2 1 / 2
or Varanasi (quoted by Tien in Ref. [80]):
f ( x ) = 2 x π 1 exp π x 2 1 / 2 .
The following approximation was also proposed (referred to as “Goldman 1” in Figure 14 and Figure 15) [81]:
f ( x ) x 1 + π x 2 1 / 2
derived from the function F ( x ) = x / 1 + 2 x by the transformation ( 4 π ) F ( x π / 4 ) . Finally, the last considered expression is as follows (referred to as “Goldman 2” in Figure 14 and Figure 15) [81]:
f ( x ) x 1 + π x 2 α 1 2 α with α = 5 / 4 .
A comparison of the accuracy of the four approximations, Equations (23)–(26), was already performed by Goldman [81] using the values obtained by Kaplan and Eggers [65] as exact values.
It turns out that the calculations were carried out for (approximation – exact function)/exact function. It is seen that the maximum error is about 17, 6, 8, and 1 percent for approximations (23), (24), (25) and (26), respectively. The largest errors in using Equations (23)–(25) occur for intermediate values of x, between x = 0.2 and x = 2 , while the error in using Equation (26) lies close to zero near x = 0.5. Approximation (26) is more accurate than the others and yet simple enough to manipulate mathematically. Approximation (25) is not as accurate as Equation (26), but it is convenient for a least squares fitting of a set of experimental data to the theoretical curve.
Martin, Olivares and Sotomayor [82] proposed the following approximant:
I 1 ( x ) = x cosh ( x ) ( 1 + λ 2 x 2 ) 3 / 4 p 0 + 2 π λ 3 / 2 q 1 x 2 ( 1 + q 1 x 2 )
with λ = 1 / 5 and p 0 = 1 / 2 and
q 1 = 3 4 λ 2 3 8 4 2 π λ 3 / 2 1 .
In terms of λ only, one has the following:
I 1 ( x ) = x cosh x ( 1 + λ 2 x 2 ) 3 / 4 1 2 + 2 π λ 3 / 2 3 4 λ 2 3 8 4 2 π λ 3 / 2 1 x 2 1 + 3 4 λ 2 3 8 4 2 π λ 3 / 2 1 x 2 ,
and with numerical values,
I 1 ( x ) = x cosh ( x ) 2 ( 1 + 0.04 λ 2 x 2 ) 3 / 4 1 + 0.0574403 x 2 1 + 0.40244 x 2 .
For the integral I 0 ( x ) Olivares, Martin and Valero proposed the following approximation [83]:
I 0 ( x ) = cosh ( x ) ( 1 + λ 2 x 2 ) 1 / 4 p 0 + p 1 x 2 ( 1 + q 1 x 2 )
with λ = 1 / 2 , p 0 = 1
p 1 = 3 4 λ 2 3 8 4 2 π λ 3 / 2 1 .
and
q 1 = 1 4 λ 2 4 1 2 π λ 1 / 2 .
In terms of λ only, one has the following:
I 0 ( x ) = cosh x ( 1 + λ 2 x 2 ) 1 / 4 1 + 2 π 1 λ 2 1 2 π λ 1 / 2 λ 1 / 2 x 2 4 1 + 1 λ 2 1 2 π λ 1 / 2 x 2 4
and with numerical values,
I 0 ( x ) = cosh ( x ) ( 1 + 0.25 x 2 ) 1 / 4 1 + 0.24273 x 2 1 + 0.43023 x 2 .
Figure 14 and Figure 15 show the different approximations over two different ranges.
Figure 16 shows the relative difference between the exact value of the Ladenburg–Reiche function and the approximation obtained using the approximants of the Bessel functions I 0 and I 1 published by Olivares et al. [83] and Martin et al. [82], respectively. Figure 17 displays the same quantities as Figure 14 but over a wider range. Figure 18 represents the relative difference with respect to the exact value for different approximations of the Ladenburg–Reiche function. The Tien and Varanasi formulas are given by Equations (23) and (24), respectively, and the Goldman (1 and 2) formulas are given by Equations (25) and (26), respectively. The plotted quantity corresponds to (Exact – Approximation)/Exact and is expressed in %.

