Stark Zeeman Line Shape Model For Multi Electron Radiators in Hot and Dense Plasmas Submitted To Large Magnetic Fields S FERRI

Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Stark-Zeeman line shape model for multi-electron

radiators in hot and dense plasmas submitted to large


magnetic fields
Sandrine Ferri, Olivier Peyrusse, Annette Calisti

To cite this version:


Sandrine Ferri, Olivier Peyrusse, Annette Calisti. Stark-Zeeman line shape model for multi-electron
radiators in hot and dense plasmas submitted to large magnetic fields. 2021. �hal-03404246�

HAL Id: hal-03404246


https://hal.science/hal-03404246
Preprint submitted on 26 Oct 2021

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
Stark-Zeeman line shape model for multi-electron radiators in hot and dense
plasmas submitted to large magnetic fields.

Sandrine Ferri,1 Olivier Peyrusse,1 and Annette Calisti1


Aix-Marseille Univ., CNRS, PIIM, UMR7345, Marseille, France.

(*Electronic mail: corresponding author: [email protected])

(Dated: 15 September 2021)

We present a Stark-Zeeman spectral line shape model and the associated numerical code,
PPPB, designed to provide fast and accurate line shapes for arbitrary atomic systems for
a large range of plasma conditions. PPPB is based on the coupling of the PPP code, a
Stark broadened spectral line shape code, developed for multi-electron ion spectroscopy
in hot and dense plasmas, and the MASCB code, recently developed to generate B-field
dependent atomic physics. The latter provides energy levels, statistical weights and re-
duced matrix elements of multi-electron radiators by diagonalizing the atomic Hamiltonian
which includes the well know B-dependent term. They are used as input in PPP working in
the standard line broadening approach, i.e. using the quasi-static ion and impact electron
approximations. The ion dynamics effects are introduced by the mean of the frequency
fluctuation model (FFM). The physical model of the electron broadening is based on the
semi- classical impact approximation including the effects of a strong collision term, of
interference and cyclotron motion. Finally, to account for polarization effects, the output
profiles are calculated for a given angle of observation with respect to the direction of the
magnetic field. The potential of such model is presented through Stark-Zeeman spectral
line shape calculations performed for various experimental conditions.

PACS numbers: Spectral line shapes modeling, Plasmas, Magnetic fields

1
I. INTRODUCTION

Measurements of magnetic fields are of importance in many studies of laboratory or space


plasmas. Among other methods, spectroscopic measurements are often used in plasmas to infer
temperatures from line intensity ratio and Doppler broadening, electron densities from the Stark
broadening and magnetic field strengths from distinct line shape features. The spectroscopy di-
agnostic techniques are based on the comparison between the observed and the modeled spectra.
Therefore, their reliable implementation requires accurate calculations of emission or absorption
spectra, which imply the use of analytic methods and computer codes of different complexity
and limit of applicability. Line shape modeling in plasmas has a long history1 but the existence
of intense magnetic fields in astrophysical objects (e.g. white dwarfs) and in various types of
plasmas created in laboratory (e.g., the magnetic- and inertial-confinement fusion devices) revives
the interest for atomic physics developments in such extreme conditions.

The presence of magnetic fields increases the complexity of line shape calculations in plasmas.
A magnetic field has three essential effects on Stark-broadened spectral lines: - a partial polar-
ization of the emitted light, - an additional splitting according to value of the magnetic quantum
number, m, and - the bending of the colliding charged particle trajectories into a helical path around
the magnetic lines of forces. These effects have been studied for several decades both theoretically
and experimentally since the initial work of N. Hoe and colleagues2 . Different methods have been
developed or have been extended to magnetized plasmas, such as numerical simulations3–6 , or
various theoretical models7–16 . Most of them are based on simplifying assumptions depending on
the relative importance of the Stark and Zeeman effects. A measure of this relative importance is
given by the ratio τ between the two respective average energy shifts2 . For example, for hydrogen,
3 1/3
with the normal electric field strength F0 = e/re2 and re = ( 4πNe
) , where e is the elementary
charge and Ne is the electron density, expressed in cm−3 , τ is given by

2/3
τ = 5.15 × 10−11 nNe /B, (1)

with n the principal quantum number and B the magnetic field strength, expressed in tesla. The
line profile coincides with the pure Stark profile if τ  1 and deviates progressively as τ decreases.
When τ ∼ 1, profiles broadened by the combined Stark-Zeeman effect are an intricate function of
Ne and B. Such cases are found for low−n hydrogen line series emitted in Tokamak edge regions

2
where Ne ∼ 1014 cm−3 , Te ∼ 1 eV and B ∼ few teslas, or in white dwarfs, where signature of in-
tense magnetic fields (over few hundred teslas) is observed at higher densities (Ne ∼ 1017 cm−3 ).
In laser-produced plasmas (100 eV < Te < 1 keV and 1021 < Ne < 1024 cm−3 ), high magnetic
fields over few hundreds teslas are generated and can strongly affect the emission of highly ion-
ized atoms17,18 . These conditions require a simultaneous treatment of Stark and Zeeman effects
on the line broadening.

The goal of this work is to present the main feature of the Stark-Zeeman line shape code PPPB
through various applications related to strongly magnetized plasmas.

II. ATOMIC PHYSICS IN PRESENCE OF MAGNETIC FIELDS

Atomic data necessary for a Stark-Zeeman calculation are usually generated by atomic physics
codes free of any external field. In practice, we use the Cowan Hartree-Fock atomic struc-
ture code19 , the Multi Configuration Dirac Fock code (MCDF)20 , the Flexible Atomic Code
(cFAC)21,22 or homemade codes for neutrals. In its original form, PPPB was designed to solve
the problem within the strong or the weak magnetic field approximations. The energy levels,
statistical weights and reduced dipole matrix elements were externally generated and neglecting
the quadratic contribution, the Zeeman contribution to the Hamiltonian was introduced in PPPB.

The Zeeman Hamiltonian reads

HZ = µB B(Lz + gs Sz ), (2)

where B is the magnitude of the magnetic field along the z direction, µB is the Bohr magneton,
gs = 2.0023192 is the anomalous gyromagnetic ratio for the electron spin, and Lz and Sz , the
projections of the total orbital and spin angular momenta of the atom, respectively. For sufficiently
weak B-field values, the off-diagonal matrix elements of HZ that connect states of different values
of the modulus, J, of the total angular momentum of the system, J~ =~L+~S, are negligible compared
to the contributions of the Coulomb and spin-orbit interactions. The contribution of the magnetic
field to the energy can be calculated as a perturbation. The following expression of the diagonal
matrix element of HZ for the state |γ J Mi was retained in PPPB

hγ J M| (Lz + gs Sz ) |γ J Mi = gγJ M, (3)

3
where gγJ is the Landé factor of the level γ J 19 . For magnetic fields sufficiently strong to disrupt
the coupling between orbital ~L and spin ~S momenta, Lz and Sz are easily evaluated for a state
|γ L S ml ms i. The strong field approximation through the following expression was then retained

hγ L S ml ms | (Lz + gs Sz ) |γ L S ml ms i = ml + gs ms . (4)

One can consider that the weak- and strong-field approaches are no longer valid when the mag-
netic field and the spin-orbit contributions are of the same order of magnitude. In this intermediate
case, the diagonalization of the whole Hamiltonian ξ~L · ~S + µB~B(~L + gs~S) is necessary.

