Paper The following article is Open access

Multiply-excited states and their contribution to opacity in CO2 laser-driven tin-plasma conditions

, , and

Published 22 January 2021 © 2021 IOP Publishing Ltd
, , Citation J Sheil et al 2021 J. Phys. B: At. Mol. Opt. Phys. 54 035002 DOI 10.1088/1361-6455/abcedf

0953-4075/54/3/035002

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 TZ 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 TeffTZ. In the second part of this work, we perform extensive level-resolved LTE opacity calculations at TZ. 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.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 4.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Powerful sources of short-wavelength radiation play a crucial role in many high-profile scientific and industrial endeavours [1]. Examples include high-resolution biological imaging [24], materials science [5, 6] and the manufacturing of integrated circuits (ICs) in the semiconductor industry [7, 8]. The use of short-wavelength radiation in this latter application is expected to enable the printing of sub-10 nm feature sizes on ICs [9], thus sustaining the production of cheaper and more powerful computer chips into the future.

The key technology driving feature-size miniaturisation in the semiconductor industry today is extreme ultraviolet lithography (EUVL) [1013]. As the name suggests, this technology uses EUV light to print features on ICs. In an industrial setting, this EUV light is generated in a laser-produced plasma (LPP) formed by irradiating tin microdroplets with CO2 laser light (λ = 10.6 μm) [14, 15]. The ensuing laser–droplet interaction generates a plasma having moderate electron temperatures and densities (Te ≈ 20–40 eV, ne ≈ 1018–1019 cm−3) comprising large populations of Sn8+ – Sn14+ ions. Spectra recorded from such plasmas are known to exhibit an intense, narrowband feature peaked near 13.5 nm [1620]. Importantly, this emission feature provides optimum wavelength overlap with the 2% reflective bandwidth (bw) centred at 13.5 nm defined by molybdenum/silicon multilayer mirrors (MLMs) employed in lithography tools [12]. This combination of (i) optimum emission characteristics and (ii) a debris-limited droplet-target (key for minimising MLM contamination from plasma expansion [21]) established tin LPPs as the light source of choice for EUVL. In fact, after more than 20 years of research and development, EUVL has now entered the realm of high-volume manufacturing of new-generation ICs [22].

The aforementioned intense, narrowband emission feature observed in laser-produced tin plasmas has, to date, been predominantly attributed to overlapping Δn = 0 transition arrays of the form 4p64dm − (4p54dm+1 + 4p64dm−14f) in charge states Sn8+–Sn14+ (m = 6–0) [2325]. Such claims have been supported by charge-state-resolved spectroscopic experiments [2629] as well as numerous theoretical efforts employing advanced atomic structure codes [3039]. Although serving as the general consensus for many years, a recent study of Torretti et al has challenged this viewpoint [13]. In the most detailed tin-ion opacity calculations performed to-date, it was found that transitions between multiply-excited configurations in Sn11+–Sn14+ are in fact the dominant contributors to EUV emission in Nd:YAG-driven (λ = 1.064 μm) tin LPPs. Surprisingly, the well-known 4p64dm − (4p54dm+1 + 4p64dm−14f) transitions play only a minor role; they comprise 11% of the opacity in the 5–20 nm region and 19% of the opacity in the 2% bw. Remarkably, it is transitions between complex, multiply-excited configurations (doubly-, triply- and quadruply-excited configurations) which dominate the opacity spectrum. Strong support for their spectroscopic dominance was evidenced through comparisons of experimental tin-droplet EUV spectra with one-dimensional radiation transport simulations. The incorporation of multiply-excited transitions in these simulations was deemed crucial for the reproduction of the experimental spectra.

The purpose of the present paper is to elucidate the role of multiply-excited states in CO2-driven tin plasmas which are currently used in the industry to provide light for EUVL. The main difference between the present work and previous atomic modelling work of Sasaki and co-authors [37, 40, 41] is the extensivity of the atomic structures employed. For example, in reference [37, 40, 41], a less extensive number of doubly-excited states were included in their atomic models. Their inclusion was mainly for the purpose of understanding satellite transition array emission of the form 4p64dm−1 nl − (4p54dm nl + 4p64dm−24fnl). The configuration sets adopted in the present work include large numbers of doubly-, triply- and quadruply-excited configurations and are therefore significantly more extensive than those considered previously. Importantly, we find that transitions between multiply-excited states make a substantial contribution to the opacity in the 5–20 nm region in plasma conditions pertinent to those generated in CO2 laser-driven tin plasmas.