4.4. General Considerations About the Merging of Lines (Blanketing)

In 2007, Abdallah et al. introduced a novel approach to spectral line binning [84], leveraging the observation that many spectral lines often appear in close proximity—particularly in cases such as intermediate coupling in low-Z elements. When line centers are clustered and their individual line widths do not differ significantly, it becomes possible to sum their strengths into a single aggregate entity before generating the line profile. If the variation in line widths is modest, a weighted average can be used to assign a representative width to the resulting “super-line”. However, when line widths exhibit large disparities—such as the coexistence of both narrow and broad features—a more refined strategy is required. Abdallah et al.’s technique segments the bound–bound photon energy domain into very fine histogram bins, typically on the scale of one-tenth (or less) of the resolution step used in the final opacity grid. The method employs histogram arrays with sizes on the order of 106, which is dramatically smaller than the number of individual spectral lines, which may reach hundreds of millions depending on physical conditions such as temperature and density. The entire spectral line database is first scanned to identify lines whose centers fall within each bin, and for each bin, the total line strength, a weighted average of the line widths, and a weighted average of the inverse widths are computed. These statistical measures help distinguish between bins dominated by broad or narrow lines. A second pass through the data separates the lines accordingly and aggregates them within their respective categories. Finally, for each populated bin, two distinct super-lines are constructed—one representing the broad lines, the other the narrow ones—mapped onto the final opacity calculation grid. Any bins with zero total line strength are discarded.
The mixed-UTA (MUTA) method [52] combines detailed and statistical treatments within the same transition array, as well as for a set of transition arrays used in the spectral or opacity calculations of light and mid-Z elements. By selecting the strongest lines within each transition array, the method delivers a spectral description with accuracy comparable to a fully detailed treatment, where all lines are explicitly included in the spectral calculation. The weaker lines are represented using a UTA-like functional form. As a result, the computational cost of this method is similar to that of the statistical UTA method. The approach was implemented in the Cowan atomic structure code and successfully applied to the calculation of spectra for both local thermodynamic equilibrium (LTE) and non-LTE plasmas of two mid-Z elements, xenon and iron. However, it is important to keep in mind that such an approach is difficult to use in intermediate coupling. Indeed, the MUTA approach involves gathering the weak lines (with the approach mentioned in Section 2) into a “global” structure, a kind of “pillow of lines”, and keeping only the strongest ones. This is possible close to pure LS coupling, when the UTA approach applies, or close to pure jj coupling, where the SOSA formalism is appropriate. In the latter case, the transition array shows two main relativistic sub-structures (see Figure 19) (the third one is very small and not visible). It is clear that in that case, using a single “pillow” (typically modeled by a Gaussian) would be deleterious, filling the gap between the two ensembles of SOSAs (i.e., the two relativistic subarrays). In that case, the approach may be extended, in the sense that we could attribute some of the weak lines to one of the subarrays, and the others to the second one. But in intermediate coupling, which is the most frequent case and the only way to account properly for configuration interactions, the applicability of the method may seem questionable. For instance, let us consider the transition array Br XI 3d4 4f3 − 3d3 4f4. In that case, two main relativistic substructures, corresponding to the relativistic subarrays 3d3/2 − 4f5/2 and 3d5/2 4f7/2 (the subarray 3d5/2 − 4f5/2 does exist, but it is too small to be clearly visible), are not clearly separated because the atomic number is not sufficiently high (spin–orbit splitting scales as Z 4 ).
Colgan et al. studied the spectra produced from a short-wavelength long-pulse (nanosecond) laser incident on Mg plasma [85]. A very complex experimental spectrum was analyzed in detail via comparison with large-scale atomic kinetics calculations using the recently developed mixed-UTA (MUTA) model. The authors found that the experimental spectrum exhibited lines from many inner-shell transitions from ions ranging from neutral Mg to Li-like Mg. In particular, lines from transitions such as 1s-3, 1s-4 and from hollow atoms (in this context, hollow atoms refer to ions with an empty 1s subshell) were observed.
The resolved transition array (RTA) approach is a different kind of statistical method in which lines are picked up at random in a joint distribution of line energies and amplitudes. Unlike the UTA model, in which the main characteristics (mean energy variance) of the transition arrays are exact (calculated in the framework of the second-quantization formalism and Racah angular-momentum techniques), the RTA approach is approximate (the lines are not at their exact energy and do not have their exact strength) but is very useful in order to take into account the “porosity” of transition arrays when the detailed computations are not feasible and the usual statistical methods (UTA, SOSA) are not applicable (at low density, for instance). The RTA method was recently successfully applied in the context of kilonova emission [86].
Finally, let us finish this section with a rather surprising link between line merging in spectroscopy and number theory. Interference broadening coefficients describe the non-Lorentzian effect that arises as pressure causes lines to overlap. These coefficients, one for each line, are at moderate pressures related linearly to absorption and dispersion. In the course of a work on interference broadening coefficients [87,88], Eakin obtained the following expression:
1 exp 1 ( x a ) 2 + b 2 + 1 ( x + a ) 2 + b 2 d x = 2 π b 1 + p = 2 ( 1 ) p 1 [ 2 ( p 1 ) ] ! ( 2 b ) 2 p 2 p ! [ ( p 1 ) ! ] 2 × 1 + r = 1 p 1 p ! [ 2 ( p r 1 ) ] ! ( 2 r 1 ) ! ( p 2 ) ! b 2 r [ ( p r 1 ) ! ] 2 ( 2 p 3 ) ! ( p + 2 r ) ! ( a 2 + b 2 ) r × k = 0 r 1 2 ( 1 ) k ( p + 3 r 2 k 1 ) ! ( 2 r k ) ! k ! ( r 2 k 1 ) ! ,
giving, for completely overlapping lines, an identity for Mersenne numbers [89,90]:
2 n 1 = r = 1 n ( n + 1 ) ! [ 2 ( n r ) ] ! ( 2 r 1 ) ! ( n 1 ) ! [ ( n r ) ! ] 2 ( 2 n 1 ) ! ( n + 2 r + 1 ) ! k = 0 r 1 2 ( 1 ) k ( n + 3 r 2 k ) ! ( 2 r k ) ! k ! ( r 2 k 1 ) ! .
Note that the left-hand side of Equation (27) is precisely the form for which Bykov and Hien [16] found an expression based on Bessel functions (see also Section 4.3).

5. Conclusions

We have presented analytical formulas in order to find the maximum line strength in a transition array. We have also given an approximation of the Ladenburg–Reiche formula for the equivalent width, and of the Elsasser function, which is a key ingredient of the overlapping model. We have also reviewed the different approximations of overlapping and blanketing, and we discussed the possibility of using a statistical indicator such as the Chernoff bound and the Gini coefficient (see Appendix B and Appendix C). The statistical properties of spectra are important for checking opacity databases [11,13]. We plan to investigate the merging of lines in the framework of other distributions, such as Holtzmark’s one [91]:
ϕ ( u ) = 5 sin 2 π 5 4 π w 1 1 + u w 5 / 2 .
Stark broadening of hydrogen-like lines in plasma is predominantly ion quasi-static in character. Griem has proposed a generalized Holtsmark profile to describe the ion quasi-static profile of hydrogen-like lines [92]. The analytical expression in Equation (28) matches his function in the line wings and differs from it by no more than 10% over the line width. In practice, the central, ion quasi-static Stark structure is smoothed out by Doppler and electron impact broadening, and for many applications, the generalized Holtsmark and Equation (28) adequately describe the Stark profile of hydrogen-like lines.
In order to improve our understanding of the statistical properties of transition arrays, we plan to compute the moments of the Porter–Thomas distribution [93,94] and to compare them with values inferred from measured or computed complex atomic spectra. We believe that the strong connections with random matrix theory need to be developed further [95].

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author. The data are not publicly available due to privacy and ethical restrictions.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

The following abbreviations are used in this manuscript:
GOEGaussian Orthogonal Ensemble
LTELocal Thermodynamic Equilibrium
MUTAMixed Unresolved Transition Arrays
RTAResolved Transition Arrays
UTAUnresolved Transition Arrays
SOSASpin–Orbit–Split Arrays

Appendix A. On the Computation of the Error Function

An expansion of the error function, which converges more rapidly for all real values of x than a Taylor expansion, is obtained using Bürmann’s theorem [96]:
erf x = 2 π sgn x · 1 e x 2 1 1 12 1 e x 2 7 480 1 e x 2 2 5 896 1 e x 2 3 787 276480 1 e x 2 4 = 2 π sgn x · 1 e x 2 π 2 + k = 1 c k e k x 2 .
The resulting approximation shows its largest relative error at x = ± 1.3796 , where it is less than 0.0036127:
erf x 2 π sgn x · 1 e x 2 π 2 + 31 200 e x 2 341 8000 e 2 x 2 .