An estimate of the critical B-field values, Bc (in tesla), for which both contributions have to be
considered at the same level in the Hamiltonian, is given for one-electron configuration23

4  2
Z∗ e2 me e4

Bc ∼ , (5)
n h̄c µB h̄2
with Z ∗ the effective charge, given by the Slater’s rule: Z ∗ = Z − σ , where σ is the screening
constant24 .

Table I shows the values of Bc for some H-like (Z ∗ = Z) and Li-like (Z ∗ = Z − 2σ1s , with
σ1s = 0.85) ions of interest. Radiative transitions from n = 2 are assumed.

Element Bc (tesla)

H-like Li-like

H (Z=1) 0.78 −−
C (Z=6) 103 270
Si (Z=14) 30 × 103 18 × 103
Ar (Z=18) 80 × 103 55 × 103

TABLE I. Critical value of the magnetic field (in tesla) for which the spin-orbit interaction is of the same
order of magnitude as the magnetic interaction. Estimations are made for Hydrogen-like and Lithium-like
ionization stages for elements of interest in this work.

To go beyond the weak and strong field approximations mentioned above, the atomic physics

4
code MASCB has been developed to generate B-field dependent atomic physics quantities.

MASCB follows the usual approach where the N-electron atomic Hamiltonian H is divided
into two parts. A separable part where electrons are supposed to be independent in a central po-
tential. The corrective part is treated using perturbation theory. In this framework, multi-electron
atomic states appear formally as a combination of Slater determinants. This means that, matrix
elements of H, J 2 and Jz must be calculated and the eigenvectors of these operators obtained after
diagonalization are the eigenstates of the atomic system. After an identification of the useful (i.e.
restricted, see for instance25,26 ) Slater determinant belonging to a set of configurations, matrix
elements are calculated using Condon rules summarized in classical textbooks24,27 . The restric-
tion of the Slater determinant set consists in considering just one possible value of M (eigenvalue
of Jz ) which is common to all states of a set of configurations. The subsequent calculations of
transition matrix elements coupling states of different M make use of the Wigner-Eckart theorem
and of operators J+ or J− when needed. After a sequential treatment of a list of configuration, it
is still possible to diagonalize the atomic Hamiltonian in the basis of the states belonging to all of
the considered configurations, which is just the so-called superposition of configurations method.

The description is non-relativistic, i.e. based on the Schrödinger equation, but it incorporates
the main relativistic corrections to the central field potential19 . This central potential, from which
mono-electronic energies and orbitals are obtained self-consistently, is built in the framework of
the Optimized Effective Potential method28,29 .

After the primary treatment of the isolated atom, one diagonalizes the part of the Hamiltonian
describing the interaction with the magnetic field. This part reads (in SI units)

e2 B2 N 2 2
HB = HZ + ri sin θi , (6)
8me ∑ i

where HZ is the linear Zeeman term (Eq.2), the second term being the quadratic term. Here ri , θi
are the usual polar coordinates of electron i in a system where the polar axis is taken along the B
axis, i.e. the z-axis30 .

It is worth noting that the quadratic term of HB introduces a mixing between states of different
n making the matrix of infinite order. Here the eigenvalues are necessarily approximated by those

5
of a truncated Hamiltonian matrix for HB . Truncated because this matrix is limited to a preselected
set of configurations. For a given value of the magnetic field, it is then necessary to check whether
the addition of more configurations introduces a significant change on the eigenvalues of interest.
For hydrogen or helium atoms this method has proven to yield sufficiently accurate results30 . The
figure 1 shows the effect of the quadratic terms on the Balmer series lines of hydrogen (up to H − ε
that corresponds to the radiative transition from n = 7 to n = 2)31,32 for plasma conditions relevant
to white dwarf atmospheres33 . One compares the Stark-Zeeman profiles obtained accounting or
not for the quadratic B-field effects for three magnetic field values: a) B = 100 tesla, b) B =
500 tesla and c) B = 1 ktesla. The calculations have been performed over the entire Balmer series
at once. Stark coupling between upper levels with different principal quantum numbers have
been accounted for. Note that our atomic structure calculation is performed in a configuration
interaction mode where the mixing is introduced by the quadratic term. This last point is crucial
for proper consideration of this term.

FIG. 1. Stark-Zeeman broadened Balmer series lines for Ne ∼ 1017 cm−3 , kB T = 5 eV and a) B = 100 tesla,
b) B = 500 tesla and c) B = 1 ktesla, without quadratic terms (dash) and with quadratic terms (solid). The
direction of observation is transversal to the B-field direction.

The quadratic terms give rise to additional structures on the line shapes and to a global shift
that increases as the principal quantum number increases. For higher magnetic fields, the Zeeman
components from different principal quantum numbers merge, so that the line series resemble as
complex set of indiscernible lines. Inferring the B-field from those lines is no longer possible.

6
In addition, the measure of the electron density by the mean of the Inglis-Teller limit34 can lead
to an overestimation of the electron density if Zeeman effects are not considered in the calculation.
As for example, the largest principal quantum number (PQN) given by the Inglis-Teller limit is
nearly equal to 7 for an electron density equal to 1017 cm−3 . The Stark-Zeeman Balmer series
verify this limit for a magnetic field of the order of 100 tesla. But, the lines emitted from large
quantum numbers start to merge as the B-field increases. For B = 500 tesla, the last resolved lines
correspond to the Hδ transitions, i.e. from the upper level n = 6. Using this level as the largest
PQN to infer the electron density would give Ne ∼ 2.7 × 1017 cm−3 instead of Ne = 1017 cm−3
as the calculation was performed. For B = 1 ktesla, the last resolved lines correspond to the Hγ
transitions, i.e. from the upper level n = 5. Hence the corresponding inferred electron density
would be Ne = 1018 cm−3 , an order of magnitude higher than the right one.

III. HELICAL TRAJECTORIES

Another difficulty to overcome in magnetized plasmas, is that, the charged particles follow
helical trajectories. The gyromotion of the electrons and ions around the magnetic lines of forces
may alter the dynamics of the plasma particles. Recent studies in strongly coupled magnetized
plasmas, i.e., when the Coulomb interaction exceeds the kinetic energy of the particles, have re-
ported a strong influence of the helical trajectories on the transport properties35–37 . A significantly
curved trajectory will also change the emitter-perturber interaction dynamics. Of particular current
interest are the electric microfield statistical properties in the case of strongly magnetized plasmas.
The Stark-Zeeman broadening mechanisms of spectral line shape could then be affected.

The effect of the helical trajectory on the emitter-perturber interaction can be estimated by the
q p
kB T
ratio of the Debye length λD = 4πN ee
to the Larmor radius r L = kB T mc2 /ZeB, with m, T and
Z the particle mass, temperature and charge, respectively. When the Larmor radius remains the
same order of the Debye length, the pertuber gyration occurs on the same time and length scales
as the Coulomb interaction. Above a critical B-field Bh , that corresponds to λD /rL = 1, the helical
trajectory has then to be accounted for. As illustration, values of the critical B-field for electrons
in magnetized plasmas of interest are summarized in Table II.