The current paper is organised as follows: in section 2, we describe the atomic structure and opacity calculations performed with the Los Alamos suite of atomic physics and population kinetics codes. An examination of non-LTE- and LTE-computed tin ionisation distributions and configuration populations in CO2 laser-driven plasma conditions follows in section 3. In section 4 we present level-resolved LTE-computed opacities of Sn11+–Sn14+. Finally, a summary of this work is given in section 5.

2. Atomic structure and opacity calculations

In the present work, we have employed the Los Alamos national laboratory (LANL) suite of atomic structure and collision codes [42] as well as the population kinetics code Atomic [43, 44] to calculate the opacities of a tin plasma at densities and temperatures relevant to those attained in a CO2 laser-driven plasma. Given the central role that these codes play in this study, we devote the following subsections to their description.

2.1. Atomic structure calculations

The first step in the present opacity calculations is the generation of all relevant atomic data (energy levels, oscillator strengths, ionisation potentials, etc). In Torretti et al [13], considerable effort was devoted to benchmarking the calculated atomic structures with available experimental data and therefore we have utilised these atomic structures in the present study. In brief, the Cats code [45] was employed in the calculation of level energies and absorption oscillator strengths. This code is based on Cowan's original atomic structure code [46, 47] and exhibits a favourable feature where one can scale the magnitude of the radial integrals in order to bring theoretical level structures into excellent agreement with experimental observations. This semi-empirical approach has proven to be an indispensable tool in atomic spectroscopy, especially for systems exhibiting complicated level structures [29, 4851]. In the work of Torretti et al [13], optimum agreement between atomic structure calculations and previous experimental observations [17, 26, 27], was found by applying scaling factors of 0.87/0.75 to the radial integrals associated with singly-/multiply-excited energy levels. These two scaling factors were adopted for all charge states considered in this work. Extensive configuration sets were employed in the calculations through use of the so-called '2-mode' approach. In these calculations, configuration interaction (CI) was retained for a large number of configurations associated with Δn = 0 and Δn = 1 single, double and triple excitations from the n = 4 manifold of the ground-state configuration. Additional configurations were included using intermediate coupling (only interactions between levels of the same configuration are considered) [39]. This approach retains a significant level of detail provided by full CI calculations however at a much reduced computational cost.

2.2. Opacity calculations

Knowledge of opacities is crucial for understanding radiation transport in a host of plasma environments, such as those encountered in stellar interiors, fusion devices and short-wavelength plasma light sources [30, 5255]. In this context, an opacity quantifies how photons, travelling in a plasma, are absorbed and scattered through interactions with the plasma constituents. For a given photon frequency ν, the opacity ${\kappa }_{\nu }^{\mathrm{T}}$ is defined as the contribution of four mechanisms:

Equation (1)

The first three terms represent bound–bound (photoabsorption), bound–free (photoionisation) and free–free (inverse bremsstrahlung) contributions to the opacity and each contain a factor correcting for stimulated emission. The final term represents photon scattering. In the present work, we are mainly concerned with the bound–bound contribution to the opacity (since all other contributions are minor in the wavelength range of interest for these plasma conditions) and therefore we drop the superscript 'BB' notation in the following. The bound–bound opacity κν is written

Equation (2)

where ρ is the mass density and the summation term is known as the absorptivity. Here, ${\sigma }_{\nu }^{ij}$ is the cross section for photoabsorption from level i to level j and ni(j) are the population densities (simply called 'populations' in the following) of energy levels i(j) having statistical weights gi(j). The sum runs over all transitions i→j associated with photon absorption at frequency ν. The second term in the summation is the aforementioned correction for stimulated emission. The cross section for photoabsorption ${\sigma }_{\nu }^{ij}$ is given by

Equation (3)

where e, me and c are the electron charge, electron mass and speed of light, respectively. The term fij is the absorption oscillator strength for a transition i → j and ϕ(ν) is the line profile function satisfying ${\int }_{0}^{\infty }\phi \left(\nu \right)\mathrm{d}\nu =1$.

As one can see from equations (2) and (3), evaluation of the bound–bound opacity requires knowledge of both atomic structure (ν and fij ) and the level populations ni(j). Evaluation of this latter quantity requires knowledge of the plasma environment in which the atomic species is embedded. Atomic, using as input the mass density ρ and electron temperature Te, can calculate these level populations in LTE or non-LTE plasma conditions (the radiation temperature Tr can also be specified in non-LTE calculations). The LTE-mode of Atomic uses the ChemEOS equation-of-state (EOS) model to derive level populations and charge state distributions (CSDs) [56, 57]. This model is based on the free-energy-minimisation approach and the occupation probability formalism (this latter procedure allows plasma density effects to be included in a consistent manner). Importantly, non-ideal plasma effects, such as plasma microfields and strong coupling are included in the free energy term. This leads to modified Boltzmann and Saha relations from which level populations and CSDs are derived.

