Multiply-excited states and their contribution to opacity in CO 2 laser-driven tin-plasma conditions - IOPscience

 
CONTINUE READING
Multiply-excited states and their contribution to opacity in CO 2 laser-driven tin-plasma conditions - IOPscience
Journal of Physics B: Atomic, Molecular and Optical Physics

PAPER • OPEN ACCESS

Multiply-excited states and their contribution to opacity in CO2 laser-
driven tin-plasma conditions
To cite this article: J Sheil et al 2021 J. Phys. B: At. Mol. Opt. Phys. 54 035002

View the article online for updates and enhancements.

                                 This content was downloaded from IP address 46.4.80.155 on 13/07/2021 at 18:37
Multiply-excited states and their contribution to opacity in CO 2 laser-driven tin-plasma conditions - IOPscience
Journal of Physics B: Atomic, Molecular and Optical Physics

J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002 (11pp)                                                                    https://doi.org/10.1088/1361-6455/abcedf

Multiply-excited states and their
contribution to opacity in CO2 laser-driven
tin-plasma conditions
J Sheil1,∗ , O O Versolato1,2 , A J Neukirch3 and J Colgan3
1
  Advanced Research Center for Nanolithography, Science Park 106, 1098 XG Amsterdam, The
Netherlands
2
  Department of Physics and Astronomy, and LaserLaB, Vrije Universiteit, De Boelelaan 1081, 1081 HV
Amsterdam, The Netherlands
3
  Los Alamos National Laboratory, Los Alamos, NM 87545, United States of America

E-mail: j.sheil@arcnl.nl and jcolgan@lanl.gov

Received 22 September 2020, revised 17 November 2020
Accepted for publication 30 November 2020
Published 22 January 2021

Abstract
A recent study (2020 Nat. Commun. 11 2334) has found that transitions between
multiply-excited configurations in open 4d-subshell tin ions are the dominant contributors to
intense EUV emission from dense, Nd:YAG-driven (laser wavelength λ = 1.064 μm) tin
plasmas. In the present study, we employ the Los Alamos Atomic code to investigate the
spectral contribution from these transitions under industrially-relevant, CO2 laser-driven
(λ = 10.6 μm) tin plasma conditions. First, we employ Busquet’s ionisation temperature
method to match the average charge state Z of a non-local-thermodynamic equilibrium
(non-LTE) plasma with an LTE one. This is done by varying the temperature of the LTE
calculations until a so-called ionisation temperature T Z is established. Importantly, this
approach generates LTE-computed configuration populations in excellent agreement with the
non-LTE populations. A corollary of this observation is that the non-LTE populations are
well-described by Boltzmann-type exponential distributions having effective temperatures
T eff ≈ T Z . In the second part of this work, we perform extensive level-resolved LTE opacity
calculations at T Z . It is found that 66% of the opacity in the industrially-relevant 2%
bandwidth centred at 13.5 nm arises from transitions between multiply-excited states. These
results reinforce the need for the consideration of complex, multiply-excited states in
modelling the radiative properties of laser-driven plasma sources of EUV light.
Keywords: opacity, non-LTE, effective temperature, multiply-excited states
(Some figures may appear in colour only in the online journal)

1. Introduction                                                                          [2–4], materials science [5, 6] and the manufacturing of inte-
                                                                                         grated circuits (ICs) in the semiconductor industry [7, 8]. The
Powerful sources of short-wavelength radiation play a cru-                               use of short-wavelength radiation in this latter application is
cial role in many high-profile scientific and industrial endeav-                         expected to enable the printing of sub-10 nm feature sizes on
ours [1]. Examples include high-resolution biological imaging                            ICs [9], thus sustaining the production of cheaper and more
                                                                                         powerful computer chips into the future.
∗
 Author to whom any correspondence should be addressed.                                     The key technology driving feature-size miniaturisation in
                  Original content from this work may be used under the terms
                  of the Creative Commons Attribution 4.0 licence. Any further
                                                                                         the semiconductor industry today is extreme ultraviolet lithog-
distribution of this work must maintain attribution to the author(s) and the title       raphy (EUVL) [10–13]. As the name suggests, this technology
of the work, journal citation and DOI.                                                   uses EUV light to print features on ICs. In an industrial setting,

0953-4075/21/035002+11$33.00                                                         1                                © 2021 IOP Publishing Ltd    Printed in the UK
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                 J Sheil et al