7
Magnetized plasmas Ne (cm−3 ) Bh (tesla)

Tokamak edge plasmas (B ∼ a few tesla) 1013 1


White Dwarf (B ∼ 100 T − 1 ktesla) 1017 100
Laser plasmas (B < 500 T ) 1020 3.2 × 103
Imploded targets (B ∼ a few ktesla) 1023 100 × 103

TABLE II. Columns list plasma of interest (with the magnetic field values commonly measured in the
concerned plasmas), electron density (Ne ) and critical B-field Bh for electrons.

Recent studies of the influence of B-fields on the electron trajectories in hydrogen plasmas have
been performed in the context of magnetic fusion and white dwarfs38–41 . It has been shown that
the introduction of helical trajectories reduces the characteristic duration of the perturbation to
the order of the inverse of the Larmor frequency, τL = 2π/ωL = 2πme c/eB. This results in a line
shape narrowing. Such results suggest a modification of the electron collision operator generally
used to describe the electronic Stark effect in line shape modeling, as we will see in section IV C.

Concerning the effects of helical trajectories on the interaction between the radiator and the
ionic perturbers, investigations of the statistical properties of the ionic microfields using classical
Molecular Dynamics simulations have been performed for short-pulse laser experiment conditions
where magnetic fields of a few thousands of tesla have been measured3 . It has been shown that
the modification of ionic field distribution function, W (F), by the magnetic field was negligible.
The same conclusion was exposed by C. Deutsch42 : "The low frequency component of the electric
microfields is seen to be rigorously unaffected by magnetic field in a thermal plasma", but, to
conclude: "in presence of a very strong magnetic field [ ] the slow electrons are to be added to
the ionic part of the low frequency component.". Very recent MD simulations43 , in the context of
high B-field generation using laser-plasmas interactions17 have shown that the helical trajectories
of the electrons along the B-line of forces may affect the distribution of the electrons around the
ions and thus may indirectly affect the ionic microfield distribution functions. Such results are
preliminary and more investigations have to be done.

8
IV. PPPB : A STARK-ZEEMAN LINE SHAPE CODE

The spectral line shape code PPPB has been designed to provide Stark-Zeeman broadened line
shapes for a wide range of density, temperature, and magnetic field values14 . It is based on the
PPP code, a Stark-broadened spectral line shape code44,45 , developed some years ago for multi-
electron ion spectroscopy in inertial confinement fusion plasmas46,47 .

Line shapes are usually modeled by working in the ”standard” quasi-static ion/impact electron
limit. The line shape function is given by
Z ∞
Is (ω) = W (Fi )I(ω, Fi )dFi , (7)
0

where, ω is the photon frequency and I(ω, Fi ) is the electron broadened line profile for a given
value of the microfield Fi following the static ion microfield distribution function, W (Fi ). The
field-dependent profile reads
1 n
† −1
o
I(ω, Fi ) = ReTr d [iω − iL(Fi ) + φ (ω)] ρd (8)
π
where L(Fi ) = [HZ (Fi ), I] is the Liouville operator associated to the B-field dependent Hamiltonian
of the emitter HZ , ρ is the emitter density operator, d is the emitter dipole operator and φ (ω) is
the electron broadening operator (see section IV C).

A peculiarity of Zeeman effect is a quantization axis imposed by the magnetic field. This
implies the emission to be polarized, following the selection rules for the dipole radiation, ∆J =
J 0 − J = 0, ±1 (J 0 = J = 0 not allowed), q = ∆M = M 0 − M = ±1 (σ± polarizations) and q =
∆M = 0 (π polarizations), assuming the magnetic field in the direction z. As the symmetry is
broken, the integration over the ionic microfield implies to consider the three directions of space
separately. One defines ~Fk and ~F⊥ the parallel and perpendicular ionic microfields to the direction
of the magnetic field, respectively. If θ is the angle between the magnetic and the electric fields,
p
then Fk = Fi µ and F⊥ = Fi 1 − µ 2 , with µ = cos θ . Therefore, the line profile given by Eq.(7) is
written as
1 ∞
Z +1 Z
Is,q (ω) = W (Fi ) Iq (ω, Fi , µ)dµdFi , (9)
2 0 −1
where Iq (ω, Fi , µ) represents the q−polarized line profile emitted by an ion in an external mag-
netic field and in a static ion field Fi .

9
The emission profile observed in the line of sight of the observer is given, by

I(ω, α) = Ik (ω)cos2 (α) + I⊥ (ω)sin2 (α), (10)

where α is the angle between the line of sight of the observer and the direction of ~B and

Ik (ω) = I+1 (ω) + I−1 (ω), (11)

1
I⊥ (ω) = I0 (ω) + (I+1 (ω) + I−1 (ω)), (12)
2
where I−1 (ω), I0 (ω) and I+1 (ω) are the q-polarized components given in Eq (9), q = −1, 0, +1
respectively.

The calculation of the integrant in Eq.(9) involves the inversion and the product of matrices in
the complex domain. In principle the calculation of has to be done for every static ion microfield
point and for every frequency point in the line shape, Is (ω). This is the most time-consuming
task in a line shape code. Hence, computer power and efficient algorithms are essential to make
complex emitter line shape calculations practical.

A. Static profile and Stark dressed transitions

The PPPB code performs block diagonalization of the resolvent in Eq.(8), afforded by the
selection rules of the atomic matrix elements. It performs an eigen-decomposition of the resolvent
in the { f , µ}-dependent bases (see ref.14,44,48 for more details)

h i−1
(2) (G)
Is,q (ω) = ∑ W f ∑ Wµ Im  dq† |M f ,µ ω1 − Ld ( f , µ) M −1
f ,µ |dq ρ0 , (13)
f µ

where M f ,µ is the matrix that diagonalizes the Liouville operator L: M −1 d


f ,µ L( f , µ)M f ,µ = L ( f , µ).
The integration over Fi and µ are replaced by i) a two-point integration weight, used for the sum-
mation over the discrete ionic field intensities f , and, ii) a Gauss-Legendre quadrature weight,
used for the angle summation49 . The number of microfields necessary to well describe the static
distribution function is n f ∼ 50 and a value of nµ ∼ 30 is enough to have a good convergence in
the angle summation.

10
In the PPPB code the static microfield distribution functions are either estimated from classical
Molecular Dynamics (MD) simulations50 , or by the Adjustable Parameter EXponential (APEX)
model51,52 . The latter is computationally fast and suited for weakly as well as strongly coupled
plasmas. Note to mention that, the Hooper static-field distribution function is used for the calcula-
tions involving neutral emitters53 . Accordingly, all those models do not depend on B and consider
the plasma surrounding the emitter isotrope.

This procedure leads to the concept of the Stark spectral components emitted by a set of
dressed two-level radiators: the Stark-dressed transitions (SDT). The SDT are characterized by
two complex numbers namely the generalized intensity aq,k + icq,k and the generalized frequency
fq,k + iγq,k .

The static q-polarized profile is then described by a sum of rational fractions which are gener-
alized Lorentzian spectral components of the line,

nq,k
ck (ω − fq,k ) + aq,k γq,k
Is,q (ω) = ∑ 2
(14)
k=1 (ω − fq,k )2 + γq,k

with nq,k = n f × nµ × ne × ng the number of SDT, that is also proportional to ng and ne the num-
ber of ground and upper selected energy levels, respectively, that define the studied atomic system.