Atomic also has the capability of modelling plasmas under non-LTE conditions [58, 59]. Unlike LTE plasmas, the calculation of level populations in non-LTE conditions requires a microscopic viewpoint in which the rates of all relevant atomic populating and depopulating processes are calculated. The most general definition of a non-LTE plasma is one which exhibits any characteristic departure from LTE conditions, such as the existence of a non-Maxwellian electron energy distribution function (EEDF) or a non-Planckian radiation field. As such, there are a variety of ways that a plasma can be defined as under non-LTE conditions. Consider an atomic state (energy level, configuration or superconfiguration) i associated with some charge state ζ in the plasma. In non-LTE plasmas, the population of this state evolves according to a collisional-radiative (CR) equation of the form [60, 61]

Equation (4)

Here, ${R}_{j{\zeta }^{\prime }\to i\zeta }={R}_{j{\zeta }^{\prime }\to i\zeta }^{\mathrm{C}}+{R}_{j{\zeta }^{\prime }\to i\zeta }^{\mathrm{R}}$ denotes the rate of collisional (${R}_{j{\zeta }^{\prime }\to i\zeta }^{\mathrm{C}}$) and radiative (${R}_{j{\zeta }^{\prime }\to i\zeta }^{\mathrm{R}}$) processes connecting j (associated with some charge state ζ') to i. In a similar fashion, rates describing ij are denoted R . The first term in equation (4) represents population influx into i and the second term describes population outflux from i. A CR equation of the form equation (4) exists for all atomic states i in each charge state ζ present in the plasma. Coupling these states yields the so-called rate equation

Equation (5)

Here, N is a vector of atomic state populations (n N) and A is the so-called rate matrix containing all relevant rates ${R}_{j{\zeta }^{\prime }\to i\zeta ,i\zeta \to j{\zeta }^{\prime }}^{\mathrm{C},\mathrm{R}}$.

The choice of atomic processes to be included in the rate matrix A is highly case-dependent; in coronal plasmas, for example, one can neglect the role of three-body recombination. Such simplifications are not possible in the current case and therefore we must consider a wide range of atomic processes (see section 3 for more details). Fortunately, Atomic has the capability of modelling a variety of (de-)populating processes such as electron-impact excitation/ionisation, photoexcitation, photoionisation, autoionisation (AI), etc. Cross sections for electron-impact excitation were calculated in the plane-wave Born (PWB) approximation using the CATS code. Electron-impact ionisation cross-sections were obtained from the GIPPER code (photoionisation cross sections can also be obtained if an external radiation field is included in the non-LTE model) [62]. Atomic uses these cross sections to calculate the rates which form the entries of A. Invoking a conservation equation of the form

Equation (6)

where Ni is the number of atomic states and nion is the ion density allows one to solve equation (5) for the atomic state populations n . These populations can then used to produce opacities [equation (2)] in a similar fashion as in the LTE mode of Atomic.

Finally, it should be emphasized that a consistent atomic structure has been employed in all stages of the calculation, ranging from the evaluation of collisional cross-sections to the derivation of atomic state populations and level-resolved opacities. This procedure ensures self-consistency in all aspects of the calculation.

3. Tin ionisation distributions and configuration populations in CO2 laser-driven plasma conditions

The study of CO2 laser irradiation of tin targets comprises a large body of work (see [20] and references therein). A very important aspect of this work is the study of tin-ion opacities in the relevant plasma conditions [32, 3739, 63]. Plasmas generated by CO2 laser irradiation lie in a rather unique position in (ne, Te) space; electron densities are not high enough to allow for LTE conditions to prevail, however not so low to justify the use of coronal equilibrium relations. In this respect, the atomic modelling of CO2-driven laser–plasmas should, in theory, employ the CR approach described in section 2.