this EUV light is generated in a laser-produced plasma (LPP)            are therefore significantly more extensive than those consid-
formed by irradiating tin microdroplets with CO2 laser light            ered previously. Importantly, we find that transitions between
(λ = 10.6 μm) [14, 15]. The ensuing laser–droplet interaction           multiply-excited states make a substantial contribution to the
generates a plasma having moderate electron temperatures and            opacity in the 5–20 nm region in plasma conditions pertinent
densities (T e ≈ 20–40 eV, ne ≈ 1018 –1019 cm−3 ) compris-              to those generated in CO2 laser-driven tin plasmas.
ing large populations of Sn8+ –Sn14+ ions. Spectra recorded                 The current paper is organised as follows: in section 2,
from such plasmas are known to exhibit an intense, narrow-              we describe the atomic structure and opacity calculations per-
band feature peaked near 13.5 nm [16–20]. Importantly, this             formed with the Los Alamos suite of atomic physics and
emission feature provides optimum wavelength overlap with               population kinetics codes. An examination of non-LTE- and
the 2% reflective bandwidth (bw) centred at 13.5 nm defined             LTE-computed tin ionisation distributions and configuration
by molybdenum/silicon multilayer mirrors (MLMs) employed                populations in CO2 laser-driven plasma conditions follows
in lithography tools [12]. This combination of (i) optimum              in section 3. In section 4 we present level-resolved LTE-
emission characteristics and (ii) a debris-limited droplet-target       computed opacities of Sn11+ –Sn14+ . Finally, a summary of
(key for minimising MLM contamination from plasma expan-                this work is given in section 5.
sion [21]) established tin LPPs as the light source of choice for
EUVL. In fact, after more than 20 years of research and devel-
opment, EUVL has now entered the realm of high-volume                   2. Atomic structure and opacity calculations
manufacturing of new-generation ICs [22].
   The aforementioned intense, narrowband emission feature              In the present work, we have employed the Los Alamos
observed in laser-produced tin plasmas has, to date, been               national laboratory (LANL) suite of atomic structure and colli-
                                                                        sion codes [42] as well as the population kinetics code Atomic
predominantly attributed to overlapping Δn = 0 transition
                                                                        [43, 44] to calculate the opacities of a tin plasma at densities
arrays of the form 4p6 4dm − (4p5 4dm+1 + 4p6 4dm−1 4 f) in
                                                                        and temperatures relevant to those attained in a CO2 laser-
charge states Sn8+ –Sn14+ (m = 6–0) [23–25]. Such claims
                                                                        driven plasma. Given the central role that these codes play
have been supported by charge-state-resolved spectro-
                                                                        in this study, we devote the following subsections to their
scopic experiments [26–29] as well as numerous theoretical
                                                                        description.
efforts employing advanced atomic structure codes [30–39].
Although serving as the general consensus for many years,
a recent study of Torretti et al has challenged this view-              2.1. Atomic structure calculations
point [13]. In the most detailed tin-ion opacity calculations
                                                                        The first step in the present opacity calculations is the gen-
performed to-date, it was found that transitions between
                                                                        eration of all relevant atomic data (energy levels, oscillator
multiply-excited configurations in Sn11+ –Sn14+ are in
                                                                        strengths, ionisation potentials, etc). In Torretti et al [13], con-
fact the dominant contributors to EUV emission in Nd:YAG-
                                                                        siderable effort was devoted to benchmarking the calculated
driven (λ = 1.064 μm) tin LPPs. Surprisingly, the well-known
                                                                        atomic structures with available experimental data and there-
4p6 4dm − (4p5 4dm+1 + 4p6 4dm−1 4 f) transitions play only a
                                                                        fore we have utilised these atomic structures in the present
minor role; they comprise 11% of the opacity in the 5–20 nm
                                                                        study. In brief, the Cats code [45] was employed in the cal-
region and 19% of the opacity in the 2% bw. Remarkably, it is
                                                                        culation of level energies and absorption oscillator strengths.
transitions between complex, multiply-excited configurations            This code is based on Cowan’s original atomic structure
(doubly-, triply- and quadruply-excited configurations) which           code [46, 47] and exhibits a favourable feature where one
dominate the opacity spectrum. Strong support for their spec-           can scale the magnitude of the radial integrals in order to
troscopic dominance was evidenced through comparisons of                bring theoretical level structures into excellent agreement with
experimental tin-droplet EUV spectra with one-dimensional               experimental observations. This semi-empirical approach has
radiation transport simulations. The incorporation of multiply-         proven to be an indispensable tool in atomic spectroscopy,
excited transitions in these simulations was deemed crucial             especially for systems exhibiting complicated level structures
for the reproduction of the experimental spectra.                       [29, 48–51]. In the work of Torretti et al [13], optimum
   The purpose of the present paper is to elucidate the                 agreement between atomic structure calculations and previous
role of multiply-excited states in CO2 -driven tin plasmas              experimental observations [17, 26, 27], was found by applying
which are currently used in the industry to provide light for           scaling factors of 0.87/0.75 to the radial integrals associated
EUVL. The main difference between the present work and                  with singly-/multiply-excited energy levels. These two scal-
previous atomic modelling work of Sasaki and co-authors                 ing factors were adopted for all charge states considered in
[37, 40, 41] is the extensivity of the atomic structures                this work. Extensive configuration sets were employed in the
employed. For example, in reference [37, 40, 41], a less exten-         calculations through use of the so-called ‘2-mode’ approach.
sive number of doubly-excited states were included in their             In these calculations, configuration interaction (CI) was
atomic models. Their inclusion was mainly for the purpose               retained for a large number of configurations associated with
of understanding satellite transition array emission of the             Δn = 0 and Δn = 1 single, double and triple excitations
form 4p6 4dm−1 nl − (4p5 4dm nl + 4p6 4dm−2 4 fnl). The config-         from the n = 4 manifold of the ground-state configuration.
uration sets adopted in the present work include large numbers          Additional configurations were included using intermediate
of doubly-, triply- and quadruply-excited configurations and            coupling (only interactions between levels of the same con-

                                                                    2
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                          J Sheil et al

figuration are considered) [39]. This approach retains a signif-            approach and the occupation probability formalism (this lat-
icant level of detail provided by full CI calculations however              ter procedure allows plasma density effects to be included in
at a much reduced computational cost.                                       a consistent manner). Importantly, non-ideal plasma effects,
                                                                            such as plasma microfields and strong coupling are included
2.2. Opacity calculations                                                   in the free energy term. This leads to modified Boltzmann
                                                                            and Saha relations from which level populations and CSDs are
Knowledge of opacities is crucial for understanding radi-
                                                                            derived.
ation transport in a host of plasma environments, such as
                                                                                Atomic also has the capability of modelling plasmas under
those encountered in stellar interiors, fusion devices and short-
                                                                            non-LTE conditions [58, 59]. Unlike LTE plasmas, the cal-
wavelength plasma light sources [30, 52–55]. In this context,
                                                                            culation of level populations in non-LTE conditions requires
an opacity quantifies how photons, travelling in a plasma, are
                                                                            a microscopic viewpoint in which the rates of all relevant
absorbed and scattered through interactions with the plasma
                                                                            atomic populating and depopulating processes are calculated.
constituents. For a given photon frequency ν, the opacity κTν is
                                                                            The most general definition of a non-LTE plasma is one which
defined as the contribution of four mechanisms:
                                                                            exhibits any characteristic departure from LTE conditions,
                  κTν = κBB   BF   FF   SCAT                                such as the existence of a non-Maxwellian electron energy dis-
                         ν + κν + κν + κν    .                   (1)
                                                                            tribution function (EEDF) or a non-Planckian radiation field.
   The first three terms represent bound–bound (photoab-                    As such, there are a variety of ways that a plasma can be
sorption), bound–free (photoionisation) and free–free (inverse              defined as under non-LTE conditions. Consider an atomic state
bremsstrahlung) contributions to the opacity and each contain               (energy level, configuration or superconfiguration) i associ-
a factor correcting for stimulated emission. The final term rep-            ated with some charge state ζ in the plasma. In non-LTE
resents photon scattering. In the present work, we are mainly               plasmas, the population of this state evolves according to a
concerned with the bound–bound contribution to the opac-                    collisional-radiative (CR) equation of the form [60, 61]
ity (since all other contributions are minor in the wavelength                                                           
                                                                                         dniζ
range of interest for these plasma conditions) and therefore                                  =    n jζ  R jζ →iζ − niζ    Riζ→jζ  .           (4)
we drop the superscript ‘BB’ notation in the following. The                               dt                              
                                                                                                   jζ                      jζ
bound–bound opacity κν is written
                                                                              Here, R jζ →iζ = RCjζ →iζ + RRjζ →iζ denotes the rate of colli-
                 κν = ρ−1 σνi j (ni − gi g−1
                                          j n j),            (2)            sional (RCjζ →iζ ) and radiative (RRjζ →iζ ) processes connecting j
                                 i< j
                                                                            (associated with some charge state ζ  ) to i. In a similar fash-
where ρ is the mass density and the summation term is known                 ion, rates describing i → j are denoted Riζ→jζ  . The first term in
as the absorptivity. Here, σνi j is the cross section for photoab-          equation (4) represents population influx into i and the second
sorption from level i to level j and ni( j) are the population den-         term describes population outflux from i. A CR equation of the
sities (simply called ‘populations’ in the following) of energy             form equation (4) exists for all atomic states i in each charge
levels i( j) having statistical weights gi( j) . The sum runs over          state ζ present in the plasma. Coupling these states yields the
all transitions i→j associated with photon absorption at fre-               so-called rate equation
quency ν. The second term in the summation is the aforemen-
tioned correction for stimulated emission. The cross section                                             dN
                                                                                                            = AN.                                 (5)
for photoabsorption σνi j is given by                                                                    dt

