Next Article in Journal
UDR-GS: Enhancing Underwater Dynamic Scene Reconstruction with Depth Regularization
Previous Article in Journal
MFC-RMA (Matrix Factorization and Constraints- Role Mining Algorithm): An Optimized Role Mining Algorithm
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

O4 -Symmetry-Based Non-Perturbative Analytical Calculations of the Effect of the Helical Trajectories of Electrons in Strongly Magnetized Plasmas on the Width of Hydrogen/Deuterium Spectral Lines

Physics Department, Auburn University, 380 Duncan Drive, Auburn, AL 36849, USA
Symmetry 2024, 16(8), 1009; https://doi.org/10.3390/sym16081009
Submission received: 5 June 2024 / Revised: 13 July 2024 / Accepted: 29 July 2024 / Published: 8 August 2024
(This article belongs to the Section Physics)

Abstract

:
The effects of the helical trajectories of the perturbing electrons in magnetized plasmas on the dynamical Stark width of hydrogen or deuterium spectral lines have been studied analytically in our previous two papers—specifically in the situation where the magnetic field B is so strong that the dynamical Stark width of these lines reduces to the so-called adiabatic Stark width because the so-called nonadiabatic Stark width is completely suppressed. This situation corresponds, for example, to DA and DBA white dwarfs. We obtained those analytical results by using the formalism of the so-called conventional (or standard) theory of the impact Stark broadening: namely, by performing calculations in the second order of the Dyson perturbation expansion. The primary outcome was that the dynamical Stark broadening was found to not depend on the magnetic field B (for sufficiently strong B). In the present paper, we use the O4 symmetry of hydrogen atoms for performing the corresponding non-perturbative analytical calculations equivalent to accounting for all orders of the Dyson perturbation expansion. The results, obtained by using the O4 symmetry of hydrogen atoms, differ from our previous ones not only quantitatively, but—most importantly—qualitatively. Namely, the dynamical Stark broadening does depend on the magnetic field B, even for strong B. These results should be important for revising the interpretation of the hydrogen Balmer lines observed in DA and DBA white dwarfs. We also address confusion in the literature on this subject.

1. Introduction

Among the analytical theories of the dynamical Stark broadening of hydrogen (or deuterium) spectral lines in plasmas, the most practically useful are the semiclassical theories that started from the pioneering works by Baranger [1,2]. These were then further developed in paper [3] into what was later called the conventional or standard theory, where it was assumed that the ion microfield was quasistatic (from the point of view of radiating atoms), while the electron microfield was dynamical and treated in the impact approximation. The primary feature of the impact approximation is treating the electron microfield as a sequence of completed binary collisions of the electrons with the radiating atom. The allowance for incomplete collisions was later made in papers [4,5,6], calling this version the unified theory.
In the intervening dozens of years, many further advances have been made, especially in treating the dynamical part of the ion microfield. We refer the readers to books [7,8,9,10,11,12,13,14,15,16], listed in chronological order, and the references therein.
The effects of the helical trajectories of the perturbing electrons in magnetized plasmas on the dynamical Stark width of hydrogen or deuterium spectral lines has been studied analytically for the first time in papers [17,18]—specifically in the situation where the magnetic field B is so strong that the dynamical Stark width of these lines reduces to the so-called adiabatic Stark width because the so-called nonadiabatic Stark width is completely suppressed. This situation corresponds, for example, to DA and DBA white dwarfs. The analytical results were obtained in papers [17,18] by using the formalism of the so-called conventional or standard theory of the impact Stark broadening [3,19]: namely, by performing calculations in the second order of the Dyson perturbation expansion. The primary outcome was that the dynamical Stark broadening was found not to depend on the magnetic field B (for sufficiently strong B).
In the present paper, we use the O4 symmetry of hydrogen atoms for performing the corresponding non-perturbative analytical calculations equivalent to accounting for all orders of the Dyson perturbation expansion. The results, obtained by using the O4 symmetry of hydrogen atoms, differ from those obtained in papers [17,18] not only quantitatively, but—most importantly—qualitatively: namely, the dynamical Stark broadening does depend on the magnetic field B, even for strong B.