As outlined in the introduction, the goal of the present work is to evaluate the role of multiply-excited states in CO2 laser-driven plasma conditions. As such, we must inherit the complicated level structures considered in reference [13]. As discussed in reference [13, 39], the use of extensive atomic models for Sn11+–Sn14+ is necessary to ensure (i) correct wavelength positions of Δn = 0 transitions (which are sensitive to the choice of underlying atomic structure as a result of CI) and (ii) partition function convergence [56]. Unfortunately, this choice of extensivity over simplicity makes the solution of equation (5) for these level structures intractable; there are simply too many atomic levels to consider. As discussed in reference [13], the calculation of detailed, level-resolved opacities for such complex structures is currently only possible using the LTE approach in Atomic.

3.1. Busquet's ionisation temperature method

In order to circumvent computational intractabilities associated with level-resolved, non-LTE opacity calculations involving multiply-excited configurations, we have employed the ionisation temperature method of Busquet [64, 65] to effectively 'mimick' non-LTE opacities using LTE computations. The key observation underlying this method is that non-LTE CSDs (performed at some Te) often resemble LTE CSDs undertaken at a temperature TZ different to Te [66]. The first step in Busquet's approach is to identify the LTE temperature TZ which yields optimum agreement in average charge state ⟨Z⟩ between the non-LTE and LTE calculations, i.e.

Equation (7)

In the current context, TZ is known as the 'ionisation temperature'. Once TZ is determined, one then performs detailed, level-resolved LTE opacity calculations at TZ . Higher-order moments of the CSD, such as the width, are not considered in this method. Although this approach provides a numerically cost-efficient means of incorporating non-LTE opacities (and EOS quantities) into hydrodynamic codes, Klapisch and Bar-Shalom have stressed that it should only be applied to plasmas not far from pure LTE conditions [66].

We wish to test the validity this method in the context of the current problem. Guided by previous modelling efforts of Sasaki and co-workers [37, 40, 41], we have undertaken a non-LTE configuration-averaged (CA) Atomic calculation at ρ = 2 × 10−4 g cm−3 and Te = 36 eV. The electron density of this plasma is ne = 1.27 × 1019 cm−3, close to the critical electron density for CO2 laser light (${n}_{\text{crit},{\mathrm{C}\mathrm{O}}_{\mathrm{2}}}=1.1{\times}1{0}^{19}\enspace {\mathrm{c}\mathrm{m}}^{-3}$). The CA approach generates appropriately averaged energies, oscillator strengths and PWB electron-impact excitation cross sections from which configuration-to-configuration rates are derived [59]. This approach is employed in the present context to ensure computational tractability. Configurations sets employed in the CA CATS calculations for Sn12+ and Sn14+ are provided in table 1. The calculations were performed under steady-state conditions, i.e. dN/dt = 0. This approximation is valid for plasmas where the time taken to generate a given CSD (through ionisation and recombination processes) occurs on timescales shorter than the laser pulse duration. This is certainly true for industrial LPP EUV sources, which are typically driven by multi-10 nanosecond-long CO2 pulses. We calculated the rates of electron-impact excitation/ionisation and AI (as well as their inverse processes) assuming a Maxwell–Boltzmann EEDF at Te = 36 eV. The calculations were performed without an external radiation field, i.e. photoexcitation and photoionisation processes were not considered (see subsection 3.2 below for further discussion). The process of spontaneous emission was included in the model.

Table 1. Configuration sets adopted for Sn12+ and Sn14+ in the non-LTE/LTE CA calculations (see main text). Configurations are grouped according to excitation degree. Here, angular momenta l = 0 – 4 (sg), l' = 1–4 (pg) and $\tilde {l}=2\enspace --\enspace 4\enspace \left(d-g\right)$, respectively.

Sn12+ Sn14+
Ground configuration Ground configuration
4s24p64d2 4s24p6
Singly-excited configurations Singly-excited configurations
4s24p64d + {4f, 5l}4s24p5 + {4d, 4f, 5l}
4s24p5 + {4d3, 4d24f, 4d25l}4s14p6 + {4d, 4f, 5l}
4s4p6 + {4d3, 4d24f, 4d25l} 
Doubly-excited configurations Doubly-excited configurations
4s24p6 + $\left\{4{f}^{2},4f5l,5s5l,5p5{l}^{\prime },5d5\tilde {l},5f5g\right\}$ 4s24p4 + {4d2, 4d4f, 4f2, 4d5l, 4f5l, 5s5l, 5p5l', 5d5f, 5d5g, 5f5g}
4s24p5 + {4d4f2, 4d5s2, 4d4f5l, 4d5s5p, 4d5s5d}4s14p5 + {4d2, 4d4f, 4f2, 4d5l, 4f5l, 5s5l, 5p5l', 5d5f, 5d5g, 5f5g}
4s24p4 + {4d4, 4d34f, 4d35l, 4d24f2}4p6 + {4d2, 4d4f, 4d5l, 4f5l}
4s4p54d4  
Triply-excited configuration Triply-excited configurations
4s24p34d5 4s24p34d3, 4s24p34d24f, 4s4p44d3, 4s4p44d24f

