• Aucun résultat trouvé

Quantifying the statistical noise in computer simulations of Stark broadening

N/A
N/A
Protected

Academic year: 2021

Partager "Quantifying the statistical noise in computer simulations of Stark broadening"

Copied!
18
0
0

Texte intégral

(1)

HAL Id: hal-03215646

https://hal.archives-ouvertes.fr/hal-03215646

Submitted on 11 May 2021

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires

Quantifying the statistical noise in computer simulations of Stark broadening

J. Rosato, Y. Marandet, R. Stamm

To cite this version:

J. Rosato, Y. Marandet, R. Stamm. Quantifying the statistical noise in computer simulations of

Stark broadening. Journal of Quantitative Spectroscopy and Radiative Transfer, Elsevier, 2020, 249,

pp.107002. �10.1016/j.jqsrt.2020.107002�. �hal-03215646�

(2)

Quantifying the statistical noise in computer simulations of Stark broadening

J. Rosato, Y. Marandet, and R. Stamm

Aix-Marseille Universit´ e, CNRS, PIIM UMR 7345, F-13397 Marseille, France

Abstract

Computer simulations employed in Stark broadening calculations are re- examined within the perspective of giving error bars to results. As a rule, the calculated spectra exhibit a noisy structure, which is inherent to the ran- dom number generators involved in the numerical method. Using a variance estimator, we quantify the statistical noise on simulated line shapes. Two expressions for the radiated power spectrum, which are analytically equiv- alent but lead to different noise levels, are considered. A discussion of this difference is carried out based on an analytical model.

Keywords:

1. Introduction

In plasma spectroscopy, the shape of an atomic line is determined by the perturbation of the energy levels due to the charged particles surrounding the emitter or absorber under consideration. This is the celebrated Stark broadening problem [1]. It still has no solution, in the sense that no gen- eral formula has been found for a line shape relative to an arbitrary atomic species. The Dyson series, which provides a solution to the time-dependent Schrdinger equation, is only a formal relation hardly applicable to calcula- tions in realistic conditions. The computer simulation technique has been developed in the seventies and the early eighties with the purpose of repro- ducing the exact solution as closely as possible [2, 3, 4]. Its advantages with respect to other models is that it is free from constraining assumptions such

Email address: joel.rosato@univ-amu.fr ()

(3)

as, for example, the binary condition involved in collision operator-based approaches (impact approximation [5, 6]) or the ad hoc description of the microfield dynamics used in stochastic models (like the MMM [7, 8] and the FFM [9, 10, 11]). It has been recognized that computer simulations can serve as benchmark for line shape models; cross-checks between models and simulations have been reported in the literature and workshops [e.g., at the International Workshop on Radiative Properties of Hot Dense Matter (RPHDM), at the Spectral Line Shapes in Plasmas code comparison work- shop (SLSP)] [12, 13, 14, 15, 16, 17, 18]. Essentially, a simulation consists in numerically integrating the time-dependent Schrdinger equation that gov- erns the dynamics of an atom perturbed by a fluctuating electric field, itself being generated from a numerical integration of the Newtonian equations of motion for the charged particles moving in the vicinity of the emitter. The initial conditions for the perturbers are generated randomly and, due to this, the simulated line shape exhibits a noisy behavior, which can be reduced only by increasing the number of runs. In this work, we examine the noise present in simulated spectra. Two variants of the simulation method are considered:

one involves a direct calculation of the power spectrum, based on the Fourier transform of the dipole operator, and the other one involves an evaluation of the dipole autocorrelation function as intermediary step. Each of these approaches have been used and reported in the literature; they yield identical results [19] but the noise level is different. The article is organized as follows:

section 2 gives an overview of the Stark broadening formalism, with a focus on the numerical simulation method; the quantification of statistical noise is addressed in section 3 in terms of a variance estimator; finally, the two variants of the simulation method are examined in section 4. The difference in noise level obtained from the two approaches is discussed in terms of an analytical model.