B. The Frequency Fluctuation Model: stochastic mixing of the SDT

The quasi-static approximation is a useful approximation. But it is well known that a qua-
sistatic treatment of the ion perturbation can lead to large errors for plasma conditions that yield
substantial microfield fluctuations54 . Depending on the time scale of the line emission, the fluctu-
ations of the microfields, produced by the moving ions, have to be accounted for. This is the most
difficult part of the line broadening problem due to the stochastic behavior of the microfields55 .

The ion dynamics producing microfield fluctuations is modeled, in PPPB, by using the Fre-
quency Fluctuation Model (FFM)14,56 .

11
The FFM is based on the assumption that an atomic system perturbed by a fluctuating mi-
crofield behaves like a set of SDTs that are subject to a stationary Markov mixing process induced
by the field fluctuation. This results in an effective exchange between two-level transitions follow-
ing a Poisson process with the fluctuation rate of νi = vth /ri , with vth the ion thermal velocity and
ri = ( 43 πNi )−1/3 the mean distance between the ions, assuming a ionic density Ni .

Working in the Liouville space of the SDT, the Stark-Zeeman line shape accounting for the ion
dynamics and polarization is written as
1  −1
d
Id,q (ω) = Re ∑ i Dd,k ω1 − L − iΓ + iW Dq, j pq, j , (15)
π kj

with Ld the Liouville operator involving the transition frequencies of the SDT, Dq, j = rq
p
1 + ic j /a j
the matrix elements of the dipole moement for the SDT in the q polarization state (rq2 = ∑k aq,k )
and, pq, j = aq, j /rq2 is the instantaneous probability of state j in the q polarization state. Γ is
defined as the diagonal matrix of inverse lifetimes with Γk j = νi δk j . W is the matrix transitions
rates between different states, such as W = νi pq,k .

The particular form of W avoids matrix inversion. According to56 , defining the quasistatic
propagator
 −1
Gs (z) = z − iLd − iΓ (16)

which has only diagonal matrix elements, the total propagator can be written as

Gd (z) = Gs (z) − iGs (z) ·W · Gd (z). (17)

and introducing the previous expression in (15), we get


(aq,k +icq,k )/rq2
rq2 ∑k νi +γq,k +i(ω−ωq,k )
Id,q (ω) = Re aq,k /rq2
. (18)
π 1 − νi ∑k νi +γ
q,k +i(ω−ωq,k )

The above expression is used to calculate the Stark-Zeeman line shape along the line of sight
given by Eqs.(10), (11) and (12).

C. Electron broadening operator

The physical model of electron broadening used in PPPB is based on a semi-classical im-
pact approximation including the effects of a strong collision term57 , and of interference58 . It is

12
supposed that the emitter interacts with the plasma by binary collisions considering independent
pseudo-electrons. The Debye length represents an upper cutoff beyond which the electrons do not
collide. Moreover the pseudo-electrons move at constant velocities along straight trajectories. A
lower cutoff is introduced to avoid the Coulomb divergence at short distances. Using the pertur-
bation theory up to second-order in the emitter-electron interaction, the Maxwell-average operator
is given by r
4π 2me h̄ 2 ~ ~  
Φ(∆ω) = − Ne R · R Cn + G(∆ω) , (19)
3 πkB Te me
where ∆ω is the frequency detuning from the line center, ~R is the (emitter) electron position op-
erator operating in the subspace of principal quantum number n and Cn is the n-dependent strong
collision term, with C2 = 1.5, C3 = 1.0, C4 = 0.75, C5 = 0.5 and Cn = 0.4 for n > 5.

There are many ways to estimate G(∆ω) (see16 and references therein). We use the semi-
classical GBK model57
Z ∞ −x  h̄n2 2  ∆ω 2 + ω 2 
1 e c
G(∆ω) = dx with y ≈ (20)
2 y x 2Z EH kB Te
where the cutoff frequency, ωc , linked to the upper cutoff ρmax = vth /ωc , given in Eq. (15) in57 .
vth = (kT /me )1/2 is the average thermal velocity. From the line center to ωc the collision operator
is essentially frequency independent, limiting the impact regime.

The collective properties of the electrons are usually assumed to occur through a time that
p
corresponds to the inverse of the electron plasma frequency, ω p = 4πNe e2 /me . However, if any
process reduces the characteristic duration of the perturbation, the correlation can be considered
lost. This is the case, for example, for high-n series lines of hydrogen, studied for plasma condi-
tions relevant to magnetic fusion and gas discharges experiments, where the line widths are larger
than the plasma frequency59 . It is also the case, for high density (Ne > 1018 cm−3 ) but relatively
low temperature (Te ≈ 1 eV ) plasmas. Due to the large number of electrons in the Debye sphere,
the correlation is lost when the electron configuration changes, i.e. when the electrons move. We
use the inverse time corresponding to a configuration change ωe = 2π/τe , where the characteristic
time of the interaction is τe = re /vth , with the average distance re = ( 34 πNe )−1/3 . Moreover,
as mentioned in the section III, it has been shown that the introduction of helical trajectories
reduces the characteristic duration of the perturbation to the order of the inverse of the Larmor
frequency. In this case, a cutoff at ωL = eB/me c should be used60 . Finally, for non-degenerated

13
FIG. 2. G(∆ω) calculated for the hydrogen Lyman−α line at Ne = 1017 cm−3 and kB T = 5 eV , for different
different B-field values that modify the cutoff frequency ωL : B = 100 tesla (short-dash), B = 500 tesla
(dot-dash) and B = 1 ktesla (double-dot-dash). For comparison, the G(∆ω) functions are shown for the
non-magnetized case (solid line).

systems, an additional cutoff at frequency ωαα 0 between the state α and α 0 has been retained too44 .

Hence, the cutoff frequency ωc , in Eq. (20) has been modified in PPPB to account for electron-
electron correlations and the helical trajectories. Here, ωc = max(ω p , ωe , ωL , ωαα 0 ).

The influence of the Larmor frequency variation on G(∆ω) is shown in Fig. 2, for conditions
relevant to white dwarfs atmosphere (Lyman-α line at Ne = 1017 cm−3 and kB T = 5 eV and
B = 100, 500 and103 tesla). The G(∆ω) calculated for B = 100 tesla is superposed to the non-
magnetized results as ωe is larger than ωL for those plasma conditions. For higher B-field values,
the cutoff is at the Larmor frequency. As the B-field value increases, the value of G(∆ω = 0)
decreases as well as the derivative of this function. The impact region is then extended. Such
results lead to a reduction of the Stark-Zeeman line width with increasing B40 .

The impact limit, G(∆ω = 0), is generally used in PPPB. It has been checked that for multi-

14
charged ions, this approximation only affect the wing of the lines for values ω of the order of or
larger than vth /rW (rW = h̄n2 /Zmvth being the Weisskopf radius).

V. SELECTED CALCULATIONS

In this section, we present different examples of applications in a broadband of plasma con-


ditions encountered in dense gas jet discharges and laser-produced plasma experiments in which
strong, controlled, static magnetic fields could be generated, as for example, by the mean of the
capacitor-coil target technique61–63 . In this technique, two parallel disks linked by a coil are irra-
diated by a high power nanosecond laser. Escaping hot electrons charge the target giving rise to a
strongh current passing through the coil. This strong current generates sub-ns duration B-fields of
strength up to one kilo-tesla.

A. Lyman−α lines of hydrogen-like Carbon in dense laser-produced plasmas submitted to


strong external magnetic fields: a need to account for intermediate B-fields.