The CSD resulting from this calculation is shown in black in figure 1. Importantly, the relevant EUV-emitting Sn11+–Sn14+ ions are produced under such conditions, with Sn12+ and Sn13+ clearly the dominant species. The average charge state of this plasma is ⟨Znon-LTE = 12.6. Also illustrated in figure 1 is an LTE CA calculation (shown in red) performed at ρ = 2 × 10−4 g cm−3 and Te = 26 eV. These calculations yield ⟨ZLTE = 12.6, identical to the non-LTE calculation. Also, we note that the CSDs are very nearly identical. An ionisation temperature of TZ = 26 eV is thus established. We note that the ratio Te/TZ = 36/26 ≈ 1.4 is larger than that obtained by the scaling law ${T}_{\mathrm{e}}/{T}_{Z}\approx {\left(1+\left(5.36{\times}1{0}^{13}\right){T}_{\mathrm{e}}^{7/2}{n}_{\mathrm{e}}^{-1}\right)}^{0.215}\approx 1.2$ obtained by Busquet et al [65]. Much better agreement with this scaling law is obtained in the Nd:YAG-relevant case considered in [13] (ne = 1.27 × 1020 cm−3, TZ = 32 eV, Te = 34.5 eV), where Te/TZ = 34.5/32 ≈ 1.08 and the scaling law yields Te/TZ ≈ 1.02.

Figure 1.

Figure 1. Charge-state distributions for plasmas calculated in non-LTE (Te = 36 eV; shown in black) and LTE (TZ = 26 eV; shown in red) conditions. The plasma mass density used in both calculations was ρ = 2 × 10−4 g cm−3.

Standard image High-resolution image

It is interesting to consider why optimum agreement in CSDs is obtained for Te > TZ . Exact reproduction of an LTE CSD is possible using a Te = 26 eV non-LTE calculation, however one must include an external radiation field in the calculation characterised by Tr = Te. In this way, all atomic processes are in detailed balance. However, as detailed above, our non-LTE calculations do not include an external radiation field. In order to match our present calculations with a TZ = 26 eV LTE calculation, we must therefore compensate for the missing (de)excitation and ionisation channels provided by photoexcitation and photoionisation by increasing Te above TZ . The difference TeTZ 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 TeTZ = 2.5 eV (where Te = 34.5 eV and TZ = 32 eV) highlights the dominance of collisional rates over radiative rates at near-LTE conditions (ne = 1.27 × 1020 cm−3). The same cannot be said for the CO2-relevant conditions, where a value TeTZ = 10 eV is indicative of truly non-LTE conditions, i.e. collisional rates do not significantly outweigh radiative rates.

3.2. Population distributions and effective temperatures

The replication of a non-LTE CSD using LTE computations, such as that shown in figure 1, is key for justifying the use of LTE opacities as an approximation for non-LTE opacities. An even stronger argument in favour of this approximation is to establish equivalence between the non-LTE and LTE-computed configuration populations (the cumulative sum of which yields the CSD—see equation (6)). To this end, we present in figure 2 non-LTE- and LTE-computed configuration populations (Pnon-LTE, PLTE) as a function of CA energy (Eav) for Sn12+ (upper panel) and Sn14+ (lower panel). The non-LTE configuration population calculations are shown in figure 2(a) (Sn12+) and figure 2(d) (Sn14+). The corresponding LTE calculations are shown in figures 2(b) and (e). Configurations associated with single-, double- and triple-electron excitations are shown as blue, orange and green circles, respectively. The ground-state configurations, fixed at Eav = 0 eV, are shown as grey circles. Perhaps the most striking feature of these calculations is the near-coincidence in non-LTE- and LTE-computed configuration populations. This is quantified in figures 2(c) and (f), where, for a given configuration C, we plot the non-LTE-normalised Configuration Population Difference (CPD) defined as (PC,non−LTEPC,LTE)/PC,non−LTE. For the majority of configurations, this quantity takes on values below 1%. This indicates a near-coincidence in non-LTE and LTE-computed configuration populations. The few configurations where the non-LTE-normalised CPD values are above 1% all have LTE-computed populations greater than the corresponding non-LTE populations.

Figure 2.