2. Stark broadening formalism

The shape of spectral lines observed in a plasma depends upon the envi-

ronmental conditions under which they are formed. The microfield is usually

large enough so that it alters the energy level structure of the emitters and

this alteration results in a broadening of the lines. We give hereafter the

main lines of the formalism involved in the modeling of Stark line broaden-

ing. We consider an atomic system that emits a photon in the presence of

charged particles located at its vicinity; this system will be referred to as an

(4)

“atom” in the following, but it can also be an ion or a molecule. According to quantum electrodynamics, the power spectrum is given by the following relation

P (ω) = ω 4 3πε 0 c 3 lim

T →∞

1 T

αβ

p α

T /2

T /2

dt β | d(t) | α e iωt

2 ⟩

. (1)

Here, d(t) denotes the dipole operator expressed in the Heisenberg picture, the sum is performed over the upper (α) and lower (β) states that contribute to the line under consideration, p α is the probability of the atom being ini- tially in the state α (p α = α | ρ | α with ρ being the density operator) and the brackets ... stand for an average over the charged particles. Equation (1) is a generalization of the Larmor formula to quantum emitter. This expres- sion has been reported as a starting point for plasma spectroscopy models in early textbooks and articles, e.g. [5, 20, 21]. In the appendix, we provide a derivation based on the Poynting theorem expressed in quantum mechan- ical terms, using operators in the Heisenberg picture. It is common to use an alternative to Eq. (1) that employs the Fourier transform of the dipole autocorrelation function. Assuming that the dipole is a stationary process, the power spectrum reads

P (ω) = ω 4

3πε 0 c 3 × 2 Re

0

dt Tr[ρ ⃗ d(0) · d(t)] e iωt . (2) This relation has been widely used for the development of Stark broadening models because it provides direct information on the dipole’s stochastic prop- erties, through the autocorrelation function Tr[ρ ⃗ d(0) · d(t)] ; furthermore, the formula does not involve a limiting process as in Eq. (1). The dipole at time t is obtained from the relation d(t) = U (t) d(0)U (t), where U (t) is the evolution operator; it obeys the time-dependent Schr¨ odinger equation

i ~ dU

dt (t) = [H 0 + V (t)]U (t), (3)

where V (t) = d · E(t) is the Stark effect term that involves the plasma

microfield E(t). According to Eq. (2), one needs to solve the Schr¨ odinger

equation on a time scale which is of the order of the dipole’s characteristic

decorrelation time in order to get a meaningful representation of the spec-

trum; this time is referred to as the “time of interest” in textbooks (e.g., [1]).

(5)

If the electric field is nearly constant during the time of interest, one can neglect its evolution (quasistatic approximation), the Schr¨ odinger equation can be solved exactly and the power spectrum can be written explicitly in terms of the microfield probability density function. The other limiting case can also be tackled analytically by using a set of assumptions that involve the microfield fluctuation time scale (impact approximation); the treatment yields an exponential behavior for the autocorrelation function in Eq. (2) and an explicit relation, involving Lorentzian functions, is obtained for the power spectrum. The issue we focus on hereafter concerns the use of computer simulations in Stark line shape calculations. The main idea of such simula- tions is to reproduce the microfield evolution in a way as realistic as possible by generating particles in a box and making them move around the emitter following prescribed trajectories; the latter are obtained from a molecular dy- namics algorithm. For a set of simulated microfield histories corresponding to different initial conditions, the Schr¨ odinger equation is solved numerically and the resulting dipole operator is evaluated. It is next inserted in the power spectrum formula, either through equation (1) or equation (2). Computer simulations can serve as a benchmark for other models. In early works, they were devoted to address the influence of ion dynamics on Stark line shapes [2, 3, 4]. Several codes have been written and are still under development to- day; applications include astrophysics, inertial and magnetic fusion research, and other laboratory plasma studies, e.g., see [12, 19, 22, 23, 24]. A feature of the simulations is that they involve (pseudo-)random number generators for the initial position and velocity of the perturbers. Simulated spectra are affected by statistical noise and this noise can be important if the number of histories is not large enough. This point is illustrated in Figure 1. The Ly- man α line of an Ar XVIII emitter in a deuterium plasma has been calculated assuming N e,i = 2 × 10 23 cm 3 and T e,i = 1 keV, with 10 and 1000 histories.