We investigate the Stark-Zeeman effect on the C VI Lyman-α line at 367.55 eV in a dense


laser-produced plasma submitted to strong external magnetic fields experiment. For the selected
plasma conditions (Ne = 1019 − 1020 cm−3 and T = 100 eV ), the Stark broadening is less than
0.02 eV and the fine structure of the line, that corresponds to a splitting of 0.05 eV , can be seen.
Such lines are interesting because, under a magnetic field of the order of few hundred teslas, the
Zeeman splitting is sufficiently strong to prevail over the Stark broadening effect and, moreover,
none of the weak- or strong- field approximations are valid. The calculation of the atomic physics
within intermediate B-field approximation is required.

The figures 3, 4 and 5 show the Stark-Zeeman (SZ) broadened polarized profiles for three
different B-fields. They are calculated using atomic data generated within the weak-field approx-
imation (black line) or within the intermediate-field approximation (MASCB, red line). For the
sake of clarity, the π components are plotted with negative intensities. The pure Stark profiles
(dot black line) are also plotted to show the modification due to the Zeeman effect on those lines.
For B = 100 tesla, both approximations give quite the same Zeeman splitting and the profiles
are mostly identical. For B = 500 tesla, the weak field approximation starts to be critical and

15
differences appear on the SZ line shapes. For B = 103 tesla, the Zeeman splitting being of the
order of magnitude of the spin-orbit interaction, the weak-field approximation gives a drastically
different B-field signature on the spectrum compared to the intermediate approximation. Using
the latter, one can see a global blue-shift of the lines and different structures. In the limit of strong
fields, e.g B ∼ 104 tesla, it has been checked that the spectra calculated within the intermediate
B-field approximation tend to the ones calculated within the strong field approximation, in which
the spin-orbit interaction is neglected.

FIG. 3. Stark-Zeeman Lyman−α line profiles of CVI, using the weak-field approximation (dash) and the
intermediate field approximation (solid) for B = 100 tesla, Ne = 5 × 1019 cm−3 , T = 100 eV . Short-dash
line corresponds to the pure Stark profile.

16
FIG. 4. The same as FIG. 3 except for B = 500 tesla

FIG. 5. The same as FIG. 3 except for B = 103 tesla

17
B. Prospect on Lithium-like isoelectronic C IV, NV and O VI n = 4 − n = 5 lines submitted
to strong external magnetic fields

Measurements of Stark broadened profiles of the n = 4 to n = 5 transitions for the Lithium-


like isoelectronic sequences have been reported in64 . They were observed in a gas-liner pinch
discharge, where the plasma conditions, 1018 6 Ne (cm−3 ) 6 2.8 × 1018 and 8.6 6 kTe 6 17 eV ,
were independently diagnosed by Thomson scattering. The spectrometer resolution was suffi-
ciently high to well resolve the Stark broadened profiles. The width of the apparatus profile is 3%,
25% and 43% of the Stark broadened C IV, NV and O VI n = 4 − n = 5 transitions linewidths,
respectively. Not to mention that, the width of the corresponding Doppler profiles is below 0.1%
in all cases. The experimental set up described in64 provided a benchmark for models. We have
investigated the effects of strong B-fields on those lines.

Stark-Zeeman calculations are challenging here because the number of fine structure energy
levels, line transitions and Stark coupling between energy levels increase with the principal quan-
tum number. For the present calculations, levels belonging to n = 4, 5 and 6 have been considered.
154 fine structure energy levels associated with the upper (initial) levels and 32 fine structure
energy levels associated with the lower (final) ones are then accounted for and over 1,200 electric
dipole allowed transitions are taken into account in the calculations (including ∆n 6= 0).

The figure 6 shows the modifications of the calculated C IV n = 4 to n = 5 line shapes under
magnetic fields up to B = 500 tesla, for Ne = 2 × 1018 cm−3 and Te = Ti = 10 eV . In both fig-
ures, the pure Stark broadened line profile, i.e. B = 0 calculations, are plotted together with the
measured one to illustrate the very good agreement between PPP and the experimental data (taken
from64 ). The Zeeman patterns of the n = 4 to n = 5 line transitions show interesting features as
the line shapes corresponding to the σ components present two distinguishable peaks that split as
the B-field increases, whereas the Stark-Zeeman line shapes corresponding to the π components
do not really vary with B-field values.

Similar tendency is seen on the SZ line shapes of lithium-like Nitrogen and Oxygen. The figure
7 shows the Stark-Zeeman effect on the σ and π components of the C IV, NV and OVI n = 4 to
n = 5 transitions.

18
FIG. 6. Stark-Zeeman C IV n = 4 to n = 5 polarized line profiles for Ne = 2 × 1019 cm−3 , T = 10 eV , B = 0
(dot), B = 100 tesla (solid), B = 200 tesla (dash), B = 300 tesla (dot-dash) and B = 500 tesla (double-dot-
dash). a) σ components; b) π components. The experiments64 are plotted with solid-plus signs.

FIG. 7. Comparisons of the Stark-Zeeman broadened n = 4 to n = 5 line transitions for C IV (solid), N


V (dash), and O VI (dot-dash) at B = 500 T , Ne = 2 × 1018 cm−3 , T = 10 eV . a) σ components; b) π
components.

By recording experimentally the σ − and π−components simultaneously, it is possible to char-


acterize the polarization degree of the different Stark-Zeeman emission lines

Iπ (ω) − Iσ (ω)
P(ω) = (21)
Iπ (ω) + Iσ (ω)

The Fig. 8 shows the polarization degree calculated for the C IV n = 4 to n = 5 transitions for
different B-field values. As the π and σ components present very different line profiles the varia-
tion of the polarization degree goes up to 70%. This case is very favorable to infer the magnetic
fields because the Zeeman patterns are well observable among the Stark broadening. Neverthe-

19
FIG. 8. Polarization degree of the C IV n = 4 to n = 5 lines for Ne = 2×1018 cm−3 , T = 10 eV , B = 100 tesla
(solid), B = 200 tesla (dash), B = 300 tesla (dot-dash) and B = 500 tesla (double-dot-dash).

less, for cases where the Zeeman patterns tend to be masked by other broadening mechanisms,
few percents of polarization degree are still experimentally measurable65 .

The Fig. 9 shows the σ and π SZ line shapes of the O VI n = 4 to n = 5 transitions calcu-
lated for a magnetic field B = 100 tesla and convolved with a Lorentzian apparatus profile with
a FWHM of 0.18 nm that correspond to the detection system used in64 . Accounting for this ad-
ditional broadening, the σ components are no longer resolved and show a profile similar to the π
component one. A measure of B-field from the Zeeman patterns would be unreliable, whereas it
would be still feasible using the corresponding polarization degree that gives 5% at the center of
the line.
The comparison of the observed and calculated polarization degree can then be in principle
used as diagnostic tool to infer the magnetic field even if the Zeeman patterns are masked.