Figure 2. Configuration populations for Sn12+ (top) and Sn14+ (bottom) calculated for a Te = 36 eV non-LTE plasma [panels (a) and (d)] and a Te = 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+.

Standard image High-resolution image

In addition to validating the use of Busquet's approach in the current context, the equivalence in configuration populations reveals an interesting characteristic of the non-LTE-computed populations in that they appear to follow a single Boltzmann-esque temperature law of the form

Equation (8)

where ${g}_{C}={\sum }_{{J}_{C}}\left(2{J}_{C}+1\right)$ is the total statistical weight of configuration C (JC is an energy level associated with C) and Teff is a so-called effective temperature [67] (the Boltzmann constant kB typically appearing in equations of the form equation (8) is here 'absorbed' into the definition of Teff).

In order to evaluate Teff, we plot in figure 3 the natural logarithm of the ratio PC,non−LTE/gC as a function of Eav for (a) Sn12+ and (b) Sn14+. First-order polynomial fits to the data are shown in red. Excellent fits to the data are characterised by R2 values of 0.99 attained in both cases. The effective temperatures Teff are obtained from the inverse-negative of the slopes of the lines of best fit. As indicated in the plots, the fits yield values of Teff = 26.6 (Sn12+) and 26.9 eV (Sn14+), clearly very similar to the LTE ionisation temperature TZ = 26 eV. This close agreement between Teff and TZ is perhaps not too surprizing given the near-coincidence in non-LTE and LTE-computed configuration populations.

Figure 3.

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

Standard image High-resolution image

This observation, that non-LTE-computed configuration populations follow Boltzmann-like temperature laws has been studied extensively in the works of Bauche, Bauche-Arnoult and others [6772]. The emergence of temperature laws is attributed to the complex interplay of atomic processes defining the steady-state CR model. More specifically, the combination of (i) the microreversibility nature of collisional processes (which drives the population distributions towards LTE conditions) and (ii) correlation laws associated with spontaneous emission (high/low-lying upper states typically decay towards high/low-lying lower states) tend to generate Boltzmann-type population distributions [71]. This is in fact a general phenomenon of non-LTE hot plasmas [68]. Finally, it should be noted that the approach of Bauche et al [70], does not ascribe a single effective temperature Teff to a population distribution. Rather, configurations are grouped according to superconfigurations and equation (8) is used to derive superconfiguration-specific temperatures. These temperatures may vary significantly from superconfiguration-to-superconfiguration (see figure 1 of [70]). Although this approach differs to that presented in figure 3 (where we have performed a 'global' fit to all configurations), the non-LTE-computed population distributions are clearly well-described by a single-temperature Boltzmann-type law.

Finally, we wish to comment on the importance of a non-zero radiation field in our non-LTE computations. As outlined in section 3, all non-LTE calculations presented thus far have not included an external radiation field in the underlying atomic kinetics. Indeed, identifying an external radiation field appropriate for the current study is rather difficult as it is a purely non-local quantity, i.e. it depends on the radiative characteristics of the plasma surrounding the studied (ρ, Te) plasma. Proper evaluation of an external radiation field requires complex simulations involving the coupling of non-LTE atomic kinetics with a model for radiation transport in a multi-(ρ, Te, Tr) plasma. Such an endeavour is outside the scope of the present work. Instead, we have performed an additional non-LTE calculation where we have included a Planckian radiation field (characterised by a radiation field Tr). We fixed Tr and tuned Te until equation (7) is established with a TZ = 26 eV LTE plasma. We find that a Te = 33 eV and Tr = 21 eV plasma, for example, reproduces both equation (7) and the LTE CSD. Moreover, the configuration populations associated with Sn12+ and Sn14+ are also found to be in excellent agreement with the LTE-computed configuration populations. This is demonstrated in figure 4, where, in a similar vein to figure 3, we plot the natural logarithm of the ratio PC,non−LTE/gC as a function of Eav for (a) Sn12+ and (b) Sn14+. Effective temperatures obtained from the slopes of the lines of best fit (shown in red) are Teff = 26.6 eV for both Sn12+ and Sn14+, clearly in excellent agreement with the data presented in figure 3. The introduction of a non-zero radiation field in these computations therefore does not break the observed Boltzmann-esque temperature law once equivalence between the non-LTE and LTE-computed CSDs is established.

Figure 4.