The power spectrum has been evaluated using Eq. (2) and only ion Stark broadening has been retained. As can be seen in the figure, the spectrum exhibits a noisy structure at 10 histories, which indicates that the number of runs is not large enough in this case. In contrast, the spectrum obtained with 1000 histories is much cleaner. As a rule, the strength of the noise decreases with the number of statistical realizations. This trend is proper to algorithms based on random number generation (Monte Carlo method).

An estimate of the fluctuation level around the exact result is provided by

variance estimators. We address this issue, together with the convergence of

the simulation toward the exact result, in the next section.

(6)

3315 3320 3325 -0.2

0.0 0.2 0.4 0.6 0.8 1.0

10 histories

1000 histories

Lineshape(arb.units)

Ar XVIII Ly-

D plasma

N e,i

= 2x10 23

cm -3

T e,i

= 1 keV

Energy (eV)

Figure 1: Plot of the Ar XVIII Lyman α line calculated in high density plasma conditions

with 10 and 1000 simulated microfield histories. For the sake of clarity, only ion Stark

broadening has been retained. The noise present in the spectrum is a feature of the random

number generation involved in the algorithm. It decreases with the number of histories.

(7)

3. Variance estimator

We use the same terminology as in the literature on Monte Carlo simu- lation methods, e.g. [25]. The average ... involved in the power spectrum is identical to the expectation value of a random variable X; namely,

X ω 4 3πε 0 c 3 lim

T →∞

1 T

αβ

p α

T /2

T /2

dt β | d(t) | α e iωt

2

(4)

if Eq. (1) is used, and X ω 4

3πε 0 c 3 × 2 Re

0

dtTr[ρ ⃗ d(0) · d(t)]e iωt (5) if Eq. (2) is used. According to the law of large numbers, the expectation value of X can be estimated as the empirical mean of a set of realizations x 1 , x 2 ... of this random variable

X ⟩ ≃ 1 n

n j=1

x j x ¯ n . (6)

A strict equality holds at the n → ∞ limit. In the numerical simulation method, n is kept finite and Equation (6) is used for the spectrum evaluation.

The fluctuation level around the exact result can be estimated in confidence interval terms. As in standard statistical analysis, we define the variance estimator

s 2 n = 1 n 1

n j=1

(x j x ¯ n ) 2 . (7)

If a 99% confidence interval is sought for the result, error bars at 3 × s n (referred to as “3σ”) can be devised. An illustration is given in figure 2. The same spectrum as in the previous section, with 3σ error bars, is shown. The size of error bars decreases with the number of histories. This is in agreement with the qualitative discussion of Fig. 1 given in the previous section. The noise on the overall spectrum can be estimated by summing the variance estimator over the frequencies (note that s n is a function of ω). Consider the following definition

d 2 n =

n ω

k=1

s 2 nk ). (8)

(8)

Here, the sum is carried out over the discretized frequency points ω k , of number n ω . Figure 3 shows a plot of d n in terms of n, assuming n = 10, 100, and 1000, with a total number of points n ω = 500. As can be seen in the figure, the noise on the overall spectrum decreases with n. The obtained values of d n (13.9, 4.33, and 1.39) are well fitted by a 1/

n law, which is in agreement with the variance estimator formula (7).

4. Comparison of the two formulas

The formula based on direct power spectrum evaluation [Eq. (1)] and the formula involving the dipole autocorrelation function [Eq. (2)] yield the same result at the n → ∞ limit, they converge with the same speed ( 1/

n), but the variances are different. The direct power spectrum evaluation involves a temporal averaging process, which is not performed if Eq. (2) is used.

This point is made apparent if one notes that Eq. (1) can be rewritten in a way formally equivalent to Eq. (2) in terms of the following autocorrelation function

C 1 (t) = lim

T →∞

1 T

T

0

dt Tr[ρ ⃗ d(t ) · d(t + t)] , (9) instead of that used in Eq. (2), namely,

C 2 (t) = Tr[ρ ⃗ d(0) · d(t)] . (10) This way of rewriting Eq. (1) was already noticed in Ref. [19]. The temporal averaging process in Eq. (9) provides better statistics on the calculated spectrum. We have performed simulations with the same conditions as above, using Eq. (1); the obtained spectra are indeed less noisy (see Fig. 4).