C. Ar K-shell emission in strongly magnetized plasmas

The design of a novel all-optical platform to magnetize laser-driven cylindrical implosions at


the OMEGA facility and their characterization through X-ray line emission has been recently
proposed66,67 . The experimental scheme combines the laser-driven MagLIF configuration for

20
FIG. 9. Stark-Zeeman line shapes, σ (solid) and π (dash) components and the corresponding polarization
degree of the O VI n = 4 to n = 5 lines for B = 100 tesla at Ne = 2 × 1018 cm−3 and T = 10 eV .

the implosion of low-density gas filled cylindrical targets68 with laser-driven seed B-fields63 .
A B-field exceeding 10 ktesla over the entire compressed core is predicted by the MHD code
GORGON69,70 . For the referred conditions, Ar K-shell spectra are expected to be observed,
thanks to the high quality spectroscopic data of Ar K-shell emission lines, with spectral resolution
E/∆E ∼ 1800, already obtained in Inertial Fusion Confinement experiments71,72 .

Three spectral properties of dopant atoms can be exploited to infer a unique measurement of
the core electron temperature and density as well as the local B-field: i) the Stark broadened line
shapes which depend strongly on the electron density, ii) the relative intensity distribution of K-
shell lines and associated satellites which are sensitive to the electron temperature and density72 ,
iii) the expected compressed B-field which is indeed strong enough to induce significant splitting,
broadening and polarization effects on the K-shell emission spectra17 .

We have performed investigation on the Stark-Zeeman broadened line shapes of Ar K-shell


X-ray transitions in Hydrogen- and Helium-like ions, namely Lyα (2p → 1s), Lyβ (3p → 1s),
Lyγ (4p → 1s) and Lyδ (5p → 1s) in H-like Ar and Heα (1s2p → 1s2 ), Heβ (1s3p → 1s2 ), Heγ
(1s4p → 1s2 ) in He-like Ar. Here a tracer amount of Ar in a deuterium plasma have been con-
sidered. A grid of plasma densities from Ne = 3 × 1022 cm−3 to Ne = 3 × 1024 cm−3 and a grid
of B-field values from 10 to 80 ktesla have been retained. Since the Stark-broadened line shapes
weakly depend on the electron temperature, only a representative value of 2 keV was chosen.
Such detailed SZ line shapes were used in the NLTE atomic kinetics code ABAKO73 to compute

21
synthetic X-ray emission spectra17,67 .

FIG. 10. Calculations of a) Ar He−α and b) He−β Stark-Zeeman spectral lines for Ne = 5 × 1023 cm−3 ,
T = 2 keV , B = 0 (solid), B = 20 ktesla (dash) and B = 40 ktesla (dot-dash). A convolution with an
instrumental resolution E/∆E = 1800 is performed. The observation is parallel to the magnetic field.

In the figure 10, the Stark-Zeeman lines shapes of the Ar He−α and He−β calculated for
two B-fields values (B = 20 ktesla and B = 40 ktesla) are compared to the corresponding Stark
broadened profiles (B = 0). The Doppler and instrumental resolution E/∆E = 1800 are accounted
for. As shown in the figure, the He−α line are more sensitive than He−β line in terms of Zeeman
splitting. The same tendency have been seen on the Ar H-like Lyman lines. As the principal
quantum number increases, the Stark broadening increases and masks the Zeeman patterns.

A possible way to measure the plasma parameters from the synthetic spectra would rely on the
following: by measuring simultaneously different emission lines one could characterize plasma
density and temperature from β −lines, through the Stark broadening. For the found plasma pa-
rameters, one could adjust the B-field value needed to reproduce the extra B-field-induced broad-
ening observed in experimental α−lines. Althought the He−α lines suffer from re-absorption for
such plasma conditions, the latter occurs in the center of the line. As the Zeeman effect splits the
lines, one can expect that the sigma components will be poorly re-absorbed, so that the B-field
diagnostic will be still feasible.

22
VI. CONCLUSION

Atomic structure can be used for magnetized plasma characterization, as hydrogen and multi-
ionized atom line emission broadens and gets polarized under strong B-fields. In this work, the
main feature of the Stark-Zeeman line shape code PPPB is presented through various applications
related to strongly magnetized plasmas encountered in astrophysics or in laboratory. PPPB allows
calculations for a wide range of plasma conditions and it is sufficiently fast to provide line shapes
to be used in radiation transport code. Zeeman effect in intermediate coupling is accounted for
by the mean of the atomic physics code MASCB that generates B-field dependent atomic physics
quantities. Investigation on hydrogen line series in highly magnetized astrophysical plasmas have
shown that the quadratic Zeeman terms give rise to additional structures and to a global shift that
increases with the principal quantum number. In high energy density plasmas, the measurement of
gigagauss (105 tesla) magnetic fields using Zeeman broadened lines from highly charged ions has
been proposed74 . Investigation of quadratic term effects on the Stark-Zeeman line shapes emitted
from highly charged ions may be of great importance as those lines are used as diagnostic tools.
Another interesting study is the effect of helical trajectories of the charged particles produced by
the presence of strong B-fields. The gyromotion of the ions and electrons may alter the dynamics
of the plasma particles and thus their interaction with the plasma emitters. The present version
of the code accounts for this effect by using a cutoff at the Larmor frequency in the electron
broadening operator. Investigation on anisotropy and screening effects on the electronic and ionic
microfield properties has to be done to improve the corresponding models in line shape codes.

ACKNOWLEDGMENTS

This work has been supported by the EUROfusion Enabling Research work programme 2017
(CfP-AWP17-IFE-CEA-02).

DATA AVAILABILITY STATEMENT

The data that support the findings of this study are available from the corresponding author
upon reasonable request.

23
REFERENCES

1 H. R. Griem, Principles of plasma spectroscopy (Principles of plasma spectroscopy. Cambridge:

Cambridge U Press, 1997).


2 H. Nguyen-Hoe, “Effet d’un champ magnetique uniforme sur les profils des raies de
l’hydrogene,” Journal of Quantitative Spectroscopy and Radiative Transfer 7, 429–474 (1967).
3 M. S. Murillo, M. E. Cox, and S. M. Carr, “Magnetized plasma microfield studies by molecular
dynamics simulation,” Journal of Quantitative Spectroscopy and Radiative Transfer 58, 811–820
(1997).
4 M. A. Gigosos and M. Á. González, “Comment on “a study of ion-dynamics and correlation
effects for spectral line broadening in plasma: K-shell lines”,” Journal of Quantitative Spec-
troscopy and Radiative Transfer 105, 533–535 (2007).
5 E. Stambulchik, K. Tsigutkin, and Y. Maron, “Spectroscopic method for measuring plasma
magnetic fields having arbitrary distributions of direction and amplitude,” Physical review letters
98, 225001 (2007).
6 J. Rosato, D. Reiter, V. Kotov, P. Börner, H. Capes, Y. Marandet, R. Stamm, S. Ferri, L. Godbert-

Mouret, M. Koubiti, et al., “Line shape modeling for radiation transport investigations in mag-
netic fusion plasmas,” High energy density physics 5, 93–96 (2009).
7 G. Mathys, “The transfer of polarized light in stark broadened hydrogen lines in the presence
of a magnetic field,” Journal of Quantitative Spectroscopy and Radiative Transfer 44, 143–151
(1990).
8 A. Derevianko and E. Oks, “Generalized theory of ion impact broadening in magnetized plasmas

and its applications for tokamaks,” Physical review letters 73, 2059 (1994).
9 S. Brillant, G. Mathys, and C. Stehle, “Hydrogen line formation in dense magnetized plasmas,”
Astronomy and Astrophysics 339, 286–297 (1998).
10 S. Günter and A. Könies, “Diagnostics of dense plasmas from the profile of hydrogen spectral
lines in the presence of a magnetic field,” Journal of Quantitative Spectroscopy and Radiative
Transfer 62, 425–431 (1999).
11 L. Godbert-Mouret, M. Koubiti, R. Stamm, K. Touati, B. Felts, H. Capes, Y. Corre, R. Guir-
let, and C. De Michelis, “Spectroscopy of magnetized plasmas,” Journal of Quantitative Spec-
troscopy and Radiative Transfer 71, 365–372 (2001).