                                      πe2                                       Here, N is a vector of atomic state populations (niζ ∈ N)
                            σνi j =        f i jφ(ν),            (3)        and A is the so-called rate matrix containing all relevant rates
                                      me c
                                                                            RC,R
                                                                              jζ →iζ,iζ→jζ  .
where e, me and c are the electron charge, electron mass and                    The choice of atomic processes to be included in the rate
speed of light, respectively. The term f i j is the absorption oscil-       matrix A is highly case-dependent; in coronal plasmas, for
lator strength for a transition
                      ∞         i → j and φ(ν) is the line profile         example, one can neglect the role of three-body recombina-
function satisfying 0 φ(ν)dν = 1.                                           tion. Such simplifications are not possible in the current case
     As one can see from equations (2) and (3), evaluation of               and therefore we must consider a wide range of atomic pro-
the bound–bound opacity requires knowledge of both atomic                   cesses (see section 3 for more details). Fortunately, Atomic
structure (ν and f i j) and the level populations ni( j) . Evaluation       has the capability of modelling a variety of (de-)populating
of this latter quantity requires knowledge of the plasma envi-              processes such as electron-impact excitation/ionisation, pho-
ronment in which the atomic species is embedded. Atomic,                    toexcitation, photoionisation, autoionisation (AI), etc. Cross
using as input the mass density ρ and electron temperature                  sections for electron-impact excitation were calculated in the
T e , can calculate these level populations in LTE or non-                  plane-wave Born (PWB) approximation using the CATS code.
LTE plasma conditions (the radiation temperature T r can                    Electron-impact ionisation cross-sections were obtained from
also be specified in non-LTE calculations). The LTE-mode of                 the GIPPER code (photoionisation cross sections can also be
Atomic uses the ChemEOS equation-of-state (EOS) model to                    obtained if an external radiation field is included in the non-
derive level populations and charge state distributions (CSDs)              LTE model) [62]. Atomic uses these cross sections to calculate
[56, 57]. This model is based on the free-energy-minimisation               the rates which form the entries of A. Invoking a conservation

                                                                        3
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                                   J Sheil et al

                                                         12+         14+
          Table 1. Configuration sets adopted for Sn    and Sn     in the non-LTE/LTE CA calculations (see main text).
          Configurations are grouped according to excitation degree. Here, angular momenta l = 0 – 4(s − g), l = 1 – 4(p − g)
          and l̃ = 2 – 4 (d − g), respectively.
          Sn12+                                                      Sn14+
          Ground configuration                                       Ground configuration
          4s2 4p6 4d2                                                4s2 4p6
          Singly-excited configurations                              Singly-excited configurations
          4s2 4p6 4d + {4 f, 5l}                                     4s2 4p5 + {4d, 4 f, 5l}
          4s2 4p5 + {4d 3 , 4d 2 4 f, 4d 2 5l}                       4s1 4p6 + {4d, 4 f, 5l}
          4s4p6 + {4d 3 , 4d 2 4 f, 4d2 5l}
          Doubly-excited configurations                              Doubly-excited configurations
          4s2 4p6 + {4 f 2 , 4 f 5l, 5s5l, 5p5l , 5d5l̃ , 5 f 5g}   4s2 4p4 + {4d 2 , 4d4 f, 4 f 2 , 4d5l, 4 f5l, 5s5l, 5p5l , 5d5 f, 5d5g, 5 f5g}
          4s2 4p5 + {4d4 f 2 , 4d5s2 , 4d4 f5l, 4d5s5p, 4d5s5d}      4s1 4p5 + {4d 2 , 4d4 f, 4 f 2 , 4d5l, 4 f5l, 5s5l, 5p5l , 5d5 f, 5d5g, 5 f5g}
          4s2 4p4 + {4d 4 , 4d 3 4 f, 4d 3 5l, 4d 2 4 f 2 }          4p6 + {4d 2 , 4d4 f, 4d5l, 4 f5l}
          4s4p5 4d 4
          Triply-excited configuration                               Triply-excited configurations
          4s2 4p3 4d5                                                4s2 4p3 4d 3 , 4s2 4p3 4d 2 4 f, 4s4p4 4d3 , 4s4p4 4d2 4 f

equation of the form                                                             this choice of extensivity over simplicity makes the solution of
                                                                                 equation (5) for these level structures intractable; there are sim-
                            
                           Z Ni −1
                                                                                 ply too many atomic levels to consider. As discussed in refer-
                                        niζ = nion ,                   (6)       ence [13], the calculation of detailed, level-resolved opacities
                            ζ =0 i =0
                                                                                 for such complex structures is currently only possible using
where N i is the number of atomic states and nion is the ion                     the LTE approach in Atomic.
density allows one to solve equation (5) for the atomic state
populations niζ . These populations can then used to produce                     3.1. Busquet’s ionisation temperature method
opacities [equation (2)] in a similar fashion as in the LTE mode
of Atomic.                                                                       In order to circumvent computational intractabilities associ-
    Finally, it should be emphasized that a consistent atomic                    ated with level-resolved, non-LTE opacity calculations involv-
structure has been employed in all stages of the calculation,                    ing multiply-excited configurations, we have employed the
ranging from the evaluation of collisional cross-sections to the                 ionisation temperature method of Busquet [64, 65] to effec-
derivation of atomic state populations and level-resolved opac-                  tively ‘mimick’ non-LTE opacities using LTE computations.
ities. This procedure ensures self-consistency in all aspects of                 The key observation underlying this method is that non-LTE
the calculation.                                                                 CSDs (performed at some T e ) often resemble LTE CSDs
                                                                                 undertaken at a temperature T Z different to T e [66]. The first
                                                                                 step in Busquet’s approach is to identify the LTE temperature
3. Tin ionisation distributions and configuration                                T Z which yields optimum agreement in average charge state
populations in CO2 laser-driven plasma conditions                                Z between the non-LTE and LTE calculations, i.e.
The study of CO2 laser irradiation of tin targets comprises a
large body of work (see [20] and references therein). A very                                         Z(Te )non−LTE = Z(TZ )LTE .                       (7)
important aspect of this work is the study of tin-ion opaci-
ties in the relevant plasma conditions [32, 37–39, 63]. Plasmas                     In the current context, T Z is known as the ‘ionisation
generated by CO2 laser irradiation lie in a rather unique posi-                  temperature’. Once T Z is determined, one then performs
tion in (ne , T e ) space; electron densities are not high enough                detailed, level-resolved LTE opacity calculations at T Z .
to allow for LTE conditions to prevail, however not so low to                    Higher-order moments of the CSD, such as the width, are
justify the use of coronal equilibrium relations. In this respect,               not considered in this method. Although this approach pro-
the atomic modelling of CO2 -driven laser–plasmas should, in                     vides a numerically cost-efficient means of incorporating non-
theory, employ the CR approach described in section 2.                           LTE opacities (and EOS quantities) into hydrodynamic codes,
    As outlined in the introduction, the goal of the present                     Klapisch and Bar-Shalom have stressed that it should only be
work is to evaluate the role of multiply-excited states in CO2                   applied to plasmas not far from pure LTE conditions [66].
laser-driven plasma conditions. As such, we must inherit the                        We wish to test the validity this method in the context of
complicated level structures considered in reference [13]. As                    the current problem. Guided by previous modelling efforts of
discussed in reference [13, 39], the use of extensive atomic                     Sasaki and co-workers [37, 40, 41], we have undertaken a
models for Sn11+ –Sn14+ is necessary to ensure (i) correct                       non-LTE configuration-averaged (CA) Atomic calculation at
wavelength positions of Δn = 0 transitions (which are sensi-                     ρ = 2 × 10−4 g cm−3 and T e = 36 eV. The electron density of
tive to the choice of underlying atomic structure as a result of                 this plasma is ne = 1.27 × 1019 cm−3 , close to the critical elec-
CI) and (ii) partition function convergence [56]. Unfortunately,                 tron density for CO2 laser light (ncrit,CO2 = 1.1 × 1019 cm−3 ).