An analysis of the noise relative to the autocorrelation functions (9) and (10) can be done in analytical terms through the classical oscillator model of the atom. In this model, perturbations are accounted for through sudden interruptions, occurring at random times following a Poisson process. After each interruption, a new random phase is generated uniformly in the [0, 2π[

interval and a new frequency is also generated according to some prescribed probability density function. Denoting z the coordinate of the oscillator, we have

z(t) =

j 0

e i(Ω j t+φ j ) θ(t t j )θ(t j+1 t), (11)

(9)

3315 3320 3325 -0.4

0.0 0.4 0.8 1.2 1.6

Lineshape(arb.units)

Ar XVIII Ly-

D plasma

N e,i

= 2x10 23

cm -3

T e,i

= 1 keV

10 histories

Energy (eV) (a)

3315 3320 3325

0.0 0.2 0.4 0.6 0.8 1.0 1.2

Lineshape(arb.units)

Energy (eV) (b)

Ar XVIII Ly-

D plasma

N e,i

= 2x10 23

cm -3

T e,i

= 1 keV

1000 histories

Figure 2: Plot of the Ar XVIII Lyman α line calculated with the same conditions as in

Fig. 1, with error bars at 3σ. The amplitude of error bars decreases with the number of

histories.

(10)

10 100 1000 1

10

d n

(arb.units)

n

Figure 3: The decrease of noise level on the overall spectrum with n can be quantified by summing the variance estimator over the frequencies. Here, the plotted total deviation d n

(see text for definition) corresponds to a calculated spectrum with 500 points. The shape of this curve is well fitted by a 1/

n law.

(11)

3315 3320 3325 0.0

0.4 0.8 1.2

Lineshape(arb.units)

Energy (eV) Eq. (1), 10 histories Ar XVIII Ly-

D plasma

N e,i

= 2x10 23

cm -3

T e,i

= 1 keV

Figure 4: The spectrum obtained using direct evaluation of the power spectrum, according to Eq. (1), is less noisy than that obtained from the relation (2) that involves the dipole autocorrelation function. Here, the plot corresponds to an Ar XVIII Lyman α line shape with the same conditions as above, obtained using Eq. (1) and assuming 10 histories.

Error bars at each frequency are smaller than those in Fig. 2(a).

(12)

where θ stands for the Heaviside step function, the times t j are generated according to the Poisson distribution

p(t j ) = θ(t j ) γ j t j j 1

(j 1)! e γt j , (12)

and φ j and Ω j are the oscillator’s phase and frequency between t j and t j+1 ; they are assumed to be independent of each other. In Eq. (12), γ is a rate denoting the number of collisions per unit time that contribute to the phase destruction. The oscillation amplitude has been set equal to 1 for simplicity sake. The autocorrelation functions defined by Eqs. (9) and (10) become

C 1 (t) = lim

T →∞

1 T

T 0

dt z (t )z(t + t) , (13) and

C 2 (t) = z (0)z(t) , (14) respectively, and they are both identical to

θ(t) e iΩ 0 t e γt . (15) The impact and static limits correspond to γ → ∞ and γ 0, respectively.

The noise level is estimated by introducing the second moments 1

T

T 0

dt 1 T

T 0

dt ′′ z(t )z (t + t)z (t ′′ )z(t ′′ + t) , (16)

z(0)z (t)z (0)z(t) , (17) and their related variances σ 2 1 and σ 2 2 , obtained by subtracting the square module of Eq. (15) from Eqs. (16) and (17). An explicit calculation yields (details are given in the appendix)

σ 1 2

T →∞

c

T , (18)

and

σ 2 2 = 1, (19)

where τ c = Re ∫

0 dt e iΩ 0 t e γt is the characteristic dipole decorrelation

time. The first variance is smaller than the second one, which indicates that

the noise level is smaller. This trend is in agreement with fully numerical

simulations.