2. Summary of Starting Formulas from Papers [17,18]

1. The radius vector R(t) of a perturbing electron traveling on a helical trajectory in a magnetized plasma is
R(t) = vztB/B + ρ[1 + (rBp/ρ) cos(ωBt + φ)] + ρ × B [rBp/(ρB)] sin(ωBt + φ).
In Equation (1), ρ is the impact parameter vector and ρ × B is its cross-product (vector product) with the magnetic field B; vz is the component of the electron velocity parallel to B (the z-axis is chosen parallel to B). Other notations are presented as follows:
rBp = vpB, ωB = eB/(mec).
In Equation (2), ωB is the Larmor frequency and vp is the component of the electron velocity perpendicular to B.
2. The electric field created by the perturbing electron at the location of the radiating hydrogen or deuterium atom:
E(t) = eR(t)/[R(t)]3.
The z-projection of this electric field is
Ez(t) = e(vzt)/[ρ2 + vz2t2 + vp2B2 + 2(ρvpB) cos(ωBt + φ)]3/2,
3. The nonadiabatic contribution to the dynamical Stark width, i.e., the contribution caused by the component of the field E(t) perpendicular to B, is completely suppressed if
B > Bcr = 2Te/(3|Xαβ|eλc) = 9.2 × 102Te(eV)/|Xαβ| Tesla,
where λc = ħ/(mec) = 2.426 × 10−10 cm is the Compton wavelength of electrons. In Equation (5), Te is the electron temperature, and
Xαβ = |na(n1 − n2)α − nb(n1 − n2)β|.
In Equation (6), n is the principal quantum number, and n1 and n2 are the parabolic quantum numbers of the upper (aα) and lower (bβ) Stark states involved in the radiative transition.
4. The adiabatic part Φad of the electron broadening operator Φab (and the corresponding Stark width) is controlled by Ez(t), given by Equation (4).
5. For the helical trajectories of perturbing electrons, the starting formula for the adiabatic Stark width Γαβ = − Re[αβad)βα] for the line component, corresponding to the radiative transition between the upper Stark sublevel α and the lower Stark sublevel β, is
Γ α β , h e l = N e d v z f 1 v z 0 d v p f 2 v p v z 2 + v p 2 1 / 2 σ v z , v p .
In Equation (7), f2(vp) is the two-dimensional Maxwell distribution and f1(vz) is its one-dimensional counterpart. The operator σ(vz, vp) discussed below has the physical meaning of the cross-section of the “optical” collisions causing the dynamical Stark broadening of the spectral line.