Appendix B. The Chernhoff Bound

In probability theory, the Chernoff bound provides a powerful way to estimate how rapidly the tail of a random variable’s distribution decreases. It relies on the moment-generating function and gives an exponentially tight upper bound on the probability that a random variable deviates from its expected value [97]. The tightest of these exponential bounds is known as the Chernoff (or Chernoff–Cramér) bound, which can sometimes decay even faster than an exponential rate—for instance, in the case of sub-Gaussian distributions. This bound is particularly effective for analyzing the sum of independent random variables, such as a collection of Bernoulli trials. Compared to classical tail inequalities based on the first or second moment—like Markov’s inequality or the Bienaymé–Chebyshev inequality—the Chernoff bound offers a much sharper estimate [98], since those older bounds typically produce slower, polynomial decay rates. However, it is important to note that the Chernoff bound applies only under the assumption that the random variables are independent, a restriction not imposed by Markov’s or Chebyshev’s inequalities.
Let us consider X 1 , X 2 , , X n arbitrary binary (i.e., X i { 0 , 1 } ) random variables, with probability p i = P ( X i = 1 ) of their respective probabilities. We have (with a > 0 ):
P i = 1 n X i > t = P exp a i = 1 n X i > t > exp ( a t ) .
According to the Markov inequality applied to the right-hand side of the latter equation, one gets
P i = 1 n X i > t E exp a i = 1 n X i exp ( a t )
or equivalently,
P i = 1 n X i > t exp ( a t ) i = 1 n E exp a X i .
Then, setting t = ( 1 + β ) N with β > 0 and a = log ( 1 + β ) , one obtains the following:
P i = 1 n X i > ( 1 + β ) N e a ( 1 + β ) N i = 1 n E 1 + β X i .
Since E 1 + β X i = p i ( 1 + β ) + ( 1 p i ) = 1 + β p i e β p i , one has
P i = 1 n X i > ( 1 + β ) N e β N 1 + β ( 1 + β ) N .
Chernoff bounds are also useful for accelerating randomized algorithms. For instance, in the case of an algorithm that outputs a guess that is the desired answer with probability p > 1 / 2 , a higher success rate can be achieved by running the algorithm n = log ( 1 / δ ) 2 p / ( p 1 / 2 ) 2 times and outputting a guess that is output by more than n / 2 runs of the algorithm. Since there is solely one such guess, and assuming that these algorithm runs are independent, the probability of having more than n / 2 correct guesses is equal to the probability that the sum of independent Bernoulli random variables X k , for which the probability of being equal to 1 is p, exceeds n / 2 . It turns out that this can be shown to be at least 1 δ via the multiplicative Chernoff bound:
P X > n 2 1 exp p 1 / 2 2 p n 2 1 δ .
It is worth mentioning that, as shown by Huber [99], the Markov, Chebyshev and Chernoff inequalities are among the most commonly used tools for estimating the tail probabilities of random variables. Notably, these bounds are considered tight—there are straightforward examples where the inequality becomes an equality. It turns out that a simple smoothing technique involving auxiliary randomness can reduce each of these bounds by a factor of two. Moreover, in many typical scenarios, this improvement can be attained even without introducing additional randomness.
There are other interesting bounds, such as the Chernoff bound with additive error. Let X 1 , X 2 , , X n be independent random variables with X i { 0 , 1 } and X = i = 1 n X i . Then, for any δ > 0 , one has the following ( E ( X ) is the expected value of random variable X):
P X > E ( X ) + δ e 2 δ 2 / n .
Replacing the condition X i { 0 , 1 } with X i [ a i , b i ] yields Hoeffding’s inequality [100]:
P | X E ( X ) | δ 2 exp 2 δ 2 i = 1 n ( a i b i ) 2 .
Finally, replacing the condition X i [ a i , b i ] with | X i | M , one obtains the Bernstein inequality [101]:
P X δ exp 2 δ 2 i = 1 n E X i 2 + M δ 3 .

Appendix C. The Gini Coefficient: Applicable to Atomic Spectra?

In economics, the Gini coefficient, also referred to as the Gini index or Gini ratio, is a measure of statistical dispersion used to represent income inequality, wealth inequality, or consumption inequality within a country or social group.
The Gini coefficient [102,103] was introduced in astronomy by Abraham et al. [104] and has since been used in morphological studies of intrinsic (unlensed) galaxies (e.g., Lotz et al. [105]). It is actually a measurement of the inequality of the distribution of light in a galaxy. Conceptually, it is calculated by ordering the pixels in a given aperture by flux, then giving the cumulative distribution function and finding the area between that curve and the curve representing the cumulative distribution function describing a galaxy with a perfectly flat profile [106].
The Gini coefficient may find applications in the statistical study of atomic spectra. It can be defined as half of the “relative mean absolute difference”, which is rigorously equivalent to the definition based on the Lorenz curve [107] (see the explanation below). The mean absolute difference is the average absolute difference of all pairs of objects (spectral lines in our case), and the relative mean absolute difference is the mean absolute difference divided by the average, x ¯ , to normalize for scale. If x i is the strength of line i, and there are n lines, then the Gini coefficient G would be given as follows:
G = i = 1 n j = 1 n x i x j 2 n 2 x ¯ = i = 1 n j = 1 n x i x j 2 n i = 1 n x i .
For an ensemble of n lines with strengths y 1 y 2 y n , one has the following:
G = 1 n n + 1 2 i = 1 n ( n + 1 i ) y i i = 1 n y i = 2 n i = 1 n i y i i = 1 n y i n + 1 n .
When the number of lines is large, the line strength distribution can be modeled as a continuous probability density function P ( x ) , where P ( x ) d x is the fraction of the lines with strength in the interval d x of about x. Then, the Gini coefficient is again half of the relative mean absolute difference:
G = 1 2 μ P ( x ) P ( y ) | x y | d x d y ,
where μ = x p ( x ) d x is the mean of the distribution, and the lower limits of integration may be replaced by zero when all strengths are positive [108]. If F ( x ) is the cumulative distribution function for f ( x ) :
F ( x ) = 0 x P ( u ) d u
and L ( x ) is the Lorenz function:
L ( x ) = 0 x u P ( u ) d u 0 u P ( u ) d u ,
then the Lorenz curve L ( F ) can be represented as a function parametric in L ( x ) and F ( x ) . The Gini coefficient can then be thought of as the ratio of the area that lies between the line of equality and the Lorenz curve ( A ) over the total area under the line of equality ( A + B ) , i.e., G = A / ( A + B ) with
B = 0 1 L ( F ) d F .
The Gini coefficient can also be calculated directly from the cumulative distribution function of the distribution F ( y ) . Defining μ as the mean of the distribution, the Gini coefficient is given as follows:
G = 1 1 μ 0 1 F ( y ) 2 d y = 1 μ 0 F ( y ) 1 F ( y ) d y ,
the latter result coming from an integration by parts.