(13)

5. Conclusion

The characterization of laboratory and astrophysical plasmas involves passive spectroscopy diagnostics. In this work, we have reexamined the nu- merical simulation method utilized in Stark broadening calculations and we have addressed the issue of statistical noise quantification. Using an estima- tor for the variance, we have shown that the noise decreases with the number of statistical realizations n and follows the power law 1/

n. Two formulas, one based on a direct calculation of the power spectrum and the other one involving an evaluation of the dipole autocorrelation function as intermediary step, have been examined in detail. Both yield the same result if a large num- ber of realizations is assumed, but the noise levels they provide are different.

Using an analytical model, we have shown that the direct evaluation of the power spectrum yields a result less noisy. This trend is in agreement with fully numerical simulations. Presently, a drawback of numerical simulations is that runs are CPU time consuming, especially if a complex atomic system, with a large number of energy levels, is considered. The design of routines that allow fast evaluation of spectra is a major issue. The method based on direct power spectrum evaluation and discussed here enter the class of the so-called variation reduction techniques. Work is presently underway in order to quantify the total computational cost (including numerical Fourier transform, matrix product etc.) involved in numerical simulations. Com- parisons to other line shape codes will be performed and confrontations to experimental spectra will also be done. The simulation methods right now are limited to classical-path treatments, which means that they cannot take into account any back-reaction of the plasma to the radiator [26], as well as any quantum mechanical effects, such as recombination [27]. An extension of this investigation to fully quantum mechanical simulations, i.e., where both the emitter and the perturbers are treated quantum mechanically, is foreseen. Techniques specific to condensed matter physics (such as quantum Monte Carlo [28]), designed to tackle quantum many-body problems, will be addressed and examined in detail.

Acknowledgements

This work has been carried out within the framework of the French Re-

search Federation for Magnetic Fusion Studies.

(14)

Appendix A.

The power spectrum formula (1) that serves as a basis in line shape calculations stems from the energy conservation relation in basic electromag- netism. According to the Poynting theorem, the rate of work delivered by an atom to the electromagnetic field is given by

P = lim

T →∞

1 T

T /2

−T /2

dt

d 3 r ⃗j(⃗ r, t) · E(⃗ r, t) , (A.1) where ⃗j(⃗ r, t) = d(t)δ(⃗ ˙ r) is the current density associated with the atomic dipole and E is the electric part of the radiation field. It is assumed that the atom is motionless, is located at r = 0, and has negligible spatial extent.

The brackets ... denote a statistical average over the perturbers. Equa- tion (A.1) holds both classically and quantum mechanically. In quantum mechanics, ⃗j and E are operators expressed within the Heisenberg picture and written in normal order [29]. At the lowest interaction order (weak cou- pling approximation), ⃗j is assumed independent of E (no radiation feedback is considered) and E is expressed as a linear function of ⃗j, according to the following relation [30]

E(⃗ r, t) = 1 ε 0 c 2

d 3 r

dt G(⃗ r r , t t ) ⃗j(⃗ ˙ r , t ). (A.2) This relation is obtained by solving the Maxwell equation set. The Green function G is defined by

G(⃗ r, t) = δ(t r/c)

4πr . (A.3)

In practice, since the atom’s spatial extent is neglected, only the first-order expansion in r of the delta function contributes to the integrals in Eqs. (A.1) and (A.2); the following substitution holds

G(⃗ r, t) ↔ − δ (t)

4πc . (A.4)

Inserting this expression into Eq. (A.2), and putting the resulting field in Eq. (A.1), one gets the following relation

P = 1

6πε 0 c 3 lim

T →∞

1 T

T /2

T /2

dt

d(t) ¨ 2

. (A.5)

(15)

The normal order prescription leads to formally perform the substitution d(t) ¨ 2 2 d ¨ ( ) (t) · d ¨ (+) (t) = 2 d ¨ (+) (t) 2 , where the plus and minus exponents denote positive and negative frequency contributions. The contribution of various atomic states to the spectrum is retained through the substitution

d ¨ (+) (t) 2

αβ

p α

β | d(t) ¨ | α 2

. Defining the spectrum P (ω) through the identity P = ∫

dωP (ω) and using the Parseval theorem, one obtains equation (1).