3. Non-Perturbative Analytical Calculations

Let us start in a general way. We consider an atomic system described by the following Hamiltonian:
H(t) = H0 + V(t), V(t) = −dE(t).
In Equation (8), H0 is the unperturbed Hamiltonian, d is the electric dipole moment operator, and dE(t) is its scalar product (also known as the dot-product) with a time-dependent electric field E(t).
The central idea is to split the interaction V(t) into two parts
V(t) = V1(t) + V2(t)
and to rewrite the Hamiltonian from Equation (8) as follows:
H(t) = H1(t) + V1(t), H1(t) = H0 + V1(t),
where the extended “unperturbed” Hamiltonian H1(t) can be diagonalized analytically at any instant of time (at least, in a subspace of a fixed principal quantum number n). In particular, for hydrogenic atoms or ions, the analytical diagonalization of the extended “unperturbed” Hamiltonian H1(t) is possible due to the higher than geometrical symmetry (O4 symmetry) of the system described by H1(t).
We remind that one of the consequences of the O4 symmetry of hydrogenic atoms or ions is that the separation of variables is possible in several different types of coordinates: in the spherical coordinates, in the parabolic coordinates, and in the elliptical coordinates. For the present paper, the most convenient is the utilization of the parabolic coordinates, in which the projection dz of the dipole moment operator on the magnetic field can be diagonalized within any subspace of the fixed principal quantum number n.
We choose the z-axis along the magnetic field B and use the following specific breakdown of the interaction V(t) into two parts: the adiabatic part
V1(t) = −dzEz(t),
and the nonadiabatic part
V2(t) = −dperpEperp(t),
where dperp and Eperp are the components of the electric dipole moment and of the electric field, respectively, perpendicular to the magnetic field B. Under the condition (5), the nonadiabatic part of the interaction is completely suppressed, so that only the adiabatic part V1(t) remains.
We calculate the adiabatic contribution to the dynamical Stark broadening of hydrogenic spectral lines exactly, nonperturbatively (rather than in the second order of the Dyson perturbation expansion) [20,21,22]. This is achieved along the lines of the so-called old adiabatic theory—see, e.g., papers [23,24] and Appendix A of the present paper. In the formalism of the old adiabatic theory, the cross-section σ of the optical collisions has the following form:
σ = 2 π 0 d ϱ ϱ 1 cos dtd z E z ( t ) / ħ .
In Equation (13), the symbol <…> stands for the average over angular or phase variables; dz is the matrix element of the z-projection of the electric dipole moment:
dz/ħ = 3Xαβħ/(2mee).
The integral over time in Equation (13) vanishes for the odd part of Ez(t). Thus, it suffices to perform the integration only for the even part Ez,even(t) of Ez(t):
Ez(t)even = [Ez(t) + Ez(−t)]/2.
The latter was calculated in papers [17,18] as follows.
The integral over the impact parameter ρ was broken into two parts
σ = σ1 + σ2,
σ1 corresponding to the integral from ρ0 to infinity and σ2 corresponding to the integral from zero to ρ0, where
ρ0 = vpB.
For calculating σ1, Ez,even(t) was expanded in terms of the small parameter ρ0/ρ = vp/(ωBρ). After keeping the first non-vanishing term of the expansion, the following was obtained:
Ez(t)even = (sin φ) (3eρvpvzB) t[sin(ωBt)]/(ρ2 + vz2t2)5/2.
For calculating σ2, Ez,even(t) was expanded in terms of the small parameter ρ/ρ0 = ωBρ/vp. After keeping the first non-vanishing term of the expansion, the following was obtained:
Ez(t)even = (sin φ) (3eρvpvzB) t[sin(ωBt)]/(vp2B2 + vz2t2)5/2.
Using expressions (18) and (19), obtained in papers [17,18], we proceed now to calculating
σ 1 = 2 π ϱ 0 d ϱ ϱ 1 cos dtd z E z ( t ) even / ħ
and
σ 2 = 2 π 0 ϱ 0 d ϱ ϱ 1 cos dtd z E z ( t ) even / ħ .
For σ1, after calculating the integral over time, the expression within the symbol <…> in Equation (20) becomes
1 − cos[(sinφ)3XαβħvpωBK1Bρ/|vz|)/(me|vz|3)],
where K1(s) is the modified Bessel function of the 2nd kind. Then, the averaging of the expression (22) over φ yields
<…> = 1 − J0[3XαβħvpωBK1Bρ/|vz|)/(me|vz|3)],
where J0(u) is the Bessel function. For the range of ρ under consideration, the Bessel function K1Bρ/|vz|) is exponentially small, so that the argument of the Bessel function J0[…] is much smaller than unity. Therefore, Equation (23) can be simplified to the following:
<…> ≈ [3XαβħvpωBK1Bρ/|vz|)/(2me|vz|3)]2.
Then, the integration over impact parameters in Equation (20) yields the following:
σ1 ≈ (π/4)3/2[3Xαβħvp/(mevz2)]2 MeijerG[{{},{3/2}}, {{0, 0, 2},{}}, vp2/vz2],
where MeijerG[…] is the Meijer G-function.
Equation (24) for σ1 coincides with σ1 from papers [17,18]. It does not depend on the magnetic field B (provided that B satisfies the condition (5)). However, below it is shown that our nonperturbative result for σ2 significantly differs—both quantitatively and qualitatively—from the perturbatively obtained result for σ2 from papers [17,18]. In particular, our nonperturbative result depends on the magnetic field B, while the perturbative result from papers [17,18] did not depend on B.
For σ2, after calculating the integral over time, the expression within the symbol <…> in Equation (21) becomes
1 − cos[(sinφ)3XαβħvpωB2K1(vp/|vz|) ρ/(me|vz|3)].
Then, the averaging of the expression (26) over φ yields
<…> = 1 − J0[3XαβħωB2K1(vp/|vz|) ρ/(me|vz|3)].
The subsequent integration over impact parameters in Equation (27) leads to the following result for σ2:
σ2 = (πvp2B2)[1 − J1(w)/w], w = 3XαβħvpωBK1(vp/|vz|)/(me|vz|3).
Obviously, σ2 depends on ωB and thus on the magnetic field B.
For presenting the results in plots, we calculate the effective average values <vp> and <|vz|>, as well as their ratio, as follows (compare to Equation (7)):
< v p > = d v z f 1 v z 0 d v p f 2 v p v z 2 + v p 2 1 / 2 v p ,
< v p > = d v z f 1 v z 0 d v p f 2 v p v z 2 + v p 2 1 / 2 | v z | .
As a result, we find
<vp> = 1.60 <vT>, <|vz|> = 0.665 <vT>, <vp>/<|vz|> = 2.41,
where vT = (2Te/me)1/2 is the mean thermal velocity of plasma electrons. Then, we use these effective average values in Equation (25) for σ1 and in Equation (28) for σ2 in the subsequent plots where all quantities will be in atomic units.
Figure 1 shows the dependence of the ratio Γ/(NevT), where Γ is the adiabatic Stark width calculated by combining Equations (7), (25), and (28), on the Larmor frequency ωB and on the mean thermal velocity of electrons vT for Xαβ = 6, Xαβ being defined in Equation (9). We note that Xαβ = 6 corresponds to either na = 3, (n1 − n2)α = 2, nb = 1 (which is a component of the Ly-beta line), or to na = 4, (n1 − n2)α = 2, nb = 2, (n1 − n2)β = 1, (which is a component of the Balmer-beta line), or to na = 4, (n1 − n2)α = 3, nb = 3, (n1 − n2)β = 2, (which is a component of the Paschen-alpha line), or to na = 6, (n1 − n2)α = 1, nb = 1, (which is a component of Lyman-epsilon line), or to na = 6, (n1 − n2)α = 2, nb = 3, (n1 − n2)β = 2, (which is a component of the Paschen-gamma line), or to na = 4, (n1 − n2)α = 2, nb = 3, (n1 − n2)β = 2, (which is a component of the Paschen-alpha line), or to na = 6, (n1 − n2)α = 3, nb = 4, (n1 − n2)β = 3, (which is a component of the Brackett-beta line). Thus, the value of Xαβ = 6 represents Stark components of seven spectral lines, which is why we chose this value for Figure 1.
Figure 2 displays the dependence of the ratio Γ/(NevT) on the Larmor frequency ωB for the mean thermal velocity of electrons vT = 0.2 a.u., corresponding to 2.7 eV, for Xαβ = 6 (solid line). The corresponding perturbative result from papers [17,18] is shown by the dashed line.
From Figure 2, it is seen that the perturbative calculation from papers [17,18] overestimated the Stark width. It also illustrates that the perturbative calculation from papers [17,18] did not produce any dependence of the Stark width on the magnetic field, while the nonperturbative (more accurate calculation) demonstrates a significant dependence of the Stark width on the magnetic field.
The physical reason for the decrease in the dynamical Stark width as the magnetic field becomes stronger is the following. The dynamical Stark width is Γ~NevTσ, where σ is the cross-section of the optical collisions responsible for Stark broadening of hydrogen or deuterium lines [9]. In its turn, σ~(vT/Ω)2, where Ω is the characteristic frequency of the variation of the electric field of the perturbing ions at the location of the radiating atom. Thus, the dynamical Stark width Γ is inversely proportional to Ω2. In non or weakly magnetized plasmas, Ω~ωW~Te/(n2ħ), where ωW is the Weisskopf frequency [24]. In strongly magnetized plasmas, as the Larmor frequency ωB starts exceeding ωW, then the characteristic frequency of the variation of the electric field of the perturbing ions at the location of the radiating atom becomes Ω~ωB and the dynamical Stark width decreases with the growth of the magnetic field.
As for the dependence of the dynamical Stark width Γ on the mean thermal velocity of plasma electrons vT, physically it is the competition of the following two factors. On the one hand, at vT = 0, the dynamical Stark width vanishes because it is dynamical, while there no dynamics at vT = 0. On the other hand, at relatively large vT, the dynamical Stark width Γ decreases as vT increases. This is because at relatively large vT, so that the Weisskopf frequency exceeds the Larmor frequency, the dynamical Stark width Γ, being proportional to vTσ(vT), becomes inversely proportional to vT. Consequently, somewhere between vT = 0 and relatively large vT, Γ reaches a maximum, as it is seen in Figure 1 at a fixed ωB.