                                                                             4
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                 J Sheil et al

                                                                        CSD is possible using a T e = 26 eV non-LTE calculation,
                                                                        however one must include an external radiation field in the
                                                                        calculation characterised by T r = T e . In this way, all atomic
                                                                        processes are in detailed balance. However, as detailed above,
                                                                        our non-LTE calculations do not include an external radia-
                                                                        tion field. In order to match our present calculations with a
                                                                        T Z = 26 eV LTE calculation, we must therefore compensate
                                                                        for the missing (de)excitation and ionisation channels pro-
                                                                        vided by photoexcitation and photoionisation by increasing
                                                                        T e above T Z . The difference T e − T Z is effectively a gauge
                                                                        for how close the system is to LTE conditions; in the work
                                                                        of Torretti et al [13], for example, a value T e − T Z = 2.5 eV
                                                                        (where T e = 34.5 eV and T Z = 32 eV) highlights the domi-
Figure 1. Charge-state distributions for plasmas calculated in
                                                                        nance of collisional rates over radiative rates at near-LTE con-
non-LTE (T e = 36 eV; shown in black) and LTE (T Z = 26 eV;             ditions (ne = 1.27 × 1020 cm−3). The same cannot be said for
shown in red) conditions. The plasma mass density used in both          the CO2 -relevant conditions, where a value T e − T Z = 10 eV
calculations was ρ = 2 × 10−4 g cm−3 .                                  is indicative of truly non-LTE conditions, i.e. collisional rates
                                                                        do not significantly outweigh radiative rates.
The CA approach generates appropriately averaged ener-                  3.2. Population distributions and effective temperatures
gies, oscillator strengths and PWB electron-impact excita-
tion cross sections from which configuration-to-configuration           The replication of a non-LTE CSD using LTE computations,
rates are derived [59]. This approach is employed in the                such as that shown in figure 1, is key for justifying the use
present context to ensure computational tractability. Config-           of LTE opacities as an approximation for non-LTE opacities.
urations sets employed in the CA CATS calculations for                  An even stronger argument in favour of this approximation
Sn12+ and Sn14+ are provided in table 1. The calculations               is to establish equivalence between the non-LTE and LTE-
were performed under steady-state conditions, i.e. dN/dt = 0.           computed configuration populations (the cumulative sum of
This approximation is valid for plasmas where the time taken            which yields the CSD—see equation (6)). To this end, we
to generate a given CSD (through ionisation and recombi-                present in figure 2 non-LTE- and LTE-computed configuration
nation processes) occurs on timescales shorter than the laser           populations (Pnon−LTE, PLTE ) as a function of CA energy (Eav )
pulse duration. This is certainly true for industrial LPP EUV           for Sn12+ (upper panel) and Sn14+ (lower panel). The non-LTE
sources, which are typically driven by multi-10 nanosecond-             configuration population calculations are shown in figure 2(a)
long CO2 pulses. We calculated the rates of electron-impact             (Sn12+) and figure 2(d) (Sn14+ ). The corresponding LTE cal-
excitation/ionisation and AI (as well as their inverse pro-             culations are shown in figures 2(b) and (e). Configurations
cesses) assuming a Maxwell–Boltzmann EEDF at T e = 36 eV.               associated with single-, double- and triple-electron excitations
The calculations were performed without an external radiation           are shown as blue, orange and green circles, respectively. The
field, i.e. photoexcitation and photoionisation processes were          ground-state configurations, fixed at Eav = 0 eV, are shown as
not considered (see subsection 3.2 below for further discus-            grey circles. Perhaps the most striking feature of these calcula-
sion). The process of spontaneous emission was included in              tions is the near-coincidence in non-LTE- and LTE-computed
the model.                                                              configuration populations. This is quantified in figures 2(c)
    The CSD resulting from this calculation is shown in black in        and (f ), where, for a given configuration C, we plot the non-
figure 1. Importantly, the relevant EUV-emitting Sn11+ –Sn14+           LTE-normalised Configuration Population Difference (CPD)
ions are produced under such conditions, with Sn12+ and                 defined as (PC,non−LTE − PC,LTE )/PC,non−LTE . For the majority
Sn13+ clearly the dominant species. The average charge                  of configurations, this quantity takes on values below 1%. This
state of this plasma is Znon−LTE = 12.6. Also illustrated in          indicates a near-coincidence in non-LTE and LTE-computed
figure 1 is an LTE CA calculation (shown in red) performed at           configuration populations. The few configurations where the
ρ = 2 × 10−4 g cm−3 and T e = 26 eV. These calculations yield           non-LTE-normalised CPD values are above 1% all have LTE-
ZLTE = 12.6, identical to the non-LTE calculation. Also, we           computed populations greater than the corresponding non-LTE
note that the CSDs are very nearly identical. An ionisation tem-        populations.
perature of T Z = 26 eV is thus established. We note that the               In addition to validating the use of Busquet’s approach in
ratio T e /T Z = 36/26 ≈ 1.4 is larger than that obtained by the        the current context, the equivalence in configuration popu-
scaling law Te /TZ ≈ (1 + (5.36 × 1013 )Te n−1
                                              7/2                       lations reveals an interesting characteristic of the non-LTE-
                                                  e )
                                                      0.215
                                                            ≈ 1.2
obtained by Busquet et al [65]. Much better agreement with              computed populations in that they appear to follow a single
this scaling law is obtained in the Nd:YAG-relevant case                Boltzmann-esque temperature law of the form
considered in [13] (ne = 1.27 × 1020 cm−3 , T Z = 32 eV,                                                                
                                                                                                                     E
T e = 34.5 eV), where T e /T Z = 34.5/32 ≈ 1.08 and the scal-                            PC,non−LTE ∝ gC exp −             ,            (8)
                                                                                                                    Teff
ing law yields T e /T Z ≈ 1.02.                                                       
    It is interesting to consider why optimum agreement in              where gC = JC (2JC + 1) is the total statistical weight of con-
CSDs is obtained for T e > T Z . Exact reproduction of an LTE           figuration C (J C is an energy level associated with C) and T eff is