24
12 M. Adams, R. Lee, H. Scott, H. Chung, and L. Klein, “Complex atomic spectral line shapes in
the presence of an external magnetic field,” Physical Review E 66, 066413 (2002).
13 X.-d. Li, S.-s. Han, C. Wang, and Z.-z. Xu, “Ultrahigh magnetic field diagnostic with spec-
tral profile calculation,” Journal of Quantitative Spectroscopy and Radiative Transfer 76, 31–43
(2003).
14 S. Ferri, A. Calisti, C. Mossé, L. Mouret, B. Talin, M. A. Gigosos, M. A. González, and
V. Lisitsa, “Frequency-fluctuation model applied to stark-zeeman spectral line shapes in plas-
mas,” Physical Review E 84, 026407 (2011).
15 C. A. Iglesias, “Efficient algorithms for stark–zeeman spectral line shape calculations,” High
Energy Density Physics 9, 737–744 (2013).
16 F. Gilleron and J.-C. Pain, “Zest: A fast code for simulating zeeman-stark line-shape functions,”
Atoms 6, 11 (2018).
17 J. J. Santos, M. Bailly-Grandvaux, M. Ehret, A. Arefiev, D. Batani, F. Beg, A. Calisti, S. Ferri,
R. Florido, P. Forestier-Colleoni, et al., “Laser-driven strong magnetostatic fields with applica-
tions to charged beam transport and magnetized high energy-density physics,” Physics of Plas-
mas 25, 056705 (2018).
18 C. Liu, K. Matsuo, S. Ferri, H.-K. Chung, S. Lee, S. Sakata, K. F. F. Law, H. Morita, B. Pollock,
J. Moody, et al., “Design of zeeman spectroscopy experiment with magnetized silicon plasma
generated in the laboratory,” High Energy Density Physics 33, 100710 (2019).
19 R. D. Cowan, The theory of atomic structure and spectra, 3 (Univ of California Press, 1981).
20 I. Grant, B. McKenzie, P. Norrington, D. Mayers, and N. Pyper, “An atomic multiconfigura-
tional dirac-fock package,” Computer Physics Communications 21, 207–231 (1980).
21 M. F. Gu, “The flexible atomic code,” Canadian Journal of Physics 86, 675–689 (2008).
22 E. Stambulchik, “cFAC code in GitHub kernel description,”.
23 B. Cagnac and J. C. P. Peyroula, Physique atomique, tome 2: applications de la mecanique
quantique (Bordas, 1982).
24 J. C. Slater, “Quantum theory of atomic structure,” Tech. Rep. (1960).
25 M. Litsarev and O. Ivanov, “Multiconfiguration hartree-fock method: Direct diagonalization for
the construction of a multielectron basis,” Journal of Experimental and Theoretical Physics 111,
22–26 (2010).
26 E. Hill, “Calculation of unit tensor operators using a restricted set of slater determinants,” Journal

of Quantitative Spectroscopy and Radiative Transfer 140, 1–6 (2014).

25
27 E. U. Condon, E. Condon, and G. Shortley, The theory of atomic spectra (Cambridge University
Press, 1935).
28 J. D. Talman and W. F. Shadwick, “Optimized effective atomic central potential,” Physical Re-
view A 14, 36 (1976).
29 J. Krieger, Y. Li, and G. Iafrate, “Systematic approximations to the optimized effective potential:
Application to orbital-density-functional theory,” Physical Review A 46, 5453 (1992).
30 R. Garstang and S. Kemic, “Hydrogen and helium spectra in large magnetic fields,” Astrophysics

and Space Science 31, 103–115 (1974).


31 S. O. Kepler, I. Pelisoli, S. Jordan, S. J. Kleinman, D. Koester, B. Külebi, V. Peçanha, B. G.
Castanheira, A. Nitta, J. E. d. S. Costa, et al., “Magnetic white dwarf stars in the sloan digital
sky survey,” Monthly Notices of the Royal Astronomical Society 429, 2934–2944 (2013).
32 J. Rosato, “Hydrogen line shapes in plasmas with large magnetic fields,” Atoms 8, 74 (2020).
33 A. Raji, J. Rosato, R. Stamm, and Y. Marandet, “New analysis of balmer line shapes in magnetic
white dwarf atmospheres,” The European Physical Journal D 75, 1–4 (2021).
34 D. R. Inglis and E. Teller, “Ionic depression of series limits in one-electron spectra.” The Astro-
physical Journal 90, 439 (1939).
35 T. Ott and M. Bonitz, “Diffusion in a strongly coupled magnetized plasma,” Physical review
letters 107, 135003 (2011).
36 S. D. Baalrud and J. Daligault, “Transport regimes spanning magnetization-coupling phase
space,” Physical Review E 96, 043202 (2017).
37 D. J. Bernstein, T. Lafleur, J. Daligault, and S. D. Baalrud, “Friction force in strongly magne-
tized plasmas,” Physical Review E 102, 041201 (2020).
38 E. Oks, “Influence of magnetic-field-caused modifications of trajectories of plasma electrons on
spectral line shapes: Applications to magnetic fusion and white dwarfs,” Journal of Quantitative
Spectroscopy and Radiative Transfer 171, 15–27 (2016).
39 J. Rosato, S. Ferri, and R. Stamm, “Influence of helical trajectories of perturbers on stark line
shapes in magnetized plasmas,” Atoms 6, 12 (2018).
40 S. Alexiou, “Line shapes in a magnetic field: Trajectory modifications i: Electrons,” Atoms 7,
52 (2019).
41 S. Alexiou, “Line shapes in a magnetic field: Trajectory modifictions ii: Full collision-time
statistics,” Atoms 7, 94 (2019).

26
42 C. Deutsch, “Electric microfield distributions in plasmas in presence of a magnetic field,”
Physics Letters A 30, 381–382 (1969).
43 S. Ferri, “Study of plasma microfield properties in highly magnetized plasmas,” (2018) unpub-
lished results.
44 A. Calisti, F. Khelfaoui, R. Stamm, B. Talin, and R. Lee, “Model for the line shapes of complex
ions in hot and dense plasmas,” Physical Review A 42, 5433 (1990).
45 A. Calisti, L. Godbert, R. Stamm, and B. Talin, “Fast numerical methods for line shape studies in
hot and dense plasmas,” Journal of Quantitative Spectroscopy and Radiative Transfer 51, 59–64
(1994).
46 N. Woolsey, A. Asfaw, B. Hammel, C. Keane, C. Back, A. Calisti, C. Mosse, R. Stamm, B. Talin,

J. Wark, et al., “Spectroscopy of compressed high energy density matter,” Physical Review E 53,
6396 (1996).
47 N. Woolsey, B. Hammel, C. Keane, A. Asfaw, C. Back, J. Moreno, J. Nash, A. Calisti, C. Mosse,