4. Discussion of the Results

We showed how far the analytical semiclassical theory of the dynamical Stark broadening of hydrogen or deuterium lines in plasmas has progressed. While the pioneering works [1,2,3] presented the most simple theory, lots of subsequent improvements were made in the intervening dozens of years, as reflected, e.g., in works [4,5,6,7,8,9,10,11,12,13,14,15,16]. However, it was not until the year 2017 that the effect of helical rather than rectilinear trajectories of perturbing electrons in strongly magnetized plasmas on the dynamical Stark width of hydrogen or deuterium spectral lines was studied analytically for the first time, in paper [17].
Specifically, in papers [17,18], the situation was studied where the magnetic field B is so strong that the dynamical Stark width of these lines reduces to the so-called adiabatic Stark width because the so-called nonadiabatic Stark width is completely suppressed. The analytical calculations in papers [17,18] were perturbative: they were performed in the second order of the Dyson perturbation expansion.
In distinction, in the present paper we used the O4 symmetry of hydrogen or deuterium atoms for performing non-perturbative analytical calculations equivalent to accounting for all orders of the Dyson perturbation expansion in the spirit of the generalized theory of the dynamical Stark broadening [20,21,22].
The results of the present paper differ from those obtained in papers [17,18], not only quantitatively, but—most importantly—qualitatively. Namely, the dynamical Stark broadening does depend on the magnetic field B even for strong B, while in papers [17,18] the dynamical Stark broadening did not depend on B. In addition to this, the non-perturbative calculations yield the dynamical Stark width by about an order of magnitude smaller than the perturbative calculations from papers [17,18], as seen, e.g., in Figure 2.
The results of the present paper should be important for revising the interpretation of the hydrogen Balmer lines observed from DA and DBA white dwarfs. According to observations, the magnetic field in plasmas of white dwarfs can range from 103 Tesla to 105 Tesla (see, e.g., papers [25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42], listed in chronological order and the references therein), thus easily exceeding the critical value from Equation (5). Indeed, for the typical electron temperature Te~1 eV in the white dwarfs plasma emitting hydrogen lines, Equation (5) yields Bcr~103/|Xαβ| Tesla < 103 Tesla.
Finally, we note that in the literature there is a confusion concerning the effect that the helical trajectories of the perturbing electrons have on the dynamical Stark width of hydrogen spectral lines. We clarify this confusion in Appendix B of the present paper.