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

Standard image High-resolution image

4. Tin-ion opacities

We now present the results of detailed, level-resolved LTE opacity calculations at CO2 laser-driven plasma conditions. The LTE plasma conditions considered here are the same as those of section 3 (ρ = 2 × 10−4 g cm−3, TZ = 26 eV). Unlike section 3, we adopt a fully level-resolved approach where Atomic is used to calculate level populations and bound–bound opacities. The atomic structures employed in these calculations are the same as those considered in reference [13] (see section 2).

The results of these calculations are presented in figure 5. Panels (a)–(d) illustrate the contributions to the opacity from the four charge states Sn11+–Sn14+. In a similar fashion to Torretti et al [13], the different contributions to the opacity are plotted in a cumulative fashion; photoabsorption from levels in the ground manifold to singly-excited levels are shown in blue (we call this the 'ground opacity'); the summation of photoabsorption from singly-excited levels to doubly-excited levels ('first opacity') and the ground opacity is shown by the line bounding the orange shaded area; the summation of photoabsorption from doubly- and triply-excited levels ('higher opacity'), the ground opacity and first opacity is shown by the line bounding the green shaded area. The grey region in each panel represents a 2% bw centred at 13.5 nm.

Figure 5.

Figure 5. Panels (a)–(d): opacities of the ions Sn11+–Sn14+. The 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 ground opacity is shown by the line bounding the orange shaded region. The line bounding the green shaded region is the summation of photoabsorption from doubly- and triply-excited levels ('higher opacity'), the ground opacity and first opacity. The total opacity, obtained by summing the contributions of Sn11+–Sn14+, is shown in panel (e). The inset in panel (e) shows the opacity in a 2% bw centred at 13.5 nm.

Standard image High-resolution image

First, examining the ordinate of figures 5(b) and (c), one readily concludes that Sn12+ and Sn13+ are the dominant contributors to the opacity in this plasma. This can be understood from the CSD presented in figure 1, where these two charge states accumulate nearly 80% of the total ion population. Although the majority of the opacity in the Sn12+ ion can be attributed to the ground opacity, the first and higher opacities contribute to the spectrum at wavelengths larger than 13.5 nm. The same is true for Sn11+. One of the most striking features of these opacity spectra is the spectral dominance of a transition in Sn13+ [see figure 5 (c)]. This transition is the subvalence excitation 4p64d ${}^{2}D_{5/2}$−4p54d2 ${}^{2}F_{7/2}$. It has a calculated wavelength of 13.286 nm, in good agreement with experimental identifications of this transition at 13.3014 [17], 13.3020 [73] and 13.301 nm [29]. The opacity of this transition is κ ≈ 83 × 105 cm2 g−1, nearly a factor 3 higher than any other transition in Sn11+–Sn14+. There are two reasons for this spectral dominance. First, as illustrated in figure 2, a significant fraction of Sn12+ and Sn14+ ions are in their ground-state configurations. This also true for Sn13+. Secondly, this LS-allowed transition exhibits a large oscillator strength. This combination, which appears multiplicatively in the first term of equation (2), is the origin of this large opacity value. As in the case of Sn11+ and Sn12+, higher-order opacities are seen to contribute at wavelengths above 13.5 nm, however only on the order κ ≈ 10 × 105 cm2 g−1. Moving to the closed 4p subshell Sn14+ ion we observe an interesting shift in opacity dominance away from the ground opacity to the first opacity. The ground opacity in this wavelength region is dominated by the transition 4p6 ${}^{1}S_{0}$–4p54d ${}^{1}P_{1}$, which has a calculated wavelength of 13.332 nm. This is in excellent agreement with experimental identifications of this transition at λ = 13.344 nm [29] and λ = 13.343 [18, 74]. This transition is clearly much weaker than the $4{p}^{6}4d\enspace {}^{2}D_{5/2}-4{p}^{5}4{d}^{2}{\enspace }^{2}{F}_{7/2}$ transition in Sn13+. This is because (i) the 4p6 1 S0 level in Sn14+ has a smaller population than the $4{p}^{6}4d\enspace {}^{2}D_{5/2}$ level in Sn13+ and (ii) the transition in Sn14+ involves energy levels having low angular momenta (Jlower, upper = 0, 1) and therefore has a lower oscillator strength than the aforementioned transition in Sn13+. As is clear from figure 5(d), photoabsorption from singly-excited levels is the main source of opacity in Sn14+. Importantly, all of the opacity in the industrially-relevant 2% bw (the grey shaded region) in Sn14+ originates from higher-order opacities, the first opacity being the dominant contributor.