                                                                    5
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                      J Sheil et al

Figure 2. Configuration populations for Sn12+ (top) and Sn14+ (bottom) calculated for a T e = 36 eV non-LTE plasma [panels (a) and (d)]
and a T e = 26 eV LTE [panels (b) and (e)] plasma. The plasma mass density used in both calculations was ρ = 2 × 10−4 g cm−3 .
Configurations associated with single-, double- and triple-electron excitations are shown as blue, orange and green circles, respectively. The
ground-state configurations are shown in grey and are fixed at Eav = 0 eV. Non-LTE-normalised configuration population differences
(CPDs) (see text for description) are shown in panels (c) and (f). The vertical red lines indicate the ionisation potentials of Sn12+ and Sn14+ .

a so-called effective temperature [67] (the Boltzmann constant             states typically decay towards high/low-lying lower states)
kB typically appearing in equations of the form equation (8) is            tend to generate Boltzmann-type population distributions
here ‘absorbed’ into the definition of T eff ).                            [71]. This is in fact a general phenomenon of non-LTE hot
    In order to evaluate T eff , we plot in figure 3 the natural           plasmas [68]. Finally, it should be noted that the approach
logarithm of the ratio PC,non−LTE/gC as a function of Eav for              of Bauche et al [70], does not ascribe a single effective
(a) Sn12+ and (b) Sn14+ . First-order polynomial fits to the               temperature T eff to a population distribution. Rather, con-
data are shown in red. Excellent fits to the data are charac-              figurations are grouped according to superconfigurations
terised by R2 values of 0.99 attained in both cases. The effec-            and equation (8) is used to derive superconfiguration-
tive temperatures T eff are obtained from the inverse-negative             specific temperatures. These temperatures may vary
of the slopes of the lines of best fit. As indicated in the                significantly from superconfiguration-to-superconfiguration
plots, the fits yield values of T eff = 26.6 (Sn12+ ) and 26.9 eV
                                                                           (see figure 1 of [70]). Although this approach differs to that
(Sn14+ ), clearly very similar to the LTE ionisation temperature
                                                                           presented in figure 3 (where we have performed a ‘global’ fit
T Z = 26 eV. This close agreement between T eff and T Z is per-
                                                                           to all configurations), the non-LTE-computed population dis-
haps not too surprizing given the near-coincidence in non-LTE
                                                                           tributions are clearly well-described by a single-temperature
and LTE-computed configuration populations.
    This observation, that non-LTE-computed configuration                  Boltzmann-type law.
populations follow Boltzmann-like temperature laws has been                     Finally, we wish to comment on the importance of a non-
studied extensively in the works of Bauche, Bauche-Arnoult                 zero radiation field in our non-LTE computations. As out-
and others [67–72]. The emergence of temperature laws                      lined in section 3, all non-LTE calculations presented thus far
is attributed to the complex interplay of atomic processes                 have not included an external radiation field in the underly-
defining the steady-state CR model. More specifically, the                 ing atomic kinetics. Indeed, identifying an external radiation
combination of (i) the microreversibility nature of colli-                 field appropriate for the current study is rather difficult as it
sional processes (which drives the population distributions                is a purely non-local quantity, i.e. it depends on the radia-
towards LTE conditions) and (ii) correlation laws asso-                    tive characteristics of the plasma surrounding the studied (ρ,
ciated with spontaneous emission (high/low-lying upper                     T e ) plasma. Proper evaluation of an external radiation field

                                                                       6
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                        J Sheil et al

Figure 3. Logarithm of the statistical-weight-normalised                     Figure 4. Logarithm of the statistical-weight-normalised
non-LTE-computed configuration populations for (a) Sn12+ and                 non-LTE-computed configuration populations for (a) Sn12+ and
(b) Sn14+ . The plasma conditions correspond to ρ = 2 × 10−4 g               (b) Sn14+ . The plasma conditions correspond to ρ = 2 × 10−4 g
cm−3 , T e = 36 eV and T r = 0 eV. A first-order polynomial fit to the       cm−3 , T e = 33 eV and T r = 21 eV. A first-order polynomial fit to
data is shown in red. Effective temperatures T eff are obtained from         the data is shown in red. Effective temperatures T eff are obtained
the slopes of the lines of best fit.                                         from the slopes of the lines of best fit.

requires complex simulations involving the coupling of non-                  4. Tin-ion opacities
LTE atomic kinetics with a model for radiation transport in
a multi-(ρ, T e , T r ) plasma. Such an endeavour is outside the             We now present the results of detailed, level-resolved LTE
scope of the present work. Instead, we have performed an addi-               opacity calculations at CO2 laser-driven plasma conditions.
tional non-LTE calculation where we have included a Planck-                  The LTE plasma conditions considered here are the same
ian radiation field (characterised by a radiation field T r ). We            as those of section 3 (ρ = 2 × 10−4 g cm−3 , T Z = 26 eV).
fixed T r and tuned T e until equation (7) is established with               Unlike section 3, we adopt a fully level-resolved approach
a T Z = 26 eV LTE plasma. We find that a T e = 33 eV and                     where Atomic is used to calculate level populations and
T r = 21 eV plasma, for example, reproduces both equation (7)                bound–bound opacities. The atomic structures employed in
and the LTE CSD. Moreover, the configuration populations                     these calculations are the same as those considered in reference
associated with Sn12+ and Sn14+ are also found to be in excel-               [13] (see section 2).
lent agreement with the LTE-computed configuration popu-                        The results of these calculations are presented in figure 5.
lations. This is demonstrated in figure 4, where, in a similar               Panels (a)–(d) illustrate the contributions to the opacity from
vein to figure 3, we plot the natural logarithm of the ratio                 the four charge states Sn11+ –Sn14+ . In a similar fashion to
PC,non−LTE/gC as a function of Eav for (a) Sn12+ and (b) Sn14+ .             Torretti et al [13], the different contributions to the opac-
Effective temperatures obtained from the slopes of the lines of              ity are plotted in a cumulative fashion; photoabsorption from
best fit (shown in red) are T eff = 26.6 eV for both Sn12+ and               levels in the ground manifold to singly-excited levels are
Sn14+ , clearly in excellent agreement with the data presented               shown in blue (we call this the ‘ground opacity’); the summa-
in figure 3. The introduction of a non-zero radiation field                  tion of photoabsorption from singly-excited levels to doubly-
in these computations therefore does not break the observed                  excited levels (‘first opacity’) and the ground opacity is shown
Boltzmann-esque temperature law once equivalence between                     by the line bounding the orange shaded area; the summa-
the non-LTE and LTE-computed CSDs is established.                            tion of photoabsorption from doubly- and triply-excited levels