Funding

This research received no external funding.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The author declares no conflicts of interest.

Appendix A. Some Details on the Old Adiabatic Theory of the Stark Broadening of Spectral Lines in Plasmas

The adiabatic line broadening theory proceeds from the assumption that the electric field E(t) of the perturbing charged particles causes only the phase modulation of the atomic oscillator (i.e., a change in the phase of the wave function of the radiating atom)—see, e.g., review [34]. Then, the correlation function has the following form:
K ( τ ) = < e x p i 0 τ κ ( t ) d t > .
In Equation (A1),
κ(t) = CE(t)/e,
where C is the Stark constant of the line component in the parabolic quantization.
Next, it is assumed that the phase change κ(t) is the sum of the phase changes κj(t) due to the individual independent perturbing particles: j = 1, 2, …, N. Then, the correlation function factorizes into the product of the corresponding contributions from the individual particles:
K ( τ ) = < exp i j = 1 N 0 τ κ j ( t ) d t > N = < Π j = 1 N e x p i 0 τ κ j ( t ) d t > N = < e x p i 0 τ κ ( t ) d t > 1 N .
In Equation (A3), the symbol <…>N stands for the average over the phase volume of all particles, while the symbol <…>1 stands for the average over the phase volume of one particle. The latter average incorporates the integration with respect to the coordinates dr0/V and with respect to the velocity distribution f(v)dv of the perturbing particles (V being the spatial volume):
K ( τ ) = 1 ( 1 / V ) d r 0 f ( v ) d v 1 e x p i 0 τ κ ( t ) d t N = N p V ,
where Np is the density of the perturbing particles.
In the limit where N → ∞, V → ∞, but Np = N/V = const, one obtains
K(τ) = exp{−NpV(τ)],
where V(τ) is the so-called “collision volume:
R e [ V ( τ ) ] = d r f ( v ) 0 2 π ϱ d ϱ d z < 1 cos [ d t C E z ( t ) ] > .
In Equation (A6), the transformation to the cylindrical coordinates dr0 = 2πρdρdz has been made.
For the relatively large time τ >> Ω, where Ω is the characteristic frequency of the variation of the electric field of the perturbing particle at the location of the radiating atom, one obtains
Re[V(τ)] ≈ (dV/dτ)|τ=∞ = vστ,
where σ is given by Equation (13).