The total opacity, obtained by summing the contributions in figures 5(a)–(d), is shown in figure 5(e). As expected, most of the higher-order opacity contributions are located at wavelengths above 13.5 nm. The inset illustrates the contributions to the opacity in the 2% bw. Higher-order opacities clearly make a significant contribution to this industrially-relevant spectral region. In essence, they tend to 'fill out' regions between the ground opacities. The charge-state-specific opacity contributions in the 2% bw are illustrated in figures 6(a)–(d), with their combined opacity contribution illustrated in figure 6(e). Importantly, only 34% of the opacity arises from the ground opacity. Of the remaining 66%, approximately 42% originates from the first opacity and 24% is associated with the higher opacity. This result clearly highlights the important spectral role played by multiply-excited transitions in CO2 laser-driven plasma conditions.

Figure 6.

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 (a)–(d), is shown in panel (e).

Standard image High-resolution image

Finally, we wish to compare these results with the Nd:YAG-relevant case considered in [13]. The plasma conditions considered in reference [13] were ρ = 2 × 10−3 g cm−3 and Te = 32 eV. This LTE calculation yielded ne = 1.27 × 1020 cm−3 (an order of magnitude higher than the current case) and an average charge state ⟨Z⟩ similar to the present case. Comparing figure 5(e) with figure 3(e) of [13], we see that, relative to the ground opacity, the higher-order opacities in the Nd:YAG case make a significantly larger contribution than in the present case. Comparing Maxwell–Boltzmann distributions at Te = 32/26 eV, one observes that the former has a larger fraction of higher-energy electrons which can collisionally populate high-lying levels associated with doubly-, triply- and quadruply-excited configurations. As both calculations were performed in LTE, the populations of the levels are described by the so-called 'partition function factor', which for some level J is written gJ  exp(−EJ /Te). To enable a comparison between both cases, we plot in figure 7 the ratio of the partition function factor for Te = 32 and 26 eV (shown in black) using a continuous energy variable E. The statistical weight gJ naturally drops out of this ratio. It is clear from this figure that the population of multiply-excited (doubly-, triply- and quadruply-excited) states is much higher in the Te = 32 eV plasma than in the Te = 26 eV plasma. Examples of this ratio for the 4p54d3, 4p44d4 and 4p34d5 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 equivalent to that generated under relevant CO2-laser-driven tin plasma conditions.

Figure 7.

Figure 7. Ratio of the partition function factor for Te = 32 and 26 eV plasmas (shown in black). The ratio for the 4p54d3, 4p44d4 and 4p34d5 configurations in Sn12+ are shown in blue, orange and green, respectively.

Standard image High-resolution image

5. Conclusion

In conclusion, we have investigated the role played by multiply-excited states in industrially-relevant CO2 laser-driven tin plasma conditions. First, we employed Busquet's ionisation temperature method to match the charge state distribution of a non-LTE plasma with an LTE computation. We find excellent agreement between the non-LTE-computed configuration populations and LTE calculations for both Sn12+ and Sn14+. Moreover, we identify the existence of Boltzmann-like temperature laws in the non-LTE population distributions. In the second part of this study we performed detailed, level-resolved LTE opacity calculations at plasma conditions pertaining to CO2 laser-driven plasma conditions. Approximately 66% of the opacity in the 2% bw centred at 13.5 nm was found to arise from transitions between multiply-excited states. This result highlights the important role played by multiply-excited states in industrially-relevant plasma conditions. This work paves the way for the generation of detailed LTE opacity tables at numerous (ρ, Te, Tr) non-LTE plasma conditions for incorporation in radiation-hydrodynamic simulations. Such simulations will enable reliable predictions of EUV emission from LPPs and are therefore key for guiding experimental efforts in characterising and optimising laser-driven plasma sources of EUV light.

Acknowledgments

We would like to thank W van der Zande, W Ubachs and R Hoekstra for their useful comments on the paper. This project has received funding from European Research Council (ERC) Starting Grant No. 802648 and is part of the VIDI research programme with Project No. 15697, which is financed by the Netherlands Organization for Scientific Research (NWO). Part of this work has been carried out at the Advanced Research Center for Nanolithography (ARCNL), a public-private partnership of the University of Amsterdam (UvA), the Vrije Universiteit Amsterdam (VU), NWO and the semiconductor equipment manufacturer ASML. Part of this work was supported by the ASC PEM programme of the US Department of Energy through the Los Alamos National Laboratory. Los Alamos National Laboratory is operated by Triad National Security, LLC, for the National Nuclear Security Administration of US Department of Energy (Contract No. 89233218NCA000001).

Please wait… references are loading.