Appendix B.

We estimate the four-time averages in Eqs. (16) and (17) by separating them into product of two-time averages; this property strictly holds in the case of Gaussian process. The average z(t )z (t +t)z (t ′′ )z(t ′′ + t) factorizes as

z(t )z (t + t)z (t ′′ )z(t ′′ + t) = z(t )z (t + t) ⟩⟨ z (t ′′ )z(t ′′ + t) + z(t )z (t ′′ ) ⟩⟨ z (t + t)z(t ′′ + t) + z(t )z(t ′′ + t) ⟩⟨ z (t + t)z (t ′′ ) (B.1) . The first and second terms are identical to the square module of C 1,2 (t), C 1,2 (t t ′′ ), respectively, and the last term is identically zero. The first variance reads

σ 1 2 = 1 T

T 0

dt 1 T

T 0

dt ′′ C 1 (t t ′′ ). (B.2) It reduces to a single integral at the T → ∞ limit and equation (18) is obtained. The second variance is obtained directly by formally setting t = 0 = t ′′ in Eq. (B.1).

References

[1] H. R. Griem, Spectral Line Broadening by Plasmas, Academic Press, New York, 1974.

[2] R. Stamm, D. Voslamber, On the role of ion dynamics in the Stark

broadening of hydrogen lines, J. Quant. Spectrosc. Radiat. Transfer 22

(1979) 599.

(16)

[3] R. Stamm, Y. Botzanowski, V. P. Kaftandjian, B. Talin, E. W. Smith, Ion-Dynamics Effect on Hydrogenic Stark Profiles in Hot and Dense Plasmas, Phys. Rev. Lett. 52 (1984) 2217.

[4] M. A. Gigosos, J. Fraile, F. Torres, Hydrogen Stark profiles: A simulation-oriented mathematical simplification, Phys. Rev. A 31 (1985) 3509.

[5] M. Baranger, Problem of Overlapping Lines in the Theory of Pressure Broadening, Phys. Rev. 111 (1958) 494.

[6] A. C. Kolb, H. Griem, Theory of Line Broadening in Multiplet Spectra, Phys. Rev. 111 (1958) 514.

[7] A. Brissaud, U. Frisch, Theory of Stark broadening-II Exact line profile with model microfield, J. Quant. Spectrosc. Radiat. Transfer 11 (1971) 1767.

[8] C. Stehl´ e, Stark broadening of hydrogen Lyman and Balmer in the con- ditions of stellar envelopes, Astron. Astrophys. Suppl. Ser. 104 (1994) 509.

[9] B. Talin, A. Calisti, L. Godbert, R. Stamm, R. W. Lee, L. Klein, Frequency-fluctuation model for line-shape calculations in plasma spec- troscopy, Phys. Rev. A 51 (1995) 1918.

[10] A. Calisti, C. Moss´ e, S. Ferri, B. Talin, F. Rosmej, L. A. Bureyeva, V. S.

Lisitsa, Dynamic Stark broadening as the Dicke narrowing effect, Phys.

Rev. E 81 (2010) 016406.

[11] E. Stambulchik, Y. Maron, Quasicontiguous frequency-fluctuation model for calculation of hydrogen and hydrogenlike Stark-broadened line shapes in plasmas, Phys. Rev. E 87 (2013) 053108.

[12] J. Rosato, Y. Marandet, H. Capes, S. Ferri, C. Moss´ e, L. Godbert- Mouret, M. Koubiti, R. Stamm, Stark broadening of hydrogen lines in low-density magnetized plasmas, Phys. Rev. E 79 (2009) 046408.

[13] E. Stambulchik, Review of the 1st Spectral Line Shapes in Plasmas code

comparison workshop, High Energy Density Phys. 9 (2013) 528.

(17)

[14] S. Ferri, A. Calisti, C. Moss´ e, J. Rosato, B. Talin, S. Alexiou, M. A.