References

  1. Moszkowski, S.A. Some Statistical Properties of Level and Line Distributions in Atomic Spectra; Rand Corporation Research Memorandum RM-2610-AEC; RAND: Santa Monica, CA, USA, 1960. [Google Scholar]
  2. Bauche-Arnoult, C.; Bauche, J.; Klapisch, M. Variance of the distributions of energy levels and of the transition arrays in atomic spectra. Phys. Rev. A 1979, 20, 2424–2439. [Google Scholar] [CrossRef]
  3. Bauche, J.; Bauche-Arnoult, C. Level and line statistics in atomic spectra. J. Phys. B At. Mol. Phys. 1987, 20, 1659–1677. [Google Scholar] [CrossRef]
  4. Bauche, J.; Bauche-Arnoult, C.; Klapisch, M. Transition arrays in the spectra of ionized atoms. Adv. At. Mol. Phys. 1988, 23, 131–195. [Google Scholar]
  5. Rose, S.J. Calculations of the radiative opacity of laser-produced plasmas. J. Phys. B At. Mol. Opt. Phys. 1992, 25, 1667–1681. [Google Scholar] [CrossRef]
  6. Gaffney, J.A.; Rose, S.J. The effect of unresolved transition arrays on plasma opacity calculations. High Energy Density Phys. 2011, 7, 240–246. [Google Scholar] [CrossRef]
  7. Pain, J.-C.; Gilleron, F.; Bauche, J.; Bauche-Arnoult, C. Effect of third-and fourth-order moments on the modeling of unresolved transition arrays. High Energy Density Phys. 2009, 5, 294–301. [Google Scholar] [CrossRef]
  8. Kramida, A.; Ralchenko, Y.; Reader, J. NIST ASD Team. NIST Atomic Spectra Database, Version 5.10; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2024. [Google Scholar] [CrossRef]
  9. Kato, D.; Murakami, I.; Tanaka, M.; Banerjee, S.; Gaigalas, G.; Radžiūtė, L.; Rynkun, P. Japan-Lithuania Opacity Database for Kilonova. 2021. Available online: http://dpc.nifs.ac.jp/DB/Opacity-Database/ (accessed on 27 April 2025).
  10. Ryabchikova, T.; Piskunov, N.; Kurucz, R.L.; Stempels, H.C.; Heiter, U.; Pakhomov, Y.; Barklem, P.S. A major upgrade of the VALD database. Phys. Scr. 2015, 90, 054005. [Google Scholar]
  11. Pain, J.-C.; Ralchenko, Y. Benford’s law in atomic spectra and opacity databases. J. Quant. Spectrosc. Radiat. Transf. 2024, 322, 109010. [Google Scholar] [CrossRef]
  12. 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 Get access Arrow. Mon. Not. Roy. Astron. Soc. 2023, 519, 2862–2878. [Google Scholar] [CrossRef]
  13. Pain, J.-C.; Croset, P. Ideas and tools for error detection in opacity databases. Atoms 2023, 11, 27. [Google Scholar] [CrossRef]
  14. Elsasser, W.M. Mean absorption and equivalent absorption coefficient of a band spectrum. Phys. Rev. 1938, 54, 126–129. [Google Scholar] [CrossRef]
  15. Benedict, W.S.; Herman, R.; Moore, G.E.; Silverman, S. The strengths, widths, and shapes of infrared lines: I. General considerations. Can. J. Phys. 1956, 34, 830–849. [Google Scholar] [CrossRef]
  16. Bykov, G.A.; Hien, P.Z. Calculation of parameters of experimental spectra of resonance absorption of gamma quanta in crystals. Soviet Phys. JETP 1963, 16, 646. [Google Scholar]
  17. Porter, C.E.; Thomas, R.G. Fluctuations of nuclear reaction widths. Phys. Rev. 1956, 104, 483–491. [Google Scholar] [CrossRef]
  18. Porter, C.E. Statistical Theories of Spectra: Fluctuations; Academic Press: New York, NY, USA, 1965. [Google Scholar]
  19. Bauche, J.; Bauche-Arnoult, C. Statistical properties of atomic spectra. Comp. Phys. Rep. 1990, 12, 3–28. [Google Scholar] [CrossRef]
  20. Bauche-Arnoult, C.; Bauche, J.; Wyart, J.-F.; Fournier, K.B. Effects of the exchange Slater integrals on the shapes of transition arrays. J. Quant. Spectrosc. Radiat. Transf. 2000, 65, 57–70. [Google Scholar] [CrossRef]
  21. Rosenzweig, N.; Porter, C.E. “Repulsion of energy levels” in complex atomic spectra. Phys. Rev. 1960, 120, 1698–1714. [Google Scholar] [CrossRef]
  22. Camarda, H.S.; Georgopulos, P.D. Statistical behavior of atomic energy levels: Agreement with Random-Matrix Theory. Phys. Rev. Lett. 1983, 50, 492–495. [Google Scholar] [CrossRef]
  23. Flambaum, V.V.; Gribakina, A.A.; Gribakin, G.F. Statistics of electromagnetic transitions as a signature of chaos in many-electron atoms. Phys. Rev. A 1998, 58, 230–237. [Google Scholar] [CrossRef]
  24. Bogomolny, E.; Sieber, M. Eigenfunction distribution for the Rosenzweig-Porter model. Phys. Rev. E 2018, 98, 032139. [Google Scholar] [CrossRef]
  25. Bauche, J.; Bauche-Arnoult, C.; Wyart, J.-F. Intensity distribution in atomic transition arrays. J. Quant. Spectrosc. Radiat. Transf. 1994, 51, 41–47. [Google Scholar] [CrossRef]
  26. Bauche, J.; Bauche-Arnoult, C.; Peyrusse, O. Atomic Properties in Hot Plasmas: From Levels to Superconfigurations; Springer International Publishing: Berlin/Heidelberg, Germany, 2015. [Google Scholar]
  27. Wilson, B.G.; Rogers, F.; Iglesias, C. Random-matrix method for the simulation of large atomic E1 transition arrays. Phys. Rev. A 1998, 37, 2695–2697. [Google Scholar] [CrossRef] [PubMed]
  28. Cowan, R.D. The Theory of Atomic Structure and Spectra; University of California Press: Berkeley, CA, USA, 1981. [Google Scholar]
  29. Kendall, M.G.; Stuart, A. Advanced Theory of Statistics; Hafner: New York, NY, USA, 1969; Volume 1. [Google Scholar]
  30. Corless, R.M.; Gonnand, G.H.; Hare, D.E.G.; Jeffrey, D.J.; Knuth, D.E. On the Lambert W function. Adv. Comput. Math. 1996, 5, 329–359. [Google Scholar] [CrossRef]
  31. Caillol, J.M. Some applications of the Lambert W function to classical statistical mechanics. J. Phys. A Math. Gen. 2003, 36, 10431. [Google Scholar] [CrossRef]
  32. Veberič, D. Lambert W function for applications in physics. Comput. Phys. Commun. 2012, 183, 2622–2628. [Google Scholar] [CrossRef]
  33. Comtet, L. Inversion de yα ey et y logα y au moyen des nombres de Stirling. C. R. Acad. Sci. Paris (In French). 1970, 270, 1085–1088. [Google Scholar]
  34. Jeffrey, D.J.; Corless, R.M.; Hare, D.E.G.; Knuth, D.E. On the inversion of yα ey in terms of associated Stirling numbers. C. R. Acad. Sci. Paris 1995, 320, 1449–1452. [Google Scholar]
  35. Hassani, M. Approximation of the Lambert W function. Rgmia Res. Rep. Collect. 2005, 8, 12. [Google Scholar]
  36. Comtet, L. Advanced Combinatorics; D. Reidel Publishing Company: Berlin/Heidelberg, Germany, 1974. [Google Scholar]