                                                                         7
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                         J Sheil et al

                                                                               Figure 6. Panels (a)–(d): contributions from the ground, first and
                                                                               higher opacities in the 2% bw in Sn11+ –Sn14+ . The total opacity in
                                                                               the 2% bw, obtained by summing the contributions in panels
Figure 5. Panels (a)–(d): opacities of the ions Sn11+ –Sn14+ . The             (a)–(d), is shown in panel (e).
opacities are plotted in a cumulative fashion. Photoabsorption from
levels of the ground manifold to singly-excited levels is shown in
blue (‘ground opacity’). The summation of photoabsorption from
singly-excited levels to doubly-excited levels (‘first opacity’) and the       any other transition in Sn11+ –Sn14+ . There are two reasons
ground opacity is shown by the line bounding the orange shaded                 for this spectral dominance. First, as illustrated in figure 2,
region. The line bounding the green shaded region is the summation
of photoabsorption from doubly- and triply-excited levels (‘higher
                                                                               a significant fraction of Sn12+ and Sn14+ ions are in their
opacity’), the ground opacity and first opacity. The total opacity,            ground-state configurations. This also true for Sn13+ . Sec-
obtained by summing the contributions of Sn11+ –Sn14+ , is shown in            ondly, this LS-allowed transition exhibits a large oscillator
panel (e). The inset in panel (e) shows the opacity in a 2% bw                 strength. This combination, which appears multiplicatively in
centred at 13.5 nm.                                                            the first term of equation (2), is the origin of this large opacity
                                                                               value. As in the case of Sn11+ and Sn12+ , higher-order opaci-
(‘higher opacity’), the ground opacity and first opacity is                    ties are seen to contribute at wavelengths above 13.5 nm, how-
shown by the line bounding the green shaded area. The grey                     ever only on the order κ ≈ 10 × 105 cm2 g−1 . Moving to the
region in each panel represents a 2% bw centred at 13.5 nm.                    closed 4p subshell Sn14+ ion we observe an interesting shift
    First, examining the ordinate of figures 5(b) and (c), one                 in opacity dominance away from the ground opacity to the
readily concludes that Sn12+ and Sn13+ are the dominant con-                   first opacity. The ground opacity in this wavelength region is
tributors to the opacity in this plasma. This can be under-                    dominated by the transition 4p6 1 S0 –4p5 4d 1 P1 , which has a
stood from the CSD presented in figure 1, where these two                      calculated wavelength of 13.332 nm. This is in excellent agree-
charge states accumulate nearly 80% of the total ion popula-                   ment with experimental identifications of this transition at
tion. Although the majority of the opacity in the Sn12+ ion                    λ = 13.344 nm [29] and λ = 13.343 [18, 74]. This transition
can be attributed to the ground opacity, the first and higher                  is clearly much weaker than the 4p6 4d 2 D5/2 − 4p5 4d2 2 F7/2
opacities contribute to the spectrum at wavelengths larger than                transition in Sn13+ . This is because (i) the 4p6 1 S0 level in
13.5 nm. The same is true for Sn11+ . One of the most strik-                   Sn14+ has a smaller population than the 4p6 4d 2 D5/2 level in
ing features of these opacity spectra is the spectral dominance                Sn13+ and (ii) the transition in Sn14+ involves energy levels
of a transition in Sn13+ [see figure 5 (c)]. This transition is                having low angular momenta (J lower,upper = 0, 1) and therefore
the subvalence excitation 4p6 4d 2 D5/2 −4p5 4d2 2 F7/2 . It has a             has a lower oscillator strength than the aforementioned tran-
calculated wavelength of 13.286 nm, in good agreement with                     sition in Sn13+ . As is clear from figure 5(d), photoabsorp-
experimental identifications of this transition at 13.3014 [17],               tion from singly-excited levels is the main source of opacity
13.3020 [73] and 13.301 nm [29]. The opacity of this transi-                   in Sn14+ . Importantly, all of the opacity in the industrially-
tion is κ ≈ 83 × 105 cm2 g−1 , nearly a factor 3 higher than                   relevant 2% bw (the grey shaded region) in Sn14+ origi-

                                                                           8
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                   J Sheil et al

                                                                           this figure that the population of multiply-excited (doubly-
                                                                           , triply- and quadruply-excited) states is much higher in the
                                                                           T e = 32 eV plasma than in the T e = 26 eV plasma. Examples
                                                                           of this ratio for the 4p5 4d3 , 4p4 4d4 and 4p3 4d5 configurations
                                                                           in Sn12+ are shown in blue, orange and green, respectively.
                                                                           Interestingly, the increased population of the multiply-excited
                                                                           states in the Nd:YAG-relevant plasma is a natural by-product
                                                                           of the higher temperatures needed to generate a CSD equiv-
                                                                           alent to that generated under relevant CO2 -laser-driven tin
                                                                           plasma conditions.