Gigosos, M. A. Gonz´ alez, D. Gonz´ alez-Herrero, N. Lara, T. Gomez, C. Iglesias, S. Lorenzen, R. C. Mancini, E. Stambulchik, Ion Dynamics Effect on Stark-Broadened Line Shapes: A Cross-Comparison of Various Models, Atoms 2 (2014) 299.

[15] E. Stambulchik, A. V. Demura, Dynamic Stark broadening of Lyman-α, J. Phys. B: At. Mol. Opt. Phys. 49 (2016) 035701.

[16] J. Rosato, Report on the third SLSP code comparison workshop, High Energy Density Phys. 22 (2017) 60.

[17] E. Stambulchik, A. Calisti, H.-K. Chung, M. A. Gonz´ alez, Special Issue on Spectral Line Shapes in Plasmas, Atoms 2 (2014) 378.

[18] E. Stambulchik, A. Calisti, H.-K. Chung, M. A. Gonz´ alez, Spectral Line Shapes in Plasmas II, Atoms 7 (2019) 20.

[19] E. Stambulchik, S. Alexiou, H. R. Griem, P. C. Kepple, Stark broadening of high principal quantum number hydrogen Balmer lines in low-density laboratory plasmas, Phys. Rev. E 75 (2007) 016401.

[20] L. I. Schiff, Quantum Mechanics, McGraw-Hill, New York, 1949.

[21] M. Baranger, Atomic and Molecular Processes, Academic Press, New York, 1962, Ch. Spectral Line Broadening in Plasmas.

[22] S. Alexiou, Overview of plasma line broadening, High Energy Density Phys. 5 (2009) 225.

[23] J. Halenka, W. Olchawa, J. Madej, B. Grabowski, Pressure shift and gravitational redshift of Balmer lines in white dwarfs: rediscussion, As- trophysical J. 808 (2015) 131.

[24] T. A. Gomez, T. Nagayama, D. P. Kilcrease, M. H. Montgomery, D. E.

Winget, Effect of higher-order multipole moments on the Stark line shape, Phys. Rev. A 94 (2016) 022501.

[25] J. Spanier, E. M. Gelbard, Monte Carlo Principles and Neutron Trans-

port Problems, Dover, Mineola, 2008.

(18)

[26] E. W. Smith, C. R. Vidal, J. Cooper, Classical Path Methods in Line Broadening. I. The Classical Path Approximation, Journal of Research of the National Bureau of Standards - A. Physics and Chemistry 73A (1969) 389.

[27] T. A. Gomez, T. Nagayama, C. J. Fontes, D. P. Kilcrease, S. B. Hansen, M. C. Zammit, D. V. Fursa, A. S. Kadyrov, I. Bray, Effect of Electron Capture on Spectral Line Broadening in Hot Dense Plasmas, Phys. Rev.

Lett. 124 (2020) 055003.

[28] D. M. Ceperley, Microscopic simulations in physics, Rev. Mod. Phys. 71 (1999) S438.

[29] P. W. Milonni, The Quantum Vacuum An Introduction to Quantum Electrodynamics, Academic Press, San Diego, 1994.

[30] J. D. Jackson, Classical Electrodynamics, Wiley, Hoboken, 1999.

Références

Documents relatifs

merous measured and calculated Stark widths have been reported for neutral and singly ionized elements but no data are available for gallium and indium.. In this paper

is not purely repulsive and, therefore, that the blue molecular system observed at low density very close to the atomic line [71] J is correlated to the 1 u

We hereafter denote the lamp of the former type as ST-HCL(see-through hollow cathode lamp) and the latter as U-HCL(U-shaped cathode HCL). The experiment has shown that the shift

Abstract: The particular character of the welding arc working in pure argon, which emission spectrum consists of many spectral lines strongly broadened by Stark effect, has allowed

ing more similar to the Lorentz treatment of total coherence rup- ture. The second term fits the so-called impact approximation. It is obtained from the values of the dipole

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

3 and 4, show that intra-shell L-L transition (around 26 nm) are negligible compared to the Al IV lines (at 13.1 and 16 nm, M-L transitions) at about 8 eV electron temperature. At