  37. Mitrinović, D.S.; Mitrinović, R.S. Tableaux qui fournissent des polynômes de Stirling. Publ. Fac. Electrotech. Univ. Belgrade Ser. Math. Phys. 1960, 34, 1–23. [Google Scholar]
  38. Karanicoloff, C. Sur une représentation des nombres de Stirling dans une forme explicite. Publ. Fac. Electrotech. Univ. Belgrade Ser. Math. Phys. 1961, 67, 9–10. (In French) [Google Scholar]
  39. Fritsch, F.N.; Shafer, R.E.; Crowley, W.P. Solution of the transcendental equation wew = x. Commun. ACM 1973, 16, 123–124. [Google Scholar] [CrossRef]
  40. Bauche, J.; Bauche-Arnoult, C. Scars of symmetries in complex atomic spectra. Phys. Rev. Lett. 1992, 69, 1038–1041. [Google Scholar] [CrossRef] [PubMed]
  41. Bauche, J.; Bauche-Arnoult, C.; Wyart, J.-F.; Duffy, P.; Klapisch, M. Simulation of atomic transition arrays for opacity calculations. Phys. Rev. A 1991, 44, 5707–5714. [Google Scholar] [CrossRef] [PubMed]
  42. Dufty, P.; Klapisch, M.; Bauche, J.; Bauche-Arnoult, C. Monte Carlo simulation of complex spectra for opacity calculations. Phys. Rev. A 1991, 44, 5715–5724. [Google Scholar]
  43. Bauche, J.; Bauche-Arnoult, C. Recent progress in the global description of atomic transitions. Phys. Scr. T 1996, 65, 99–103. [Google Scholar] [CrossRef]
  44. Gilleron, F.; Bauche, J.; Bauche-Arnoult, C. A Statistical approach for simulating detailed-line spectra. J. Phys. B At. Mol. Opt. Phys. 2007, 40, 3057–3074. [Google Scholar] [CrossRef]
  45. Gumbel, E.J. Les valeurs extrêmes des distributions statistiques. Ann. L’Institut Henri Poincaré 1935, 5, 115–158. (In French) [Google Scholar]
  46. Gumbel, E.J.; Lieblein, J. Statistical Theory of Extreme Values and Some Practical Applications: A Series of Lectures; US Government Printing Office: Washington, DC, USA, 1954; Volume 33. [Google Scholar]
  47. Gumbel, E.J. Statistics of Extremes; New York Chichester; Columbia University Press: West Sussex, UK, 1958. [Google Scholar]
  48. Learner, R.C.M. A simple (and unexpected) experimental law relating to the number of weak line in a complex spectrum. J. Phys. B Atom. Mol. Phys. 1982, 15, L891–L895. [Google Scholar] [CrossRef]
  49. Pain, J.-C. Regularities and symmetries in atomic structure and spectra. High Energy Density Phys. 2013, 9, 392–401. [Google Scholar] [CrossRef]
  50. Pain, J.-C.; Croset, P. Checking the reliability of opacity databases. Eur. Phys. J. D 2023, 77, 60. [Google Scholar] [CrossRef]
  51. Porcherot, Q.; Pain, J.-C.; Gilleron, F.; Blenski, T. A consistent approach for mixed detailed and statistical calculation of opacities in hot plasmas. High Energy Density Phys. 2011, 7, 234–239. [Google Scholar] [CrossRef]
  52. Mazevet, S.; Abdallah, J. Mixed UTA and detailed line treatment for mid-Z opacity and spectral calculations. J. Phys. B At. Mol. Opt. Phys. 2006, 39, 3419–3429. [Google Scholar] [CrossRef]
  53. Robert, E. (Unpublished). Available online: https://fr.wikipedia.org/wiki/Fonction_d%27erreur (accessed on 27 April 2025).
  54. Carlitz, L. The inverse of the error function. Pacific J. Math. 1963, 13, 459–470. [Google Scholar] [CrossRef]
  55. Blair, J.M.; Edwards, C.A.; Johnson, J.H. Rational Chebyshev approximations for the inverse of the error function. Math. Comp. 1976, 30, 827–830. [Google Scholar] [CrossRef]
  56. Poppe, G.P.M.; Wijers, C.M.J. More efficient computation of the complex error function. Assoc. Comput. Mach. 1990, 16, 38–46. [Google Scholar] [CrossRef]
  57. Whitman, R.I. Effects of Three Transmission Models in the Rotational Water-Vapor Band on Radiance Calculations and Constituent Inferences; National Aeronautics and Space Administration Report 6112; NASA: Washington, DC, USA, 1971. [Google Scholar]
  58. Burn, J.W.; Uplinger, W.G. The Determination of Atmospheric Temperature Profiles from Planetary Limb Radiance Profiles; National Aeronautics and Space Administration Report 71606130; NASA: Washington, DC, USA, 1970. [Google Scholar]
  59. Ladenburg, R.; Fritz, R. Über selektive Absorption. Ann. Phys. 1923, 347, 181–209. [Google Scholar] [CrossRef]
  60. Reiche, F. Total absorption and overlapping effect due to two spectral lines of equal width and strength. J. Opt. Soc. Am. 1956, 46, 590–597. [Google Scholar] [CrossRef]
  61. Nossal, R.; Weiss, G.H. Analysis of a generalized Ladenburg-Reiche integral. J. Quant. Spectrosc. Radiat. Transf. 1968, 8, 763–770. [Google Scholar] [CrossRef]
  62. Maass, F.; Martin, P.; Olivares, J. Analytic approximation to Bessel function J0(x). Comput. Appl. Math. 2020, 39, 222. [Google Scholar] [CrossRef]
  63. Mayer, H. Methods of Opacity Calculations; Los Alamos Report LA 647; Los Alamos Scientific Laboratory: Los Alamos, NM, USA, 1947. [Google Scholar]
  64. Poirier, M. Analytical description of absorption profiles in warm dense plasmas: Discussion of the non-overlapping-line limit. High Energy Density Phys. 2013, 9, 618–625. [Google Scholar] [CrossRef]
  65. Kaplan, L.D.; Eggers, D.F. Intensity and line-width of the 15-micron CO2 band, determined by a curve-of-growth method. J. Chem. Phys. 1956, 25, 876–883. [Google Scholar] [CrossRef]
  66. Seitz, W.S.; Lundholm, D.V. Elsasser model for band absorption; series representation of a useful integral. J. Opt. Soc. Am. 1964, 54, 315–318. [Google Scholar] [CrossRef]
  67. Kiselev, K.A. Analytical calculation of the Elsasser integral. J. Appl. Spectrosc. 1973, 19, 1488–1491. [Google Scholar] [CrossRef]
  68. Arfken, G. Modified Bessel functions, Iν(x) and Kν(x). In Mathematical Methods for Physicists, 3rd ed.; Academic Press: Orlando, FL, USA, 1985; pp. 610–616. [Google Scholar]
  69. Goody, R.M.; Yung, Y.L. The Elsasser Function. In Atmospheric Radiation: Theoretical Basis, 2nd ed.; Oxford Academic: New York, NY, USA, 1989. [Google Scholar]
  70. Kaplan, L.D. Regions of validity of various absorption-coefficient approximations. J. Meteorol. 1953, 10, 100–104. [Google Scholar] [CrossRef]
  71. Plass, G.N. Regions of validity of various absorption-coefficient approximations. J. Atmos. Sci. 1954, 11, 163–164. [Google Scholar] [CrossRef]
  72. Wark, D.Q.; Wolk, M. An extension of a table of absorption for Elsasser bands. Mon. Weather Rev. 1960, 88, 249–250. [Google Scholar] [CrossRef]