                                                                           5. Conclusion
Figure 7. Ratio of the partition function factor for T e = 32 and
26 eV plasmas (shown in black). The ratio for the 4p5 4d3 , 4p4 4d 4       In conclusion, we have investigated the role played by
and 4p3 4d5 configurations in Sn12+ are shown in blue, orange and          multiply-excited states in industrially-relevant CO2 laser-
green, respectively.
                                                                           driven tin plasma conditions. First, we employed Busquet’s
                                                                           ionisation temperature method to match the charge state dis-
                                                                           tribution of a non-LTE plasma with an LTE computation. We
nates from higher-order opacities, the first opacity being the             find excellent agreement between the non-LTE-computed con-
dominant contributor.                                                      figuration populations and LTE calculations for both Sn12+
    The total opacity, obtained by summing the contributions               and Sn14+ . Moreover, we identify the existence of Boltzmann-
in figures 5(a)–(d), is shown in figure 5(e). As expected, most            like temperature laws in the non-LTE population distributions.
of the higher-order opacity contributions are located at wave-             In the second part of this study we performed detailed, level-
lengths above 13.5 nm. The inset illustrates the contributions to          resolved LTE opacity calculations at plasma conditions per-
the opacity in the 2% bw. Higher-order opacities clearly make              taining to CO2 laser-driven plasma conditions. Approximately
a significant contribution to this industrially-relevant spectral          66% of the opacity in the 2% bw centred at 13.5 nm was found
region. In essence, they tend to ‘fill out’ regions between the            to arise from transitions between multiply-excited states. This
ground opacities. The charge-state-specific opacity contribu-              result highlights the important role played by multiply-excited
tions in the 2% bw are illustrated in figures 6(a)–(d), with               states in industrially-relevant plasma conditions. This work
their combined opacity contribution illustrated in figure 6(e).            paves the way for the generation of detailed LTE opacity tables
Importantly, only 34% of the opacity arises from the ground                at numerous (ρ, T e , T r ) non-LTE plasma conditions for incor-
opacity. Of the remaining 66%, approximately 42% originates                poration in radiation-hydrodynamic simulations. Such simu-
from the first opacity and 24% is associated with the higher               lations will enable reliable predictions of EUV emission from
opacity. This result clearly highlights the important spectral             LPPs and are therefore key for guiding experimental efforts in
role played by multiply-excited transitions in CO2 laser-driven            characterising and optimising laser-driven plasma sources of
plasma conditions.                                                         EUV light.
    Finally, we wish to compare these results with the Nd:YAG-
relevant case considered in [13]. The plasma conditions con-
sidered in reference [13] were ρ = 2 × 10−3 g cm−3 and                     Acknowledgments
T e = 32 eV. This LTE calculation yielded ne = 1.27 ×
1020 cm−3 (an order of magnitude higher than the current case)             We would like to thank W van der Zande, W Ubachs and R
and an average charge state Z similar to the present case.               Hoekstra for their useful comments on the paper. This project
Comparing figure 5(e) with figure 3(e) of [13], we see that,               has received funding from European Research Council (ERC)
relative to the ground opacity, the higher-order opacities in              Starting Grant No. 802648 and is part of the VIDI research
the Nd:YAG case make a significantly larger contribution than              programme with Project No. 15697, which is financed by the
in the present case. Comparing Maxwell–Boltzmann distri-                   Netherlands Organization for Scientific Research (NWO). Part
butions at T e = 32/26 eV, one observes that the former has                of this work has been carried out at the Advanced Research
a larger fraction of higher-energy electrons which can colli-              Center for Nanolithography (ARCNL), a public-private part-
sionally populate high-lying levels associated with doubly-,               nership of the University of Amsterdam (UvA), the Vrije
triply- and quadruply-excited configurations. As both calcula-             Universiteit Amsterdam (VU), NWO and the semiconduc-
tions were performed in LTE, the populations of the levels are             tor equipment manufacturer ASML. Part of this work was
described by the so-called ‘partition function factor’, which              supported by the ASC PEM programme of the US Depart-
for some level J is written gJ exp(−EJ /T e ). To enable a com-            ment of Energy through the Los Alamos National Labora-
parison between both cases, we plot in figure 7 the ratio of               tory. Los Alamos National Laboratory is operated by Triad
the partition function factor for T e = 32 and 26 eV (shown                National Security, LLC, for the National Nuclear Security
in black) using a continuous energy variable E. The statisti-              Administration of US Department of Energy (Contract No.
cal weight gJ naturally drops out of this ratio. It is clear from          89233218NCA000001).

                                                                       9
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                                     J Sheil et al