R. Stamm, et al., “Evolution of electron temperature and electron density in indirectly driven
spherical implosions,” Physical Review E 56, 2314 (1997).
48 B. Talin, A. Calisti, L. Godbert, R. Stamm, R. Lee, and L. Klein, “Frequency-fluctuation model
for line-shape calculations in plasma spectroscopy,” Physical Review A 51, 1918 (1995).
49 M. Abramowitz and I. Stegun, “Handbook of mathematical functions. original work published
1964,” (1972).
50 B. Talin, E. Dufour, A. Calisti, M. A. Gigosos, M. A. Gonzalez, T. del Rio Gaztelurrutia, and
J. W. Dufty, “Molecular dynamics simulation for modelling plasma spectroscopy,” Journal of
Physics A: Mathematical and General 36, 6049 (2003).
51 C. A. Iglesias, H. E. DeWitt, J. L. Lebowitz, D. MacGowan, and W. B. Hubbard, “Low-
frequency electric microfield distributions in plasmas,” Physical Review A 31, 1698 (1985).
52 C. Iglesias, F. Rogers, R. Shepherd, A. Bar-Shalom, M. Murillo, D. Kilcrease, A. Calisti, and
R. Lee, “Fast electric microfield distribution calculations in extreme matter conditions,” Journal
of Quantitative Spectroscopy and Radiative Transfer 65, 303–315 (2000).
53 C. Hooper Jr, “Low-frequency component electric microfield distributions in plasmas,” Physical
Review 165, 215 (1968).
54 N. Woolsey, B. Hammel, C. Keane, C. Back, J. Moreno, J. Nash, A. Calisti, C. Mosse, L. God-
bert, R. Stamm, et al., “Spectroscopic line shape measurements at high densities,” Journal of
Quantitative Spectroscopy and Radiative Transfer 58, 975–989 (1997).

27
55 S. Ferri, A. Calisti, C. Mossé, J. Rosato, B. Talin, S. Alexiou, M. A. Gigosos, M. A. González,
D. González-Herrero, N. Lara, et al., “Ion dynamics effect on stark-broadened line shapes: A
cross-comparison of various models,” Atoms 2, 299–318 (2014).
56 A. Calisti, C. Mossé, S. Ferri, B. Talin, F. Rosmej, L. Bureyeva, and V. Lisitsa, “Dynamic stark
broadening as the dicke narrowing effect,” Physical Review E 81, 016406 (2010).
57 H. R. Griem, M. Blaha, and P. C. Kepple, “Stark-profile calculations for lyman-series lines of
one-electron ions in dense plasmas,” Physical Review A 19, 2421 (1979).
58 E. Galtier, F. Rosmej, A. Calisti, B. Talin, C. Mossé, S. Ferri, and V. Lisitsa, “Interference effects
and stark broadening in xuv intrashell transitions in aluminum under conditions of intense xuv
free-electron-laser irradiation,” Physical Review A 87, 033424 (2013).
59 S. Ferri, A. Calisti, R. Stamm, B. Talin, R. Lee, and L. Klein, “Electronic broadening model for
high-n balmer line profiles,” Physical Review E 58, R6943 (1998).
60 E. Maschke and D. Voslamber, “Stark broadening of hydrogen lines in strong magnetic fields,”
in Phenomena in Ionized Gases, Volume II, VII International Conference (1966) p. 568.
61 H. Daido, F. Miki, K. Mima, M. Fujita, K. Sawai, H. Fujita, Y. Kitagawa, S. Nakai, and C. Ya-
manaka, “Generation of a strong magnetic field by an intense co 2 laser pulse,” Physical review
letters 56, 846 (1986).
62 S. Fujioka, Z. Zhang, K. Ishihara, K. Shigemori, Y. Hironaka, T. Johzaki, A. Sunahara, N. Ya-
mamoto, H. Nakashima, T. Watanabe, et al., “Kilotesla magnetic field due to a capacitor-coil
target driven by high power laser,” Scientific reports 3, 1–7 (2013).
63 J. Santos, M. Bailly-Grandvaux, L. Giuffrida, P. Forestier-Colleoni, S. Fujioka, Z. Zhang, P. Ko-
rneev, R. Bouillaud, S. Dorard, D. Batani, et al., “Laser-driven platform for generation and char-
acterization of strong quasi-static magnetic fields,” New Journal of Physics 17, 083051 (2015).
64 S. Glenzer, T. Wrubel, S. Buscher, H.-J. Kunze, L. Godbert, A. Calisti, R. Stamm, B. Talin,
J. Nash, R. Lee, et al., “Spectral line profiles of n= 4 to n= 5 transitions in c iv, nv and o vi,”
Journal of Physics B: Atomic, Molecular and Optical Physics 27, 5507 (1994).
65 T. Fujimoto and A. Iwamae, Plasma polarization spectroscopy, Vol. 44 (Springer, 2008).
66 M. Bailly-Grandvaux, S. McGuffey, F. Beg, S. Ferri, A. Calisti, J. Davies, R. Florido,
M. Gigosos, J. Honrubia, R. Mancini, et al., “An all-optical platform to characterize strongly
magnetized hot dense plasmas at> 10 kt,” in APS Division of Plasma Physics Meeting Abstracts,
Vol. 2020 (2020) pp. BO07–005.

28
67 R. Florido, C. Walsh, M. Bailly-Grandvaux, F. Beg, C. McGuffey, A. Calisti, S. Ferri,
M. Gigosos, R. Mancini, T. Nagayama, et al., “Spectroscopic and mhd modeling of magne-
tized cylindrical implosions using a laser-produced seed b-field,” in APS Division of Plasma
Physics Meeting Abstracts, Vol. 2020 (2020) p. GP17.00009.
68 J. Davies, D. Barnak, R. Betti, E. Campbell, V. Y. Glebov, E. Hansen, J. Knauer, J. Peebles, and
A. Sefkow, “Inferring fuel areal density from secondary neutron yields in laser-driven magne-
tized liner inertial fusion,” Physics of Plasmas 26, 022706 (2019).
69 C. Walsh, K. McGlinchey, J. Tong, B. Appelbe, A. Crilly, M. Zhang, and J. Chittenden, “Pertur-
bation modifications by pre-magnetisation of inertial confinement fusion implosions,” Physics
of Plasmas 26, 022701 (2019).
70 C. Walsh, J. Chittenden, D. Hill, and C. Ridgers, “Extended-magnetohydrodynamics in under-
dense plasmas,” Physics of Plasmas 27, 022103 (2020).
71 N. Woolsey, B. Hammel, C. Keane, C. Back, J. Moreno, J. Nash, A. Calisti, C. Mosse, R. Stamm,

B. Talin, et al., “Competing effects of collisional ionization and radiative cooling in inertially
confined plasmas,” Physical Review E 57, 4650 (1998).
72 H.-K. Chung and R. Lee, “Applications of nlte population kinetics,” High Energy Density
Physics 5, 1–14 (2009).
73 R. Florido, R. Rodríguez, J. Gil, J. Rubiano, P. Martel, E. Mínguez, and R. Mancini, “Model-
ing of population kinetics of plasmas that are not in local thermodynamic equilibrium, using a
versatile collisional-radiative model based on analytical rates,” Physical Review E 80, 056402
(2009).
74 J. F. Seely, “Gigagauss magnetic field measurements using zeeman broadening of ne-like transi-
tions in highly charged ions,” Review of Scientific Instruments 92, 053535 (2021).

29

You might also like