  73. Zachor, A.S. Absorptance and radiative transfer by a regular band. J. Quant. Spectrosc. Radial. Transf. 1967, 7, 857–870. [Google Scholar] [CrossRef]
  74. Goody, R.M. A statistical model for water-vapor absorption. Quart. J. Roy. Meteorol. Soc. 1952, 78, 165–169. [Google Scholar] [CrossRef]
  75. Plass, G.N. Useful representations for measurements of spectral band absorption. J. Opt. Soc. Am. 1960, 50, 868–875. [Google Scholar] [CrossRef]
  76. Vardya, M.S. Simplified description of the line absorption coefficient in blanketing effect computation. J. Quant. Spectrosc. Radiat. Transf. 1966, 6, 539–548. [Google Scholar] [CrossRef]
  77. Sakai, H.; Stauffer, F.R. Equivalent width due to two overlapping lines. J. Opt. Soc. Am. 1964, 54, 759–762. [Google Scholar] [CrossRef]
  78. Hirono, M.; Suda, T. Band absorption by the spectral lines with a super-Lorentz line shape. J. Phys. Soc. Jpn. 1983, 52, 1157–1163. [Google Scholar] [CrossRef]
  79. Berk, A. Voigt equivalent widths and spectral-bin single-line transmittances: Exact expansions and the MODTRAN 5 implementation. J. Quant. Spectrosc. Radiat. Transf. 2013, 118, 102–120. [Google Scholar] [CrossRef]
  80. Tien, C.L. A simple approximate formula for radiation from a line with resonance contours. J. Quant. Spectrosc. Radiat. Transf. 1966, 6, 893–894. [Google Scholar] [CrossRef]
  81. Goldman, A. On simple approximations to the equivalent width of a Lorentz line. J. Quant. Spectrosc. Radiat. Transf. 1968, 8, 829–831. [Google Scholar] [CrossRef]
  82. Martin, P.; Olivares, J.; Sotomayor, A. Precise analytic approximation for the modified Bessel function I1(x). Rev. Mex. Fis. 2017, 63, 130–133. [Google Scholar]
  83. Olivares, J.; Martin, P.; Valero, E. A simple approximation for the modified Bessel function of zero order I0(x). J. Phys. Conf. Ser. 2018, 1043, 012003. [Google Scholar] [CrossRef]
  84. Abdallah, J., Jr.; Kilcrease, D.P.; Magee, N.H.; Mazevet, S.; Hakel, P.; Sherrill, M.E. Spectral line strength binning method for opacity calculations. High Energy Density Phys. 2007, 3, 309–313. [Google Scholar] [CrossRef]
  85. Colgan, J.; Abdallah, J.; Faenov, A.; Pikuz, T.A.; Skobelev, I.Y. MUTA calculations of a laser-produced Mg hollow atom spectrum. Phys. Scr. 2008, 78, 015302. [Google Scholar] [CrossRef]
  86. Deprince, J.; Carvajal Gallego, H.; Ben Nasr, S.; Maison, L.; Pain, J.-C.; Palmeri, P.; Quinet, P. Radiative and opacity data obtained from large-scale atomic structure calculations and from statistical simulations for the spectral analysis of kilonovae in their photospheric and nebular phases: The sample case of Er III. Eur. Phys. J. D 2024, 78, 105. [Google Scholar] [CrossRef]
  87. Eakin, R.T. Interference in transmission spectra of overlapping Lorentzian lines. J. Quant. Spectrosc. Radiat. Transf. 1988, 39, 225–236. [Google Scholar] [CrossRef]
  88. Rosenkranz, W. Interference coefficients for overlapping oxygen lines in air. J. Quant. Spectrosc. Radiat. Transf. 1988, 39, 281–297. [Google Scholar] [CrossRef]
  89. Dickson, L.E. Divisibility and Primality. In History of the Theory of Numbers; Dover: New York, NY, USA, 2005; Volume 1, p. 13. [Google Scholar]
  90. Eakin, R.T. A combinatorial partition of Mersenne numbers arising from spectroscopy. J. Number Theory 2012, 132, 2166–2183. [Google Scholar] [CrossRef]
  91. Irons, F.E. The escape factor in plasma spectroscopy–I. The escape factor defined and evaluated. J. Quant. Spectrosc. Radiat. Transf. 1979, 22, 1–20. [Google Scholar] [CrossRef]
  92. Griem, H.R. Stark broadening of higher hydrogen and hydrogen-like lines by electrons and ions. Astrophys. J. Lett. 1960, 132, 883–893. [Google Scholar] [CrossRef]
  93. Basor, E.L.; Tracy, C.A.; Widom, H. Asymptotics of level-spacing distributions for random matrices. Phys. Rev. Lett. 1992, 69, 5–8. [Google Scholar] [CrossRef]
  94. Majumdar, S.N.; Nadal, C.; Scardicchio, A.; Vivo, P. Index distribution of Gaussian random matrices. Phys. Rev. Lett. 2009, 103, 220603. [Google Scholar] [CrossRef]
  95. Pain, J.-C. Random-matrix theory for complex atomic spectra. Chin. J. Phys. 2012, 50, 523–532. [Google Scholar]
  96. Schöpf, H.M.; Supancic, P.H. On Bürmann’s theorem and its application to problems of linear and nonlinear heat transfer and diffusion. Math. J. 2014, 16, 1–44. [Google Scholar] [CrossRef]
  97. Chernoff, H. A Measure of asymptotic efficiency for tests of a hypothesis based on the sum of observations. Ann. Math. Stat. 1952, 23, 493–507. [Google Scholar] [CrossRef]
  98. Papoulis, A. Probability, Random Variables, and Stochastic Processes, 3rd ed.; McGraw-Hill: New York, NY, USA, 1991. [Google Scholar]
  99. Huber, M. Halving the bounds for the Markov, Chebyshev, and Chernoff inequalities using smoothing. Am. Math. Mon. 2019, 126, 915–927. [Google Scholar] [CrossRef]
  100. Hoeffding, W. Probability inequalities for sums of bounded random variables. J. Am. Stat. Assoc. 1963, 58, 13–30. [Google Scholar] [CrossRef]
  101. Bernstein, S.N. On certain modifications of Chebyshev’s inequality. Dokl. Akad. Nauk SSSR 1937, 17, 275–277. [Google Scholar]
  102. Gini, C. Variabilità e mutabilità (1912). In Memorie di Metodologica Statistica; Pizetti, E., Salvemini, T., Eds.; Libreria Eredi Virgilio Veschi: Rome, Italy, 1955. (In Italian) [Google Scholar]
  103. Gini, C. Measurement of inequality of incomes. Econ. J. 1921, 31, 124–126. [Google Scholar] [CrossRef]
  104. Abraham, R.G.; van den Bergh, S.; Nair, P. A new approach to galaxy morphology. I. Analysis of the Sloan Digital Sky Survey Early Data Release. Astrophys. J. Lett. 2003, 588, 218–229. [Google Scholar] [CrossRef]
  105. Lotz, J.M.; Madau, P.; Giavalisco, M.; Primack, J.; Ferguson, H.C. The rest-frame far-ultraviolet morphologies of star-forming galaxies at z ≈ 1.5 and 4. Astrophys. J. Lett. 2006, 636, 592–609. [Google Scholar] [CrossRef]
  106. Florian, M.K.; Gladders, M.D.; Li, N.; Sharo, K. The Gini Coefficient as a tool for image family identification in strong lensing systems with multiple images. Astrophys. J. Lett. 2016, 816, L23. [Google Scholar] [CrossRef]