Appendix B. Clarification of the Confusion in the Literature on This Subject

The analytical results from our paper [17] were obtained in frames of the so-called conventional (or standard) theory of the impact Stark broadening [3,19]. Despite this, in paper [43] it was incorrectly stated that the results in paper [1] were obtained in the frames of the so-called generalized theory of Stark broadening.
Most importantly, in paper [43] it was incorrectly stated that presumably in our paper [17] it was predicted analytically that the allowance for the helical trajectories of the perturbing electrons (HTPE) leads to a dramatic width increase in the lines Balmer-beta, Balmer-delta, and Balmer-epsilon at high densities, while simulations in paper [43] yielded a decrease in the corresponding widths. For supporting this statement, in paper [43] there were specific examples of the Balmer-beta, Balmer-delta, and Balmer-epsilon lines at the electron temperature Te = 1 eV and the electron density Ne = 2 × 1017 cm−3.
However, in reality, according to Equations (18) and (47), and Figure 1 from our paper [17], whether the allowance for HTPE increases or decreases the width of the Stark components of any hydrogen line depends on the value of the following dimensionless parameter
D = 5.57 × 10−11|Xαβ|[Ne(cm−3)]1/2/Te(eV),
where
Xαβ = naqα − nbqβ, qα = (n1 − n2)α, qα = (n1 − n2)α.
In Equation (A9), n is the principal quantum number, n1 and n2 are the parabolic quantum numbers (while q is often called the electric quantum number); we noted in paper [1] that Xαβ is the standard label of the Stark component of hydrogen or deuterium spectral lines corresponding to the radiative transition between the upper (α) and lower (β) Stark sublevels. According to Equation (47) and Figure 1 from paper [17], the allowance for HTPE increases the width of a Stark component if D > 0.44, but decreases its width if D < 0.44.
For the plasma parameters chosen in paper [43], Equation (A8) simplifies to
D = 0.025 |Xαβ|.
The critical value of D = 0.44 corresponds in this case to the critical value of |Xαβ| = 18, so that according to paper [1], the allowance for HTPE increases the width of Stark components having |Xαβ| > 18, but decreases the width of Stark components having |Xαβ| < 18.
For the Balmer-beta line, all intense Stark components have a |Xαβ| of no more than 10. So, the actual prediction from paper [17] for the Balmer-beta line at the plasma parameters chosen in paper [43] is the decrease in the Stark width (which can be also seen in Figures 2 and 3 from our paper [17]), rather than the increase in the Stark width (incorrectly stated in paper [43] with respect to the predictions from paper [17].
For the Balmer-delta line, the most intense Stark component has |Xαβ| = 6. So, based on the results from paper [17], for the plasma parameters chosen in paper [43], one should not expect the increase in the Stark width—contrary to the incorrect statement from paper [43] with respect to the predictions from our paper [17].
For the Balmer-epsilon line, the most intense Stark component has |Xαβ| = 14. So, again, based on the results from paper [17], for the plasma parameters chosen in paper [28], one should not expect the increase in the Stark width—contrary to the incorrect statement from paper [43] with respect to the predictions from our paper [17].
The statements from our paper [17] concerning the increase in the Stark width of the Balmer-delta and higher lines due to HTPE related to the electron densities Ne > 1018 cm−3, i.e., electron densities much higher than the only one value of Ne chosen in simulations from paper [43]. This situation is yet another demonstration of the superiority of analytical results over simulations: the analytical results are valid for a broad range of the electron density, while the simulations from paper [43] were performed for only one value of the electron density Ne = 2 × 1017 cm−3.