ORCID iDs                                                                  [34] MacFarlane J J, Rettig C L, Wang P, Golovkin I E and
                                                                                   Woodruff P R 2005 Emerging Lithographic Technologies IX
J Sheil https://orcid.org/0000-0003-3393-9658                                      (International Society for Optics and Photonics SPIE) vol
                                                                                   5751 ed R S Mackay pp 588–600
O O Versolato https://orcid.org/0000-0003-3852-5227                        [35] Koike F and Fritzsche S 2007 Radiat. Phys. Chem. 76 404
J Colgan https://orcid.org/0000-0003-1045-3858                             [36] de Gaufridy de Dortan F 2007 J. Phys. B: At. Mol. Opt. Phys.
                                                                                   40 599
                                                                           [37] Sasaki A, Sunahara A, Furukawa H, Nishihara K, Fujioka S,
References                                                                         Nishikawa T, Koike F, Ohashi H and Tanuma H 2010 J. Appl.
                                                                                   Phys. 107 113303
 [1] Bleiner D, Costello J, Dortan F, O’Sullivan G, Pina L and             [38] Zeng J, Gao C and Yuan J 2010 Phys. Rev. E 82 026409
       Michette A (ed) 2014 Short Wavelength Laboratory Sources:           [39] Colgan J, Kilcrease D P, Abdallah J, Sherrill M E, Fontes C
       Principles and Practices (London: The Royal Society of                      J, Hakel P and Armstrong G S J 2017 High Energy Density
       Chemistry)                                                                  Phys. 23 133
 [2] Schneider G, Guttmann P, Heim S, Rehbein S, Mueller F,                [40] Sasaki A, Sunahara A, Nishihawra K, Nishikawa T, Koike F and
       Nagashima K, Heymann J B, Müller W G and McNally J G                        Tanuma H 2008 J. Phys.: Conf. Ser. 112 042062
       2010 Nat. Methods 7 985                                             [41] Sasaki A, Sunahara A, Nishihara K, Nishikawa T, Fujima K,
 [3] McDermott G, Le Gros M A, Knoechel C G, Uchida M and                          Kagawa T, Koike F and Tanuma H 2007 High Energy Density
       Larabell C A 2009 Trends Cell Biol. 19 587                                  Phys. 3 250
 [4] Helk T, Zürch M and Spielmann C 2019 Struct. Dyn. 6 010902            [42] Fontes C J et al 2015 J. Phys. B: At. Mol. Opt. Phys. 48 144014
 [5] Sakdinawat A and Attwood D 2010 Nat. Photon. 4 840                    [43] Magee N H, Abdallah J, Colgan J, Hakel P, Kilcrease D P,
 [6] Bartnik A, Fiedorowicz H, Jarocki R, Kostecki J, Szczurek M                   Mazevet S, Sherrill M, Fontes C J and Zhang H 2004 AIP
       and Wachulak P W 2011 Nucl. Instrum. Methods Phys. Res.                     Conf. Proc. 730 168–79
       A 647 125                                                           [44] Hakel P, Sherrill M E, Mazevet S, Abdallah J, Colgan J,
 [7] Mack C 2007 Fundamental Principles of Optical Lithography:                    Kilcrease D P, Magee N H, Fontes C J and Zhang H L 2006
       The Science of Microfabrication (New York: Wiley)                           J. Quant. Spectrosc. Radiat. Transfer 99 265
 [8] Garner C M 2012 Phil. Trans. R. Soc. A 370 4015                       [45] Abdallah J J, Clark R E H and Cowan R D 1988 Theoretical
 [9] van Setten E, Schiffelers G, Psara E, Oorschot D, Davydova N,                 atomic physics code development I CATS: Cowan Atomic
       Finders J, Depre L and Farys V 2014 30th European Mask                      Structure Code Los Alamos National Laboratory Report LA-
       and Lithography Conf. ed U F W Behringer (International                     11436-M (unpublished)
       Society for Optics and Photonics SPIE) vol 9231 pp 50–63            [46] Cowan R D 1981 The Theory of Atomic Structure and Spectra
[10] Wagner C and Harned N 2010 Nat. Photon. 4 24                                  (Berkeley, CA: University of California Press)
[11] Bakshi V (ed) 2009 EUV Lithography (Washington: SPIE Press)           [47] Kramida A 2019 Atoms 7 64
[12] Bakshi V (ed) 2018 EUV Lithography 2nd edn (Washington:               [48] Wyart J-F 2011 Can. J. Phys. 89 451
       SPIE Press)                                                         [49] Ryabtsev A, Kononov E, Kildiyarova R, Tchang-Brillet W-Ü,
[13] Torretti F et al 2020 Nat. Commun. 11 2334                                    Wyart J-F, Champion N and Blaess C 2015 Atoms 3 273
[14] Fomenkov I et al 2017 Adv. Opt. Technol. 6 173                        [50] Windberger A et al 2016 Phys. Rev. A 94 012506
[15] Purvis M et al 2018 Extreme Ultraviolet (EUV) Lithography             [51] Torretti F et al 2017 Phys. Rev. A 95 042503
       IX (International Society for Optics and Photonics SPIE) vol        [52] Hubeny I and Mihalas D 2015 Theory of Stellar Atmospheres:
       10583 ed K A Goldberg pp 476–85                                             An Introduction to Astrophysical Non-equilibrium Quantita-
[16] Churilov S S and Ryabtsev A N 2006 Phys. Scr. 73 614                          tive Spectroscopic Analysis (Princeton, NJ: Princeton Univer-
[17] Churilov S S and Ryabtsev A N 2006 Opt. Spectrosc. 101 169                    sity Press)
[18] Ryabtsev A N, Kononov É Y and Churilov S S 2008 Opt.                  [53] Castor J 2004 Radiation Hydrodynamics (Cambridge: Cam-
       Spectrosc. 105 844                                                          bridge University Press)
[19] O’Sullivan G et al 2015 J. Phys. B 48 144025                          [54] Ralchenko Y 2013 Plasma Fusion Res. 8 2503024
[20] Versolato O O 2019 Plasma Sources Sci. Technol. 28 083001             [55] Scott H A 2016 Modern Methods in Collisional-Radiative
[21] Richardson M, Koay C-S, Takenoshita K, Keyser C and Al-                       Modeling of Plasmas ed Y Ralchenko (Berlin: Springer)
       Rabban M 2004 J. Vac. Sci. Technol. B 22 785                                pp 81–104
[22] Felix N M and Attwood D T 2020 Extreme Ultraviolet (EUV)              [56] Hakel P and Kilcrease D P 2004 AIP Conf. Proc. 730
       Lithography XI (International Society for Optics and Photon-                 90–199
       ics SPIE) vol 11323 ed N M Felix and A Lio (https://doi.org/        [57] Kilcrease D P, Colgan J, Hakel P, Fontes C J and Sherrill M E
       10.1117/12.2572271)                                                         2015 High Energy Density Phys. 16 36
[23] O’Sullivan G and Carroll P K 1981 J. Opt. Soc. Am. 71 227             [58] Fontes C J, Jr J A Jr, Clark R E H and Kilcrease D P 2000 J.
[24] Mandelbaum P, Finkenthal M, Schwob J L and Klapisch M 1987                    Quant. Spectrosc. Radiat. Transfer 65 223
       Phys. Rev. A 35 5051                                                [59] Fontes C J, Colgan J and Abdallah J 2016 Modern Methods in
[25] Svendsen W and O’Sullivan G 1994 Phys. Rev. A 50 3710                         Collisional-Radiative Modeling of Plasmas ed Y Ralchenko
[26] Ohashi H, Suda S, Tanuma H, Fujioka S, Nishimura H, Sasaki                    (Berlin: Springer) pp 17–50
       A and Nishihara K 2010 J. Phys. B: At. Mol. Opt. Phys. 43           [60] Rose S J 1995 J. Quant. Spectrosc. Radiat. Transfer 54 333
       065204                                                              [61] Rodríguez R, Espinosa G and Gil J M 2018 Phys. Rev. E 98
[27] D’Arcy R et al 2009 Phys. Rev. A 79 042509                                    033213
[28] D’Arcy R et al 2011 Phys. Scr. T144 014026                            [62] Clark R E H, Abdallah J J and Mann J B 1991 Astrophys. J. 381
[29] Scheers J et al 2020 Phys. Rev. A 101 062511                                  597
[30] Bakshi V (ed) 2006 EUV Sources for Lithography (Washington:           [63] Lysaght M, Kilbane D, Murphy N, Cummings A, Dunne P and
       SPIE Press)                                                                 O’Sullivan G 2005 Phys. Rev. A 72 014502
[31] White J, Hayden P, Dunne P, Cummings A, Murphy N, Sheridan            [64] Busquet M 1993 Phys. Fluids B 5 4191
       P and O’Sullivan G 2005 J. Appl. Phys. 98 113301                    [65] Busquet M, Colombant D, Klapisch M, Fyfe D and Gardner J
[32] Fujioka S et al 2005 Phys. Rev. Lett. 95 235004                               2009 High Energy Density Phys. 5 270
[33] Poirier M, Blenski T, de Gaufridy de Dortan F and Gilleron F          [66] Klapisch M and Bar-Shalom A 1997 J. Quant. Spectrosc.
       2006 J. Quant. Spectrosc. Radiat. Transfer 99 482                           Radiat. Transfer 58 687

                                                                      10
J. Phys. B: At. Mol. Opt. Phys. 54 (2021) 035002                                                                              J Sheil et al

[67] Bauche J and Bauche-Arnoult C 2000 J. Phys. B: At. Mol. Opt.        [71] Bauche J, Bauche-Arnoult C and Peyrusse O 2006 J. Quant.
       Phys. 33 L283                                                            Spectrosc. Radiat. Transfer 99 55
[68] Fournier K B, Bauche J and Bauche-Arnoult C 2000 J. Phys. B:        [72] Busquet M 2006 J. Quant. Spectrosc. Radiat. Transfer 99
       At. Mol. Opt. Phys. 33 4891                                              131
[69] Bauche-Arnoult C and Bauche J 2001 J. Quant. Spectrosc.             [73] Sugar J, Rowan W L and Kaufman V 1992 J. Opt. Soc. Am. B 9
       Radiat. Transfer 71 189                                                  1959
[70] Bauche J, Bauche-Arnoult C and Fournier K B 2004 Phys. Rev.         [74] Sugar J, Rowan W L and Kaufman V 1991 J. Opt. Soc. Am. B 8
       E 69 026403                                                              2026

                                                                    11
You can also read