  107. Sen, A. On Economic Inequality, 2nd ed.; Oxford University Press: Oxford, UK, 1977. [Google Scholar]
  108. Dorfman, R. A Formula for the Gini Coefficient. Rev. Econ. Stat. 1979, 61, 146–149. [Google Scholar] [CrossRef]
Figure 1. The two real branches of the Lambert function: W 0 and W 1 (the variable is strictly larger than 1 / e ).
Figure 1. The two real branches of the Lambert function: W 0 and W 1 (the variable is strictly larger than 1 / e ).
Plasma 08 00017 g001
Figure 2. Relative difference between Formulas (6) and (8) for S max / S , as a function of the total number of lines L. The plotted quantity is (Equation (6) − Equation (8))/(Equation (6)), in %.
Figure 2. Relative difference between Formulas (6) and (8) for S max / S , as a function of the total number of lines L. The plotted quantity is (Equation (6) − Equation (8))/(Equation (6)), in %.
Plasma 08 00017 g002
Figure 3. Expressions (6): S max = v ln 8 L 2 π ln ln 8 L 2 π + ln ln 8 L 2 π / ln 8 L 2 π and (9): S max = v 2 ln ( 2 L ) 2 for S max / S as a function of the total number of lines L.
Figure 3. Expressions (6): S max = v ln 8 L 2 π ln ln 8 L 2 π + ln ln 8 L 2 π / ln 8 L 2 π and (9): S max = v 2 ln ( 2 L ) 2 for S max / S as a function of the total number of lines L.
Plasma 08 00017 g003
Figure 4. Pd VII 4d4 − 4d3 5p computed using Cowan’s code [28]. The total number of lines is 1718.
Figure 4. Pd VII 4d4 − 4d3 5p computed using Cowan’s code [28]. The total number of lines is 1718.
Plasma 08 00017 g004
Figure 5. Transition array Cd V 4d7 5s − 4d7 5p computed using Cowan’s code. The total number of lines is 2082.
Figure 5. Transition array Cd V 4d7 5s − 4d7 5p computed using Cowan’s code. The total number of lines is 2082.
Plasma 08 00017 g005
Figure 6. Test of the relevance of approximation (14) of the erfc function ( erfc ( x ) = 1 erf ( x ) ).
Figure 6. Test of the relevance of approximation (14) of the erfc function ( erfc ( x ) = 1 erf ( x ) ).
Plasma 08 00017 g006
Figure 7. Comparison between Formulas (18) and (19) for S 0 / S .
Figure 7. Comparison between Formulas (18) and (19) for S 0 / S .
Plasma 08 00017 g007
Figure 8. Elsasser function (exact and different approximations) as a function of y for u = 0.01 . The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22). The black, red and blue curves are superimposed on each other.
Figure 8. Elsasser function (exact and different approximations) as a function of y for u = 0.01 . The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22). The black, red and blue curves are superimposed on each other.
Plasma 08 00017 g008
Figure 9. Elsasser function (exact and different approximations) as a function of y for u = 1 . The first approximation is given by Equations (20), the second one by Equation (21) and the third one by Equation (22).
Figure 9. Elsasser function (exact and different approximations) as a function of y for u = 1 . The first approximation is given by Equations (20), the second one by Equation (21) and the third one by Equation (22).
Plasma 08 00017 g009
Figure 10. Elsasser function (exact and different approximations) as a function of y for u = 10 . The black and green curves are superimposed. The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22).
Figure 10. Elsasser function (exact and different approximations) as a function of y for u = 10 . The black and green curves are superimposed. The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22).
Plasma 08 00017 g010
Figure 11. Elsasser function (exact and different approximations) as a function of u for y = 0.01 . The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22). The black and green curves are very close to each other.
Figure 11. Elsasser function (exact and different approximations) as a function of u for y = 0.01 . The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22). The black and green curves are very close to each other.
Plasma 08 00017 g011
Figure 12. Elsasser function (exact and different approximations) as a function of u for y = 1 . The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22). The black, red and blue curves are superimposed.
Figure 12. Elsasser function (exact and different approximations) as a function of u for y = 1 . The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22). The black, red and blue curves are superimposed.
Plasma 08 00017 g012
Figure 13. Elsasser function (exact and different approximations) as a function of u for y = 10 . The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22). The black, red and blue curves are superimposed. The green curve yields unrealistically small values.
Figure 13. Elsasser function (exact and different approximations) as a function of u for y = 10 . The first approximation is given by Equation (20), the second one by Equation (21) and the third one by Equation (22). The black, red and blue curves are superimposed. The green curve yields unrealistically small values.
Plasma 08 00017 g013
Figure 14. Comparison of different approximations of the Ladenburg–Reiche function. The Tien and Varanasi formulas are given by Equations (23) and (24), respectively, and the Goldman (1 and 2) formulas are given by Equations (25) and (26), respectively.
Figure 14. Comparison of different approximations of the Ladenburg–Reiche function. The Tien and Varanasi formulas are given by Equations (23) and (24), respectively, and the Goldman (1 and 2) formulas are given by Equations (25) and (26), respectively.
Plasma 08 00017 g014
Figure 15. Comparison of different approximations of the Ladenburg–Reiche function. The Tien and Varanasi formulas are given by Equations (23) and (24), respectively, and the Goldman (1 and 2) formulas are given by Equations (25) and (26), respectively. Same as Figure 14 but over a wider range.
Figure 15. Comparison of different approximations of the Ladenburg–Reiche function. The Tien and Varanasi formulas are given by Equations (23) and (24), respectively, and the Goldman (1 and 2) formulas are given by Equations (25) and (26), respectively. Same as Figure 14 but over a wider range.
Plasma 08 00017 g015