References

  1. Baranger, M. Simplified Quantum-Mechanical Theory of Pressure Broadening. Phys. Rev. 1958, 111, 481–493. [Google Scholar] [CrossRef]
  2. Baranger, M. Problem of Overlapping Lines in the Theory of Pressure Broadening. Phys. Rev. 1958, 111, 494–504. [Google Scholar] [CrossRef]
  3. Kolb, A.; Griem, H.R. Theory of Line Broadening in Multiplet Spectra. Phys. Rev. 1958, 111, 514–521. [Google Scholar] [CrossRef]
  4. Smith, E.; Cooper, J.; Vidal, C. Unified Classical-Path Treatment of Stark Broadening in Plasmas. Phys. Rev. 1969, 185, 140–151. [Google Scholar] [CrossRef]
  5. Vidal, C.; Cooper, J.; Smith, E. Hydrogen Stark Broadening Calculations with the Unified Classical Path Theory. J. Quant. Spectrosc. Radiat. Transf. 1970, 10, 1011–1063. [Google Scholar] [CrossRef]
  6. Vidal, C.; Cooper, J.; Smith, E. Unified Theory Calculations of Stark Broadened Hydrogen Lines Including Lower State Interactions. J. Quant. Spectrosc. Rad. Transf. 1971, 11, 263–281. [Google Scholar] [CrossRef]
  7. Griem, H.R. Plasma Spectroscopy; McGraw-Hill: New York, NY, USA, 1964. [Google Scholar]
  8. Griem, H.R. Spectral Line Broadening by Plasmas; Academic: New York, NY, USA, 1974. [Google Scholar]
  9. Oks, E. Plasma Spectroscopy. In The Influence of Microwave and Laser Fields; Springer: New York, NY, USA, 1995. [Google Scholar]
  10. Griem, H.R. Principles of Plasma Spectroscopy; Cambridge University Press: Cambridge, UK, 1997. [Google Scholar]
  11. Fujimoto, T. Plasma Spectroscopy; Clarendon: Oxford, UK, 2004. [Google Scholar]
  12. Oks, E. Stark Broadening of Hydrogen and Hydrogenlike Spectral Lines in Plasmas: The Physical Insight; Alpha Science International: Oxford, UK, 2006. [Google Scholar]
  13. Fujimoto, T.; Iwamae, A. Plasma Polarization Spectroscopy; Springer: Berlin/Heidelberg, Germany, 2008. [Google Scholar]
  14. Ochkin, V.N. Spectroscopy of Low Temperature Plasma; Wiley: Weinheim, Germany, 2009. [Google Scholar]
  15. Oks, E. Diagnostics of Laboratory and Astrophysical Plasmas Using Spectral Lines of One-, Two-, and Three-Electron Systems; World Scientific: Singapore, 2017. [Google Scholar]
  16. Oks, E. Advances in X-ray Spectroscopy of Laser Plasmas; IOP Publishing: Bristol, UK, 2020. [Google Scholar]
  17. Oks, E. Effect of Helical Trajectories of Electrons in Strongly Magnetized Plasmas on the Width of Hydrogen/Deuterium Spectral Lines: Analytical Results and Applications to White Dwarfs. Intern. Rev. At. Mol. Phys. 2017, 8, 61–72. [Google Scholar]
  18. Oks, E. Review of Recent Advances in the Analytical Theory of Stark Broadening of Hydrogenic Spectral Lines in Plasmas: Applications to Laboratory Discharges and Astrophysical Objects. Atoms 2018, 6, 50, Erratum in Atoms 2019, 7, 68. [Google Scholar] [CrossRef]
  19. Kepple, P.; Griem, H.R. Improved Stark Profile Calculations for the Hydrogen Lines Hα, Hβ, Hγ and Hδ. Phys. Rev. 1968, 173, 317. [Google Scholar] [CrossRef]
  20. Ispolatov, Y.; Oks, E. A Convergent Theory of Stark Broadening of Hydrogen Lines in Dense Plasmas. J. Quant. Spectr. Rad. Transf. 1994, 51, 129–138. [Google Scholar] [CrossRef]
  21. Oks, E.; Derevianko, A.; Ispolatov, Y. A Generalized Theory of Stark Broadening of Hydrogen-Like Spectral Lines in Dense Plasmas. J. Quant. Spectr. Rad. Transf. 1995, 54, 307–316. [Google Scholar] [CrossRef]
  22. Oks, E. Stark Widths of Hydrogen Spectral Lines in Plasmas: A Highly-Advanced Non-Simulative Semiclassical Theory and Tables. AIP Conf. Proc. 2006, 874, 19–34. [Google Scholar]
  23. Anderson, P.W. A Method of Synthesis of the Statistical and Impact Theories of Pressure Broadening. Phys. Rev. 1952, 86, 809. [Google Scholar] [CrossRef]
  24. Lisitsa, V.S. Stark Broadening of Hydrogen Lines in Plasmas. Sov. Phys. Uspekhi 1977, 122, 603–630. [Google Scholar] [CrossRef]
  25. Angel, J.R.P. Magnetic White Dwarfs. Ann. Rev. Astron. Astrophys. 1978, 16, 487–519. [Google Scholar] [CrossRef]
  26. Angel, J.R.P.; Liebert, J.; Stockman, H.S. The Optical Spectrum of Hydrogen at 160–350 Million Gauss in the White Dwarf GRW +70 deg 8247. Astrophys. J. 1985, 292, 260–266. [Google Scholar] [CrossRef]
  27. Wunner, G.; Roesner, W.; Herold, H.; Ruder, H. Stationary Hydrogen Lines in White Dwarf Magnetic Fields and the Spectrum of the Magnetic Degenerate GRW +70 8247. Astron. Astrophys. 1985, 149, 102–108. [Google Scholar]
  28. Henry, R.J.W.; Oconnell, R.F. Hydrogen Spectrum in Magnetic White Dwarfs: Hα, Hβ, and Hγ Transitions. Publ. Astron. Soc. Pac. 1985, 97, 333–339. [Google Scholar] [CrossRef]
  29. Bergeron, P.; Wesemael, F.; Fontaine, G.; Liebert, J. On the Surface Composition of Cool, Hydrogen-Line White Dwarfs: Discovery of Helium in the Atmospheres of Cool DA Stars and Evidence for Convective Mixing. Astrophys. J. 1990, 351, L21–L24. [Google Scholar] [CrossRef]