Figure 16. Relative difference between the exact value of the Ladenburg–Reiche function and the approximation obtained using the approximants of the Bessel functions I 0 and I 1 published by Olivares et al. [83] and Martin et al. [82], respectively.
Figure 16. Relative difference between the exact value of the Ladenburg–Reiche function and the approximation obtained using the approximants of the Bessel functions I 0 and I 1 published by Olivares et al. [83] and Martin et al. [82], respectively.
Plasma 08 00017 g016
Figure 17. Relative difference between the exact value of the Ladenburg–Reiche function and the approximation obtained using the approximants of the Bessel functions I 0 and I 1 published by Olivares et al. [83] and Martin et al. [82], respectively. Same as Figure 14 but over a wider range.
Figure 17. Relative difference between the exact value of the Ladenburg–Reiche function and the approximation obtained using the approximants of the Bessel functions I 0 and I 1 published by Olivares et al. [83] and Martin et al. [82], respectively. Same as Figure 14 but over a wider range.
Plasma 08 00017 g017
Figure 18. Relative difference with respect to the exact value for different approximations of the Ladenburg–Reiche function. The Tien and Varanasi formulas are given by Equations (23) and (24), respectively, and the Goldman (1 and 2) formulas are given by Equations (25) and (26), respectively. The plotted quantity corresponds to (Exact – Approximation)/Exact and is expressed in %.
Figure 18. Relative difference with respect to the exact value for different approximations of the Ladenburg–Reiche function. The Tien and Varanasi formulas are given by Equations (23) and (24), respectively, and the Goldman (1 and 2) formulas are given by Equations (25) and (26), respectively. The plotted quantity corresponds to (Exact – Approximation)/Exact and is expressed in %.
Plasma 08 00017 g018
Figure 19. Transition array Nd 4d8 5s1 − 4d8 5p1. Unlike the transition arrays displayed in Figure 4 and Figure 5, the present arrays shows two main relativistic substructures, corresponding to the relativistic subarrays 5s1/2 − 5p1/2 and 5s1/2 − 5p3/2.
Figure 19. Transition array Nd 4d8 5s1 − 4d8 5p1. Unlike the transition arrays displayed in Figure 4 and Figure 5, the present arrays shows two main relativistic substructures, corresponding to the relativistic subarrays 5s1/2 − 5p1/2 and 5s1/2 − 5p3/2.
Plasma 08 00017 g019
Table 1. Exact number of lines calculated for several transition arrays in different coupling schemes: LS, jj, j and IC (intermediate coupling). In the case of j coupling, the two values correspond to the non-zero spin–orbit integral (specified inside brackets).
Table 1. Exact number of lines calculated for several transition arrays in different coupling schemes: LS, jj, j and IC (intermediate coupling). In the case of j coupling, the two values correspond to the non-zero spin–orbit integral (specified inside brackets).
Transition ArrayLSjjjIC
d4 − d3 p6378591718 (p) − 1053 (d)1718
d5 − d4 f190324713286 (d) − 5470 (f)5470
d5 − d4 p115015723245 (p) − 1925 (d)3245
d9 f − d8 f2133116022780 (d) − 2727 (f)3590
p3 d2 − p2 d36129691513,611 (p) − 12,128 (d)18,237
p3 − p2 d465671 (p) − 97 (d)97
Table 2. Estimates of the maximum strength (calibrated so that the total strengths of the arrays are 2940 and 50,400 in Pd VII and Cd V, respectively [41]) for two different transition arrays: Pd VII 4d4 − 4d3 5p and Cd V 4d7 5s − 4d7 5p. The radial integrals required for the diagonalization of the transition array Pd VII 4d4 − 4d3 5p are given in Table 3.
Table 2. Estimates of the maximum strength (calibrated so that the total strengths of the arrays are 2940 and 50,400 in Pd VII and Cd V, respectively [41]) for two different transition arrays: Pd VII 4d4 − 4d3 5p and Cd V 4d7 5s − 4d7 5p. The radial integrals required for the diagonalization of the transition array Pd VII 4d4 − 4d3 5p are given in Table 3.
Pd VIICd V
4d4 − 4d3 5p4d7 5s − 4d7 5p
Total number of lines17182082
Variance in the amplitudes v1.716124.2064
Exact maximal strength S max 531050
Approximate S max (Formula (6))22.73329.27
Approximate S max (Formula (8))22.53326.55
Approximate S max (Formula (9))56.88840.68
Table 3. Values of the parameters used for the diagonalization of the 4d4 − 4d3 5p array in the Pd VII spectrum in cm−1. F k and G k are direct and exchange Slater integrals, respectively, and ζ denotes a spin–orbit integral.
Table 3. Values of the parameters used for the diagonalization of the 4d4 − 4d3 5p array in the Pd VII spectrum in cm−1. F k and G k are direct and exchange Slater integrals, respectively, and ζ denotes a spin–orbit integral.
4d44d3 5p
F 2 (4d, 4d)73,80076,300
F 4 (4d, 4d)50,50052,700
F 2 (4d, 5p) 30,323
G 1 (4d, 5p) 10,000
G 3 (4d, 5p) 8020
ζ 4 d 22002390
ζ 5 p 5784
Table 4. Number of lines with a strength greater than a fraction of the maximum strength (exact and estimated based on Formulas (13) and (17)) for transition array Pd VII 4d4 − 4d3 5p (see Figure 4). The total number of lines is 1718.
Table 4. Number of lines with a strength greater than a fraction of the maximum strength (exact and estimated based on Formulas (13) and (17)) for transition array Pd VII 4d4 − 4d3 5p (see Figure 4). The total number of lines is 1718.
ϵ No. lines S > ϵ S max No. lines S > ϵ S max No. lines S > ϵ S max
ExactFormula (13)Formula (17)
0.0037811460.331314.87
0.015241253.991054.35
0.03311934.66737.49
0.06185683.32519.56
0.320100.73118.48
0.51425.1554.41
0.6512.8739.14
0.813.4621.80
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.

Share and Cite

MDPI and ACS Style

Pain, J.-C. Strong, Weak and Merging Lines in Atomic Spectra. Plasma 2025, 8, 17. https://doi.org/10.3390/plasma8020017

AMA Style

Pain J-C. Strong, Weak and Merging Lines in Atomic Spectra. Plasma. 2025; 8(2):17. https://doi.org/10.3390/plasma8020017

Chicago/Turabian Style

Pain, Jean-Christophe. 2025. "Strong, Weak and Merging Lines in Atomic Spectra" Plasma 8, no. 2: 17. https://doi.org/10.3390/plasma8020017

APA Style

Pain, J.-C. (2025). Strong, Weak and Merging Lines in Atomic Spectra. Plasma, 8(2), 17. https://doi.org/10.3390/plasma8020017

Article Metrics

Back to TopTop