  30. MacDonald, J.; Vennes, S. How Much Hydrogen is There in a White Dwarf? Astrophys. J. 1991, 371, 719–738. [Google Scholar] [CrossRef]
  31. Bergeron, P.; Saffer, R.A.; Liebert, J. A Spectroscopic Determination of the Mass Distribution of DA White Dwarfs. Astrophys. J. 1992, 394, 228–247. [Google Scholar] [CrossRef]
  32. Fassbinder, P.; Schweizer, W. Stationary Hydrogen Lines in Magnetic and Electric Fields of White Dwarf Stars. Astron. Astrophys. 1996, 314, 700–706. [Google Scholar]
  33. Reimers, D.; Jordan, S.; Koester, D.; Bade, N. Discovery of Four White Dwarfs with Strong Magnetic Fields by the Hamburg/ESO Survey. Astron. Astrophys. 1996, 311, 572. [Google Scholar]
  34. Kowalski, P.M.; Saumon, D. Found: The Missing Blue Opacity in Atmosphere Models of Cool Hydrogen White Dwarfs. Astrophys. J. 2006, 651, L137–L140. [Google Scholar] [CrossRef]
  35. Kawka, A.; Vennes, S.; Schmidt, G.D.; Wickramasinghe, D.T.; Koch, R. Spectropolarimetric Survey of Hydrogen-Rich White Dwarfs Stars. Astrophys. J. 2007, 654, 499–520. [Google Scholar] [CrossRef]
  36. Tremblay, P.-E.; Bergeron, P. Spectroscopic Analysis of DA White Dwarfs: Stark Broadening of Hydrogen Lines Including Nonideal Effects. Astrophys. J. 2009, 696, 1755–1770. [Google Scholar] [CrossRef]
  37. Külebi, B.; Jordan, S.; Euchner, F.; Gänsicke, B.T.; Hirsch, H. Analysis of Hydrogen-Rich Magnetic White Dwarfs Detected in the Sloan Digital Sky Survey. Astron. Astrophys. 2009, 506, 1341–1350. [Google Scholar] [CrossRef]
  38. Gianninas, A.; Bergeron, P.; Ruiz, M.T. A Spectroscopic Survey and Analysis of Bright, Hydrogen-Rich White Dwarfs. Astrophys. J. 2011, 743, 138. [Google Scholar] [CrossRef]
  39. Franzon, B.; Schramm, S. Effects of Magnetic Fields in White Dwarfs. J. Phys. Conf. Ser. 2017, 861, 012015. [Google Scholar] [CrossRef]
  40. Rolland, B.; Bergeron, P.; Fontaine, G. On the Spectral Evolution of Helium-atmosphere White Dwarfs Showing Traces of Hydrogen. Astrophys. J. 2018, 857, 56. [Google Scholar] [CrossRef]
  41. Bohlin, R.C.; Hubeny, I.; Rauch, T. New Grids of Pure-hydrogen White Dwarf NLTE Model Atmospheres and the HST/STIS Flux Calibration. Astronom. J. 2020, 160, 21. [Google Scholar] [CrossRef]
  42. Hardy, F.; Dufour, P.; Jordan, S. Spectrophotometric Analysis of Magnetic White Dwarf—I. Hydrogen-rich Compositions. Mon. Not. R. Astron. Soc. 2023, 520, 6111–6134. [Google Scholar] [CrossRef]
  43. Alexiou, S. Effects of Spiralling Trajectories on White Dwarf Spectra: High Rydberg States. Atoms 2023, 11, 141. [Google Scholar] [CrossRef]
Figure 1. Dependence of the ratio Γ/(NevT), where Γ is the adiabatic Stark width calculated by combining Equations (7), (25), and (28), on the Larmor frequency ωB and on the mean thermal velocity of electrons vT for Xαβ = 6, Xαβ being defined in Equation (9). All quantities are in atomic units.
Figure 1. Dependence of the ratio Γ/(NevT), where Γ is the adiabatic Stark width calculated by combining Equations (7), (25), and (28), on the Larmor frequency ωB and on the mean thermal velocity of electrons vT for Xαβ = 6, Xαβ being defined in Equation (9). All quantities are in atomic units.
Symmetry 16 01009 g001
Figure 2. Dependence of the ratio Γ/(NevT) on the Larmor frequency ωB for the mean thermal velocity of electrons vT = 0.2 a.u., corresponding to 2.7 eV, for Xαβ = 6 (solid line). The corresponding perturbative result from papers [1,2] is shown by the dashed line. All quantities are in atomic units.
Figure 2. Dependence of the ratio Γ/(NevT) on the Larmor frequency ωB for the mean thermal velocity of electrons vT = 0.2 a.u., corresponding to 2.7 eV, for Xαβ = 6 (solid line). The corresponding perturbative result from papers [1,2] is shown by the dashed line. All quantities are in atomic units.
Symmetry 16 01009 g002
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

Oks, E. O4 -Symmetry-Based Non-Perturbative Analytical Calculations of the Effect of the Helical Trajectories of Electrons in Strongly Magnetized Plasmas on the Width of Hydrogen/Deuterium Spectral Lines. Symmetry 2024, 16, 1009. https://doi.org/10.3390/sym16081009

AMA Style

Oks E. O4 -Symmetry-Based Non-Perturbative Analytical Calculations of the Effect of the Helical Trajectories of Electrons in Strongly Magnetized Plasmas on the Width of Hydrogen/Deuterium Spectral Lines. Symmetry. 2024; 16(8):1009. https://doi.org/10.3390/sym16081009

Chicago/Turabian Style

Oks, Eugene. 2024. "O4 -Symmetry-Based Non-Perturbative Analytical Calculations of the Effect of the Helical Trajectories of Electrons in Strongly Magnetized Plasmas on the Width of Hydrogen/Deuterium Spectral Lines" Symmetry 16, no. 8: 1009. https://doi.org/10.3390/sym16081009

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop