This work presents the first experimental study on the near-threshold sputtering regime for monocrystalline low-index plane tungsten targets investigated using high-resolution emission spectroscopy. We analyzed the line shape emitted by sputtered atoms, which contains information on the angular and velocity distribution functions via Doppler broadening. Specifically, we report changes in the line profile of the resonant W I 498.4 nm transition during plasma exposure of polycrystalline and monocrystalline (100) and (111) tungsten targets at the linear plasma device PSI-2. Biasing the targets from −60 V to −100 V provided low-energy argon ions for near-threshold sputtering. The line shapes, measured along the angle of observation perpendicular to the normal of the sample, were significantly broader for the monocrystalline (100) and (111) compared to that of the polycrystalline target. In particular, the (111) target demonstrates a pronounced heart-shaped profile. The modeling captures this distribution via a cos(θ)exp(−bθ) function—θ is the polar angle—combined with a parameterized Thompson velocity distribution. Furthermore, comparing the experimental data to molecular dynamics simulations at 100 eV illustrates a reasonable agreement of the angular distribution function with the measurements.

Physical sputtering is one of the most fundamental processes in plasma surface interaction. Ions that bombard a surface may knock target atoms out of their equilibrium positions within the solid. Further interaction of the projectile and recoil atom with the crystal can create a significantly larger collisional cascade. Additionally, assuming an amorphous structure and not too high kinetic energies creates an isotropic flux within the solid. As this flux moves through the surface, it naturally has a cosine angular distribution function (ADF) and has to overcome the surface binding energy. This simple physical picture describes qualitatively why physical sputtering often creates the well-known Knudsen cosine law.1 

The experimental proof of such collisional cascades being the mechanism underlying physical sputtering was provided by Wehner.2 He exposed low-index crystalline planes of silver to low-energy Hg ions at normal incidence and collected the deposition. The measured spot-like deposition patterns depended upon the orientation of the monocrystal surface. Furthermore, the dependence on the polar angle was non-cosine-like and non-isotropic in the azimuthal direction. These results illustrate how the collisional cascade in the near-threshold regime of physical sputtering depends on the surface orientation.

This fundamental information on the angular distribution function of sputtered atoms is vital for modeling the plasma–wall interaction in thermonuclear fusion reactors.3 For example, the currently being built International Thermonuclear Experimental Reactor (ITER) will have a tungsten wall, and its erosion via plasma particle bombardment creates impurities that can drastically change the physics of the discharge.4 One issue is that the high-Z element tungsten will have bound electrons even within the high-temperature region of the plasma and can rapidly cool it down.5 Fusing a deuteron and triton at sufficient rates for energy production depends strongly on the temperature. Thus, cooling the plasma via high-Z elements6 will be a lynchpin of fusion power plants, and correctly predicting the erosion and trajectory of sputtered tungsten is of the essence.

The sources of the physical erosion in a fusion reactor are the fast charge exchange (CX) neutrals for the main wall7 and ions in plasma-wetted areas, such as the divertor.3 Erosion due to the CX neutrals is well understood due to their kinetic energy on the order of keV since this produces the linear cascade regime for the physical sputtering, which the binary collision approximation captures accurately.8,9 The plasma ions, however, have significantly less kinetic energy, so their collisional processes usually occur within the first few atomic layers of the plasma-facing component. The assumption of using binary collisions can also break,10 which is why the computationally more expensive molecular dynamics (MD) simulations may have to be used to capture the erosion correctly.11 

Another critical aspect of the plasma-facing components is their retention of hydrogen isotopes.12 For example, regulatory standards strictly limit the amount of the radioactive isotope tritium stored within the wall. This issue is one of the reasons why in situ laser-based methods for monitoring retention are a highly active area of research.13–15 Complementary to fusion reactors, specifically designed material interaction facilities enable testing the properties of the plasma-facing components within a laboratory setting.16 For instance, linear plasma devices provide fluxes of the ions onto samples similar to those expected in the divertor region of fusion reactors.

Not only isotopes of hydrogen but also seeding gas impurities bombard the divertor of fusion reactors.17 These elements, such as argon, dominate the erosion of the tungsten divertor during the so-called inter edge localized mode phases.3 Experimental data and molecular dynamics simulation suggest in the case of near-threshold sputtering of polycrystalline tungsten by Ar+ a butterfly-like ADF.18,19 However, modeling by Stepanova and Dew20 suggests an almost pure cosine angular distribution function at projectile energies as low as 100 eV. The current literature lacks experimental data on the angular distribution of particles sputtered via low-energy ions from monocrystalline tungsten targets.

Linear plasma devices, like PSI-2, are well suited to address this problem in the near-threshold energy regime. The diagnostic of choice is the high-resolution spectroscopy of the emission by the sputtered atoms themselves. Indeed, the line shape of sputtered atoms provides through the Doppler broadening information on the ADF and velocity distribution function (VDF). Essentially, the discharge fulfills two purposes: First, it creates the flux of ions onto a to-be-investigated sample, and second, it excites the sputtered atoms into energetically higher electronic states. Their radiative relaxation results in spontaneous emission of photons, which the spectrometer measures.

The pilot study at PSI-2 was limited to the one line of sight (LOS) that passes through the plasma source.21 This LOS was anti-parallel to the normal of the tungsten target (0° LOS). Adding the LOS in the direction perpendicular to the symmetry axis (90° LOS) proved that modeling both line shapes with consistent ADF and VDF is impossible within the standard model. This standard model explicitly assumes an infinite size of the target. The model proposed and benchmarked in Ref. 22 presents a method to include the geometric losses and excitation/ionization processes. The fitting of the ADF via cosine power-law scaling suggested a pronounced under-cosine distribution during near-threshold sputtering of polycrystalline tungsten by Ar+ bombardment.

Furthermore, improving the modeling requires accounting for the correct Zeeman hyperfine splitting of the odd isotope 183W, as benchmarked in Ref. 23 using Kr I and Xe I transitions. The results presented in this study continue the development of the line shape emission spectroscopy of sputtered atoms. The main goal is to validate the approach of monitoring the ADF during plasma exposure since this could provide additional information on the changes in surface structure. So, for instance, the method may allow for operando monitoring of the distribution functions of atoms sputtered from strongly reformed surfaces, such as during the growth/erosion of tungsten fuzz.24 

Before attempting such time-resolved measurements, where the interpretation of the data may be difficult, it is essential to continue benchmarking the approach under well-definable conditions. This goal is the rationale for exposing low-index crystallographic planes of tungsten—(100) (111)—and a polycrystalline sample to the plasma at PSI-2 since one expects a pronounced influence on the ADF for near-threshold sputtering.19 At the same time, this is the first experimental study investigating the low-energy regime (<100 eV) for sputtering monocrystalline tungsten targets via Ar+ projectiles.

Observing significant changes in the line shape that depend on the surface structure of the target and align with the expectation from modeling the line shape then validates the use of line shape emission spectroscopy for monitoring changes in the ADF during plasma exposure. The data presented in this study are for biases of −60 V, −80 V, and −100 V applied to the target and represents the first measurements of such line shapes by atoms sputtered from monocrystalline samples.

A steady-state arc discharge driven between a heated LaB6 cathode and a grounded Mo anode generates a plasma that an axial magnetic field confines, as Fig. 1 illustrates. This plasma exhibits an axisymmetric hollow electron density and temperature profile due to the geometry of the source, and the plasma propagates into an exposure area with a homogeneous magnetic field to study plasma–wall interaction. A radially movable Langmuir probe monitors the plasma parameters, and the overview paper of this linear plasma device (PSI-2) provides additional information on its characteristics.25 

FIG. 1.

Schematic of the linear plasma device PSI-2, showing the two LOS used for measuring the emission by sputtered tungsten. Tungsten and nickel hollow cathode lamps provide the calibration spectra. The plasma parameters are almost identical to the previous study (Ref. 22) on sputtered W at PSI-2. The peak electron density in this study was 1.8×1018m3, and the electron temperature was below 4 eV.

FIG. 1.

Schematic of the linear plasma device PSI-2, showing the two LOS used for measuring the emission by sputtered tungsten. Tungsten and nickel hollow cathode lamps provide the calibration spectra. The plasma parameters are almost identical to the previous study (Ref. 22) on sputtered W at PSI-2. The peak electron density in this study was 1.8×1018m3, and the electron temperature was below 4 eV.

Close modal

One distinct advantage of this linear plasma device for spectroscopy compared to other linear machines26–28 is the geometry of the source: The hollow and ring-shaped electrodes enable measuring optical emission spectra along the line of sight (LOS) that moves through the plasma source. In this study, we refer to this LOS as 0°. The reasoning is simply that the symmetry axis of our experiment is in this direction. Naturally, the LOS perpendicular to this direction is 90°. Figure 1 indicates these lines of sight via the two green beams that exit the vacuum vessel, and their diameters are about 2 mm to 3 mm. The LOS along 0° terminates at the surface of the target exposed to the plasma, whereas the LOS at 90° collects light from the region having an axial distance of 2 mm from the target.

The topology of the magnetic field within the exposure area created by the six Cu solenoids ensures the bombardment of the samples at normal incidence.16 In this region, the field strength of 90 mT provides the normal Zeeman splitting between a π (Δm=0) to σ (Δm=±1) component on the order of 1 GHz. A polarizing beam splitter (CCM1-PBS251, Thorlabs, 420–680 nm) mounted on the optical bench at the 90° LOS isolates the π components (cf. Fig. 13 of Ref. 29). Optical isolation due to a quarter waveplate (λ/4, B. Halle, 460–680 nm) and linear polarizer (Glan–Thompson Prism, B. Halle) distinguishes between the σ+ and σ transitions at 0°.30 Both lines of sight image simultaneously onto the slit of a high-resolution Echelle spectrometer. A bandpass filter with a width of 10 nm centered around 500 nm suppresses the superposition of different spectral orders. In this study, the 1024 horizontal pixels of the charge-coupled device (CCD) camera cover a wavelength range of 0.7 nm (λ/Δλ7·105).

Two hollow cathode lamps (Ni, W) provide the calibration data needed for unambiguously deriving the wavelength dispersion of the spectrometer. The line targeted in this study is the resonant 498.4 nm transition of neutral tungsten.

Converting the Thompson distribution function of the particle flux density Φ(E,θ) [Eq. (19) of Ref. 31] into the velocity distribution function F by separating the cosine angular distribution, transforming from energy into velocity space (dE/dv=mv), and using F=Φ/v gives
(1)
where v is the velocity of the sputtered particle, M is its atomic mass, Eb denotes the surface binding energy, and vmax is the maximal velocity of the sputtered atom. The parameter s = 3 corresponds to the standard Thompson value. The maximal velocity in the binary collision approximation (BCA) after leaving the solid is
(2)
where Eimp is the impact energy of the projectile and Mpr denotes its mass. Equation (2) shows that, according to the BCA, sputtering occurs when the term within the round brackets exceeds zero. This consideration gives a threshold value for the Ar+ bombarding tungsten of 14.8 eV.32 In reality, sputtering also requires momentum reversal, further increasing the threshold.

As stated in the introduction, this study aims to verify that the line shape by transitions emitted by sputtered atoms does not only provide information on the energy30 but also about the angular distribution function. Therefore, monocrystalline (100) (111), and a polycrystalline tungsten target are exposed to Ar ion bombardment under the same plasma conditions to create sputtered atoms. In particular, in the case of monocrystalline targets, one expects a very non-cosine-like angular distribution function.19 Several approximations can describe this effect.

For instance, adding two-cosine power-law scaling terms (acosb(θ)+ccosd(θ)) and optimizing their parameters (a,b,c,d) can capture such distributions.33,34 However, applying a fitting procedure for the measured line shape using a four-parameter ADF may run into a regime with a non-unique solution as the instrumental profile is still considerable, see Sec. III C. This issue is the rationale for optimizing a one-parameter angular distribution function.

Figure 2 shows several one-parameter angular distribution functions for the probability p=G(θ,b)dΩ of detecting atoms in the solid angle dΩ. The frequently used cosine power-law scaling cosb(θ) in the description of spectra of reflected35 as well as sputtered atoms cannot describe the results of simulations19 at large off-normal angles, as illustrated in Fig. 2(a). This problem arises because a negative b creates a divergence at 90°, giving the isotropic distribution (b = 0) as the limiting case. Figure 2(b) shows an alternative ADF proposed by Yamamura et al.36 The limitation of this distribution function is at the 0° off-normal angle, as negative probabilities are unphysical, providing b = −1 as the most extreme case for large emission angles. The distribution function chosen for this study is a cosine distribution modified by an exponential term
(3)
FIG. 2.

Images showing the several one-parameter angular distribution functions considered for fitting the experimental spectra. The black-colored horizontal bar corresponds to the parameter of b, where the distribution function is the standard cosine. (c) illustrates the function chosen for the fitting of the experimental spectra. (a) Cosine power-law scaling (b) Formula by Yamamura et al.36 (c) Exponentially modified cosine distribution.

FIG. 2.

Images showing the several one-parameter angular distribution functions considered for fitting the experimental spectra. The black-colored horizontal bar corresponds to the parameter of b, where the distribution function is the standard cosine. (c) illustrates the function chosen for the fitting of the experimental spectra. (a) Cosine power-law scaling (b) Formula by Yamamura et al.36 (c) Exponentially modified cosine distribution.

Close modal

Figure 2(c) clearly shows that this function allows for describing the emission at high average off-normal angles.

The distribution functions of Eqs. (1) and (3) feed into the recently developed space-resolved emission model for sputtered atoms.22 The main idea behind the model is to transform the distribution functions into a more natural coordinate system of the problem and then integrate the contributions by the target to the whole line of sight. The first step is to project the velocity into the LOS direction, but the remaining details depend on the particular application. In this study, one transforms dΩ into dxdy as the cartesian coordinates are ideal for integrating across the surface of quadratic targets. The critical advantage of this model is that it includes apriori the geometric losses of particles out of the LOS that contribute significantly to the line shape of sputtered atoms at the linear plasma device PSI-2.22 

Here, we choose the step-profile approximation of the plasma—SRM2 of Ref. 22—due to the hollow electron density and temperature profile within PSI-2. This assumption corresponds to constant excitation and ionization rates for the 0° line of sight. In the case of the orthogonal angle of observation (90°), excitation and ionization rates are constant in front of the target and zero next to it. Section 4.1 of Ref. 22 provides a detailed description of the step-profile approximation and its validity.

Another aspect to consider is the potential confinement of the tungsten ions within the plasma column due to the magnetic field. For sputtered tungsten, one expects a most probable velocity of about 2100 ms−1 (Ref. 37) giving a Larmor radius for a 90 mT magnetic field of 40 mm. Therefore, comparing this to the size of the target (13 mm × 13 mm), the FWHM of the plasma (60 mm), and the spot size of our spectroscopy system (2 mm to 3 mm) shows that we have unconfined tungsten ions. In this case, we can neglect the recombination in the line shape analysis.

Additionally to the aforementioned broadening mechanisms, the spectral shape of the W I 498.4 nm transition depends on line splitting, as discussed below.

Naturally occurring tungsten consists of five stable isotopes, with abundancies as tabulated in Refs. 38 and 39. Their difference in isotopic weight and nuclear charge distribution shifts the center of gravity energies of the atomic levels.40  Table I presents these shifts for the 498.4 nm line.

TABLE I.

Center of gravity energy shifts of the resonant W I 498.4 nm transition compared to the line position of 186W in MHz.41 

184–186 183–186 182–186 180–186
1550 ± 6  2404 ± 9  3343 ± 6  4596 ± 12 
184–186 183–186 182–186 180–186
1550 ± 6  2404 ± 9  3343 ± 6  4596 ± 12 
The odd isotope 183W, with an abundance of 14.3%, has a nonzero nuclear spin I of 1/2. The magnetic moment associated with the spin interacts with the electron charge distribution around the nucleus. This interaction is naturally zero for the 5D0 ground state of tungsten due to its spherically symmetric electron cloud. However, the upper level (7F1°) of the 498.4 nm line has nonzero interaction, resulting in the magnetic dipole hyperfine interaction constant of (Ref. 41)
(4)
Higher-order terms, such as the electric quadrupole term, vanish since I = 1/2.
Accounting for an additional external magnetic field provides the Hamiltonian that perturbs the fine structure wave functions as (Ref. 43)
(5)
where gJ is the Landé splitting factor, μB is the Bohr magneton, B denotes the magnetic field strength, and gI is the g-factor of the nuclear magnetic dipole moment μN.44 The Hamiltonian (5) has simple solutions for the cases when either the magnetic field or hyperfine terms dominate the total perturbation. These are known as the weak-field limit and the Paschen–Back regime.

The situation regarding the 498.4 nm line of neutral 183W in a magnetic field of 90 mT is very similar to that of the Xe I lines studied with laser absorption at PSI-2. That is, Zeeman and hyperfine terms are of comparable size. We use the same code for addressing the hyperfine and Zeeman mixing, as reported in Ref. 23. The main idea is to evaluate the Hamiltonian within the basis of |I,mI,J,mJ> wavefunctions because those remain good quantum numbers in both the weak and strong field limits. The wavefunction mixing coefficients obtained for the field of 90 mT enable the calculation of relative intensities via Eq. (A.6) of Ref. 22. Solving the Hamiltonian (5) for even isotopes of tungsten is trivial because the nuclear spin I is zero, and applying the Wigner–Eckart theorem gives relative intensities as per Eq. (17.24) of Ref. 47.

Figure 3 shows the Zeeman pattern for a field of 90 mT of the Δm=0 (π) and Δm=±1 (σ±) transitions of the W I 498.4 nm line. The wavelength shift is defined (cf. Table I) relative to the heaviest naturally occurring tungsten isotope—186W. The natural abundance of the latter is 28.4%. It has only one π unshifted component and two σ components at ±1.8 pm for the 498.4 nm line (J = 1 — J = 0 transition) within the magnetic field of 90 mT. The other highly abundant 184W (30.6%) and 182W (26.5%) isotopes create the remaining two dominant lines in the π spectrum. The isotope 180W has a tiny abundance of 0.12%, responsible for an imperceptible component at about −3.8 pm. There is also a weak component at −3.4 pm due to the 183W isotope (14.3% abundance). This isotope creates four π components, two of which are dominant at the central position of the 183W 498.4 nm line. Thus, the intensity distribution resembles an almost pure Paschen–Back spectrum of the hyperfine structure—complete decoupling of I and J.

FIG. 3.

This Zeeman pattern was calculated in the same procedure as for Xe,23 assuming an external magnetic field of 90 mT, and using the Landé splitting factor 1.5442 (7F1°). The transition for even isotopes is between states with J = 0 and J = 1. Hence, the simple structure for these isotopes. 183W, highlighted with the * texture, has a more demanding pattern because of the hyperfine-Zeeman mixing.

FIG. 3.

This Zeeman pattern was calculated in the same procedure as for Xe,23 assuming an external magnetic field of 90 mT, and using the Landé splitting factor 1.5442 (7F1°). The transition for even isotopes is between states with J = 0 and J = 1. Hence, the simple structure for these isotopes. 183W, highlighted with the * texture, has a more demanding pattern because of the hyperfine-Zeeman mixing.

Close modal

The spectrometer simultaneously detects emission for the two lines of sight, as indicated in Fig. 1. One can position two lamps in front of the collecting optics of the 90° and 0° directions to obtain the pixel range covered by either LOS. Figure 4(a) presents the results of this procedure: In both cases, the vertical pixel range is about 50, and they are sufficiently separated to ensure that they do not influence each other.

FIG. 4.

(a): Backlight illumination detected by the camera. The brighter pixels correspond to those covered by the respective LOS. (b): Spectrum of two hollow cathode lamps binned across the vertical 50 pixels for the 0° LOS. (c): The instrumental profile of the spectrometer, as derived by optimizing Voigt functions, with the FWHM of the Gaussian being σ= 3 pm, and that of the Lorentzian γ= 0.3 pm. (a) Backlight illumination (b) Line identification (c) Fitting of the spectrometer profile.

FIG. 4.

(a): Backlight illumination detected by the camera. The brighter pixels correspond to those covered by the respective LOS. (b): Spectrum of two hollow cathode lamps binned across the vertical 50 pixels for the 0° LOS. (c): The instrumental profile of the spectrometer, as derived by optimizing Voigt functions, with the FWHM of the Gaussian being σ= 3 pm, and that of the Lorentzian γ= 0.3 pm. (a) Backlight illumination (b) Line identification (c) Fitting of the spectrometer profile.

Close modal

The high resolution of the spectrometer makes it necessary to pay special attention to the calibration. Several hollow cathode lamps (HCLs) that sputter metal and create the line emission of said metals are available to derive the reciprocal linear dispersion.48 As it turns out, two Ni I lines of detectable strength are sufficiently close to the target line of W I 498.4 nm. For calibrating the 0° LOS of the spectrometer, we focused the emission by the Ni and W HCLs on the same vertical pixels as those covered by the 0° LOS of Fig. 4(a). Figure 4(b) shows the vertically binned spectra of the HCLs. The Ni spectrum has an additional constant offset to distinguish it from the W spectrum. Lowercase Latin letters from a to g designate each noticeable spectral line, and Table II presents the unambiguous line identification. This identification provides the data for deriving the linear wavelength dispersion of the spectrometer. Of note is that the tungsten lamp (Heraeus, 37 mm HCL, element W) contains a bit of nickel, which explains the small line intensities at positions b and c.

TABLE II.

Line identification by using the vacuum wavelengths of neutral W and Ni as tabulated in the NIST ASD.42,45,46

Line a b c d e f g
W I  498.40 nm  ⋯  ⋯  498.12 nm  498.49 nm  498.56 nm  498.61 nm 
Ni I  ⋯  498.16  498.55  ⋯  ⋯  ⋯  ⋯ 
Line a b c d e f g
W I  498.40 nm  ⋯  ⋯  498.12 nm  498.49 nm  498.56 nm  498.61 nm 
Ni I  ⋯  498.16  498.55  ⋯  ⋯  ⋯  ⋯ 

The next step is to extract the instrumental profile from the calibration spectrum of the tungsten lamp. Here, it is of the essence to include the isotope and hyperfine splitting. Figure 4(c) shows the spectrum of the W HCL at 498.4 nm and the contribution by each isotope. The hyperfine interaction within 183W splits the transition into two lines. This splitting occurs due to the coupling of I = 1/2 and J = 1 for the upper level (7F1°), resulting in the coupled quantum number of Fu=1/2 and Fu=3/2. These states have slightly different energy eigenvalues and have allowed E1 transitions into the ground state with F = 1/2. For a detailed description of calculating the line positions and intensities for odd isotopes like 183W, see Ref. 23 

The manuscript Ref. 22 which introduced the space-resolved model, showed a distinctly different dependency of the lines of sight at 0 and 90 degrees on the angular distribution function. For example, increasing the average off-normal angles creates more Doppler shift in the 90° direction and reduces the Doppler shift for the 0° angle of observation. As opposed to that, changing the VDF affects both lines of sight the same.

For visualizing this behavior, it is possible to calculate a large set of synthetic line shapes by varying the parameters s in Eq. (1) and b of Eq. (3), as Fig. 5(a) illustrates. Calculating the spectra's full width at quarter maxima (FWQM) is an excellent measure of their general line shape. Figures 5(b) and 5(c) show the FWQM depending on the parameters s and b via a 2D contour plot.

FIG. 5.

Full width at quarter maxima (FWQM) of synthetic spectra calculated within SRM2 (Ref. 22) by assuming a kinetic energy of the bombarding Ar+ of 71 eV, a surface binding energy of 8.68 eV,49 z-distance of the 90° LOS to the target of 2 mm, and a reflectance of 54%. (a) depicts example spectra calculated for one combination of s and b. The gray-colored horizontal line indicates the FWQM position. The FWQM of both spectra are given in picometers. Varying the parameters (s, b), saving the FWQMs, and plotting them gives the contour plots of (b) and (c). (a) Synthetic spectra (b) 0° LOS (c) 90° LOS.

FIG. 5.

Full width at quarter maxima (FWQM) of synthetic spectra calculated within SRM2 (Ref. 22) by assuming a kinetic energy of the bombarding Ar+ of 71 eV, a surface binding energy of 8.68 eV,49 z-distance of the 90° LOS to the target of 2 mm, and a reflectance of 54%. (a) depicts example spectra calculated for one combination of s and b. The gray-colored horizontal line indicates the FWQM position. The FWQM of both spectra are given in picometers. Varying the parameters (s, b), saving the FWQMs, and plotting them gives the contour plots of (b) and (c). (a) Synthetic spectra (b) 0° LOS (c) 90° LOS.

Close modal

Superimposing the isolines provides a situation like those found in line-ratio methods.50 For example, Fig. 3 of Ref. 51 represents such a case, showing temperature and density-sensitive helium line ratios. However, it is important to note that the one-parameter ADF and VDF are approximate solutions and cannot fully capture the intricate details of the distribution functions. Still, fitting these to experimental data should at least provide qualitative information about the average off-normal angle.

Simulations of the outgoing angle distribution of the sputtered particles were carried out with the PARCAS Molecular Dynamics code52,53 using an adaptive time step.54 The interatomic potential by Mason et al.55 was used for the W-W interactions and the ZBL56 universal repulsive potential for Ar-Ar and Ar-W interactions. The argon atoms were given an energy of 100 eV at normal incidence to the surfaces. We focused on the (100) and (111) low-index surfaces, which can be studied during cumulative impacts. 10 000 impacts were carried out on the surfaces, and five runs were carried out for statistics. The cells were initially perfect, but could evolve during the continuous irradiation. The calculated outgoing angular distributions are the values over the whole run and for all the five identical starting positions. Single-impact simulations were carried out on the same (100) and (111) surfaces in addition to random surfaces, following the technique in Ref. 19. For single-impact simulations, five different relaxations of the same low-index surfaces were irradiated by 10 000 impacts each, and for the random surfaces, 30 random surface orientations were studied and 5 000 impacts for each surface.

The simulation cells for the cumulative impact simulations were about 6×6×12nm3, containing around 30 000 atoms, and had a two-layer structure. At the bottom, a few layers of atoms were fixed, and the rest of the atoms were mobile. The simulation was carried out in two steps, similar to previous studies on cumulative impacts,57 showing good agreement with experiments. First, an impact simulation, where the argon atom was given energy, was conducted for 5 ps. During this, no thermostat was applied, and any atom that was sputtered or reflected during the simulation was deleted after the simulation was done. An atom was considered sputtered/reflected if it was higher above the surface than the cutoff of the potential. After the impact simulation, a relaxation simulation was carried out to keep the temperature at 300 K without building up during the series. During this 5 ps simulation, a Berendsen thermostat58 was applied to all mobile atoms to extract the extra heat.

The simulation cells for the single-impact simulations were half-spheres with a radius of 5 nm, containing about 17 000 atoms. The cells had a shell-like structure, where the outermost atoms were fixed, and between the fixed atoms and the mobile atoms in the center, a few layers of atoms were thermally controlled. The same setup has been previously used for low-index and random surface orientation sputtering simulations, see Ref. 19 for details. The impacting Ar atoms always hit a square on the center of the pristine cell, with a side length of 2 nm. The simulation lasted for 5 ps, and an atom was considered sputtered/reflected if it was higher above the surface than the cutoff of the potential.

The line shapes of optical transitions by sputtered atoms measured in the linear plasma device PSI-2 generally exhibit asymmetries to the central transition wavelength. For instance, Fig. 6 presents an asymmetric spectrum of the W I 498.4 nm line measured along the 0° LOS while applying a bias of −80 V to the target. The actual kinetic energy of Ar+ bombarding the tungsten sample at normal incidence depends on the potential difference between the plasma and the metallic target. In the case of argon discharges, the plasma potential at the radial locations covered by the sample is around −9 V. A local negative plasma potential develops due to the strongly magnetized electrons with a typical Larmor radius of 0.05 mm. The Ar ions, however, have a Larmor radius of roughly 10 mm, demonstrating much larger cross field transport for the ions, such that a negative plasma potential is required to ensure the quasineutrality of the plasma.

FIG. 6.

Contribution of the even and odd tungsten isotopes to the line shape at the 0° LOS. The direct emission is blue-shifted since the particles move toward the observer. Additionally, one detects red-shifted emission due to reflection at the surface of the target.

FIG. 6.

Contribution of the even and odd tungsten isotopes to the line shape at the 0° LOS. The direct emission is blue-shifted since the particles move toward the observer. Additionally, one detects red-shifted emission due to reflection at the surface of the target.

Close modal

The most dominant spectral feature as shown in Fig. 6 is the direct emission: This emission by the sputtered atoms moving toward the observer shifts toward lower wavelengths—blue-shifted component. However, the tungsten atoms radiate isotropically, meaning they emit with the same probability into the direction of the target red-shifted light. The reflectance of the surface then decreases the signal of the red-shifted emission. In the case of mirror-like polished tungsten, one expects a reflectance of approximately 53%,59 which agrees very nicely with the value of 54% as extracted from fitting the line shape of the data presented in Fig. 6.

A similar example is exposing a mirror-like polished Al target to an argon plasma at PSI-2. Those measurements verify that accounting for light reflection at the target is crucial for modeling the line shape along the 0° LOS: the bombardment with argon ions for one hour significantly roughens the surface, resulting in a shift from specular to diffusive reflectance. Therefore, the intensity of the red-shifted component decreases substantially compared to the blue-shifted one, as Fig. 4 of Ref. 30 shows for the case of the Al I 396.15 nm line. The additional experimental proof is the Hα emission by backscattered hydrogen atoms at PSI-2. Ref. 60 shows that the reflectance plays a crucial role in the line shape, and the intensity ratio of red to blue-shifted components depends on the reflectance of the target material, as expected.

Other significant contributions to the line shape of emission by sputtered atoms, such as tungsten, are isotopic, hyperfine, and Zeeman splitting, convolved with instrumental broadening, as discussed in Sec. III.

The data plotted in Fig. 6 directly provides information on the velocity distribution function in the 0° LOS direction. That is, the typical shift from the center wavelength to the peak of the blue-shifted emission is around 3.3 pm, corresponding to a velocity of 2000 ms−1. This velocity allows for an estimation of the typical ionization length. The ionization rates from Vainshtein et al.61 and Schlummer et al.62 for an electron temperature of 3 eV give the ionization length, assuming a density of 2.5×1018m3, of approximately 40 mm. This length is significantly larger than the dimensions of the target (13 mm × 13 mm). Thus, the influence of ionization is negligible compared to geometric losses. Figures 6(a) and 6(b) of Ref. 22 show precisely this behavior based on the line shape simulation via SRM1. Furthermore, the tungsten ions are practically unmagnetized, as discussed in Sec. III.

The LOS at 90° does not intersect the target, so reflectance is irrelevant. Further assuming an azimuthal isotropy of the ADF and a plasma with homogeneous temperature and density, one would naturally expect a symmetric line shape centered around the unshifted wavelength. In this study, the target is positioned in one maximum of the hollow profile of the plasma (cf. Fig. 1). This geometry may produce a slight asymmetry: Some particles moving toward the observer (blue-shifted emission) move through the two plasma maxima. In contrast, particles moving away from the observer (red-shifted emission) pass through only one plasma maximum. Of course, the geometric losses of the particles out of the 90° LOS before reaching the second plasma maximum impose a limit to the asymmetry.

We implicitly assume azimuthal isotropy when analyzing the experimental data, and the model used in this study (SRM2 of Ref. 22) assumes excitation and ionization occur solely in the region in front of the target, producing strictly symmetric line shapes for the 90° angle of observation.

Figures 7(a) and 7(b) depict the spectra measured at 0° and 90° while applying a bias of −60 V to the target. The potential difference to the plasma provides the argon ions with a kinetic energy of 51 eV, and the magnetic field topology in the exposure area of PSI-2 ensures bombardment at normal incidence. The spectra measured along the 0° LOS have two essential differences: First, the surface structure of the three targets investigated was different—polycrystalline target, (100) and (111) planes. Second, the Zeeman pattern in the case of the W poly and W (111) targets was σ+, whereas that of W (100) was σ. This fact explains the slight difference in line position and resulted from the measurement procedure. We started by exposing the polycrystalline target, varied the bias, and rotated the quarter waveplate by 90° (Fig. 1) to prove that optical isolation worked as intended (cf. Ref. 30) For the next target, we kept this alignment, exposed the W (100) sample to the plasma, and repeated the same procedure for W (111).

FIG. 7.

Line shape of W I 498.4 nm line from mono- (100) and (111) and polycrystalline W at the 0° LOS (a) and 90° LOS (b). Each color corresponds to a different target. The spectra in (b) reveal a heart-shaped emission distribution during the sputtering of the (111) target. (a) 0° LOS, 51 eV (b) 90° LOS, 51 eV.

FIG. 7.

Line shape of W I 498.4 nm line from mono- (100) and (111) and polycrystalline W at the 0° LOS (a) and 90° LOS (b). Each color corresponds to a different target. The spectra in (b) reveal a heart-shaped emission distribution during the sputtering of the (111) target. (a) 0° LOS, 51 eV (b) 90° LOS, 51 eV.

Close modal

Comparing the spectrum due to sputtering from the poly and (111) target shows that the distribution functions must differ substantially. However, the difference in linewidth could result from changes in the VDF and ADF, as Fig. 5(b) illustrates via the isolines. Given the significant instrumental profile, it is impossible to deduce the cause for this variation by analyzing solely the 0° LOS. Therefore, we included measurements at the 90° LOS, and the variation in line shape is considerably more pronounced [cf. Fig. 7(b)]. The poly target produces a spectrum having a singular peak shape, (100) creates a flat-top-like structure with a width of a few pico meters, and (111) results in a two-peak structure owing to a considerable dip in the center of the line. This additional data are sufficient to make at least qualitative claims about the distribution functions because the width at 0° and 90° depends opposite on the ADF [cf. Figs. 5(b) and 5(c)].

Repeating these measurements at −80 V and −100 V bias provided the data needed to extract information on the VDF and ADF by fitting the one-parameter distribution functions of Eqs. (1) and (3). Again, the model underlying the analysis is SRM2,22 and we used a parameter tolerance of 0.01, resulting in decent fits to the data. Figures 8(a)–8(f) show these experimental data and the fits represented by solid lines. Tables III and IV give the optimized parameters s and b, respectively. The parameter s of the Thompson VDF is significantly different for the several targets investigated in this study. Those of the polycrystalline target are the closest to the standard value of s = 3. Quite a few mechanisms could result in this difference: The surface binding energy is not identical for all crystalline planes, and the azimuthal isotropy assumed in the modeling is only strictly fulfilled by the poly target. Furthermore, the exponentially modified cosine ADF may not capture all cases equally well.

FIG. 8.

Experimental spectra measured along the 0° and 90° LOS. The solid lines show the fit, as obtained by iteratively fitting the spectra: 0° spectra provide the data to optimize the parameter of the VDF s, and 90° spectra are used to fit the ADF parameter b. Tables III and IV list the parameters determined in this way. (a) W poly 0° (b) W (100) 0° (c) W (111) 0° (d) W poly 90° (e) W (100) 90° (f) W (111) 90°.

FIG. 8.

Experimental spectra measured along the 0° and 90° LOS. The solid lines show the fit, as obtained by iteratively fitting the spectra: 0° spectra provide the data to optimize the parameter of the VDF s, and 90° spectra are used to fit the ADF parameter b. Tables III and IV list the parameters determined in this way. (a) W poly 0° (b) W (100) 0° (c) W (111) 0° (d) W poly 90° (e) W (100) 90° (f) W (111) 90°.

Close modal
TABLE III.

VDF scaling parameter s. Note that the kinetic energy of the bombarding ion reduces by about −9 eV due to the negative plasma potential. Values presented in brackets refer to one standard deviation, as obtained from the fitting procedure, relevant to the last digits given.

−60 V −80 V −100 V
Poly  2.59(10)  2.94(4)  2.92(2) 
(100)  2.00(11)  2.62(5)  2.71(3) 
(111)  1.99(10)  2.58(6)  2.46(2) 
−60 V −80 V −100 V
Poly  2.59(10)  2.94(4)  2.92(2) 
(100)  2.00(11)  2.62(5)  2.71(3) 
(111)  1.99(10)  2.58(6)  2.46(2) 
TABLE IV.

ADF scaling parameter b. Note that the kinetic energy of the bombarding ion reduces by about −9 eV due to the negative plasma potential. Values presented in brackets refer to one standard deviation, as obtained from the fitting procedure, relevant to the last digits given.

−60 V −80 V −100 V
Poly  −1.47(10)  −1.24(5)  −0.95(5) 
(100)  −1.85(14)  −1.86(11)  −1.60(11) 
(111)  −3.5(9)  −2.51(11)  −2.04(10) 
−60 V −80 V −100 V
Poly  −1.47(10)  −1.24(5)  −0.95(5) 
(100)  −1.85(14)  −1.86(11)  −1.60(11) 
(111)  −3.5(9)  −2.51(11)  −2.04(10) 

Therefore, we compare through Figs. 9(a)–9(c) the extracted ADFs to MD simulations performed as discussed in Sec. IV. The single-impact MD and fitted ADF are fairly similar in the case of the poly target. Higher kinetic energies of the impinging argon ions create, on average, lower off-normal angles, so the distribution function slowly tends toward a cosine one. We did not conduct cumulative impact simulations for the polycrystalline target because our setup cannot handle such simulations on random surfaces. However, single-impact simulations of random surfaces are arguably sufficient for modeling the erosion of polycrystalline surfaces due to their inherently rough and variable nature.

FIG. 9.

Angular distribution functions (ADFs) and velocity distribution functions (VDFs) as obtained from the experimental spectra [Figs. 8(a)–8(f)] with SRM2.22 The black circles plotted in (a) to (c) correspond to single-impact MD simulations, whereas those depicted via squares represent the result of cumulative impact MD simulations. (a) W poly ADF (b) W (100) ADF (c) W (111) ADF (d) W poly VDF (e) W (100) VDF (f) W (111) VDF.

FIG. 9.

Angular distribution functions (ADFs) and velocity distribution functions (VDFs) as obtained from the experimental spectra [Figs. 8(a)–8(f)] with SRM2.22 The black circles plotted in (a) to (c) correspond to single-impact MD simulations, whereas those depicted via squares represent the result of cumulative impact MD simulations. (a) W poly ADF (b) W (100) ADF (c) W (111) ADF (d) W poly VDF (e) W (100) VDF (f) W (111) VDF.

Close modal

The sputtering of tungsten atoms from the (111) plane in the near-threshold regime leads to the most probable emission angle at the highest θ found in this study. The results of single and cumulative impact MD simulations at 100 eV, plotted in Fig. 9(b), suggest an even more peaked emission profile. There is a slightly better agreement between the cumulative impact simulations and the fitted ADFs, which makes sense because one expects some modification of the surface structure due to continuous plasma exposure. The chief contributor to the discrepancy between MD simulation and fitted ADFs is most likely the one-parameter approximation of the ADF, which cannot produce such a well-defined peak. However, the at least qualitative agreement in the case of the mono targets (111) and (100) verifies the line shape emission spectroscopy for monitoring the ADF of sputtered particles in plasma exposure experiments.

This study reports on the line shape of the resonant W I 498.4 nm transition by atoms sputtered in the linear plasma device PSI-2. The main goal was to verify that our approach of using two observation angles can capture variation in the angular distribution function (ADF). Work by Jussila et al. suggested pronounced differences in the ADF of low-index crystalline tungsten mono to poly targets when bombarding them with low-energy argon ions.19 

Hence, we exposed (111) (100), and polycrystalline tungsten targets to an argon plasma at the linear plasma device PSI-2. The argon ions bombarded the samples at normal incidence with kinetic energies below 100 eV, providing near-threshold sputtering. These sputtered particles transversed through the plasma and were excited via electron collisions. Measuring the resulting line-integrated electric dipole radiation via the 0° and 90° observation angles provided the experimental data analyzed in this work. Most critical is the comparison of the line shape that Figs. 7(a) and 7(b) show. Indeed, the line shape of the W I 498.4 nm transition differs depending on the target we exposed to the plasma. These results represent the first measurements using the line shape emission spectroscopy that show such distinct differences solely due to the surface structure.

This variation in line shape is consistent with the expectation, as given by the model benchmarked in Ref. 22 and the information on the ADF.19 The (111) target creates the broadest spectra along the 90° direction and the narrowest ones in the 0° direction. Fitting the experimental data with the recently developed SRM222 while using a parameterized Thompson VDF and an exponentially modified cosine ADF provided consistent fits to the line profiles. The main idea was to optimize the spectra for 0° and 90° iteratively due to their opposite dependency on the ADF, as Figs. 5(a)–5(c) illustrate.

Comparing the deduced ADFs to the ones simulated using the PARCAS MD code for 100 eV Ar+ bombarding tungsten surfaces at normal incidence gives a decent agreement for the polycrystalline target. In the case of the monocrystalline targets, we performed single and cumulative impact MD simulations. The cumulative results agree slightly better with the ADFs, as derived from fitting the emission spectra at 0° and 90°. However, the one-parameter ADF used in this study seems less effective at accurately representing the highly peaked emission of sputtered atoms within a narrow angular interval from low-index monocrystalline planes. Although a two-cosine power-law scaling can capture these distributions, the considerable instrumental broadening due to the spectrometer would cause ambiguity in the solutions. Nonetheless, the qualitative comparison is still reasonable: the (111) target creates a peaked emission profile at large off-normal angles, as illustrated by the dip in the profile at the unshifted wavelengths [cf. Fig. 7(b)].

Additionally, we did not monitor the modification of the surface morphology in this campaign. The continuous plasma exposure of the targets shifts their mirror-like specular reflectance to the diffusive one. We will address this problem soon and investigate the time evolution of the angular distribution as for instance, was proposed for Al in Ref. 30. This method can distinguish the sputtering of re-deposited and bulk tungsten.

This study demonstrates the ability of high-resolution emission spectroscopy to detect the difference in the internal structure of plasma-contacting surfaces by analyzing the spectral line shape of atomic transitions. More specifically, our work confirms its reliability in deducing information on the ADF during the sputtering of plasma-facing components. This technique's capability for capturing time-resolved changes in the line shape could be critical for studying samples where the ADF is expected to change during plasma exposure. A prerequisite for this approach is that one needs at least two observation angles to distinguish between spectral line broadening due to the VDF and ADF. Particularly, advantageous are those angles parallel and perpendicular to the normal, as contour plots of the simulated linewidth show [cf. Figures 5(b) and 5(c)].

This work has been carried out within the framework of the EUROfusion Consortium, funded by the European Union via the Euratom Research and Training Programme (Grant Agreement No 101052200—EUROfusion). Views and opinions expressed are however those of the author(s) only and do not necessarily reflect those of the European Union or the European Commission. Neither the European Union nor the European Commission can be held responsible for them.

Computer time granted by the IT Center for Science—CSC—Finland is gratefully acknowledged.

The authors have no conflicts to disclose.

M. Sackers: Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). O. Marchuk: Conceptualization (equal); Investigation (equal); Methodology (equal); Project administration (equal); Supervision (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal). S. Ertmer: Investigation (equal); Validation (equal); Writing – review & editing (equal). S. Brezinsek: Funding acquisition (equal); Project administration (equal); Resources (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal). F. Granberg: Data curation (equal); Formal analysis (equal); Software (equal); Validation (equal); Writing – original draft (equal); Writing – review & editing (equal). A. Kreter: Resources (equal); Writing – original draft (equal); Writing – review & editing (equal).

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

1.
M.
Knudsen
, “
Die Verdampfung von Kristalloberflächen
,”
Ann. Phys.
357
,
105
108
(
1917
).
2.
G. K.
Wehner
, “
Sputtering of metal single crystals by ion bombardment
,”
J. Appl. Phys.
26
,
1056
1057
(
1955
).
3.
S.
Brezinsek
,
A.
Kirschner
,
M.
Mayer
,
A.
Baron-Wiechec
,
I.
Borodkina
,
D.
Borodin
,
I.
Coffey
,
J.
Coenen
,
N.
den Harder
,
A.
Eksaeva
,
C.
Guillemaut
,
K.
Heinola
,
A.
Huber
,
V.
Huber
,
M.
Imrisek
,
S.
Jachmich
,
E.
Pawelec
,
M.
Rubel
,
S.
Krat
,
G.
Sergienko
,
G. F.
Matthews
,
A. G.
Meigs
,
S.
Wiesen
, and
A.
Widdowson
, “
Erosion, screening, and migration of tungsten in the JET divertor
,”
Nucl. Fusion
59
,
096035
(
2019
).
4.
R. A.
Pitts
,
S.
Carpentier
,
F.
Escourbiac
,
T.
Hirai
,
V.
Komarov
,
A. S.
Kukushkin
,
S.
Lisgo
,
A.
Loarte
,
M.
Merola
,
R.
Mitteau
,
A. R.
Raffray
,
M.
Shimada
, and
P. C.
Stangeby
, “
Physics basis and design of the ITER plasma-facing components
,”
J. Nucl. Mater.
415
,
S957
S964
(
2011
).
5.
T.
Pütterich
,
R.
Neu
,
R.
Dux
,
A. D.
Whiteford
,
M. G.
O'Mullane
, and
H. P.
Summers
, “
Calculation and experimental test of the cooling factor of tungsten
,”
Nucl. Fusion
50
,
025012
(
2010
).
6.
T.
Pütterich
,
E.
Fable
,
R.
Dux
,
M.
O'Mullane
,
R.
Neu
, and
M.
Siccinio
, “
Determination of the tolerable impurity concentrations in a fusion reactor using a consistent set of cooling factors
,”
Nucl. Fusion
59
,
056013
(
2019
).
7.
A.
Tabasso
,
H.
Maier
,
J.
Roth
, and
K.
Krieger
, “
Studies of tungsten erosion at the inner and outer main chamber wall of the ASDEX Upgrade tokamak
,”
J. Nucl. Mater.
290–293
,
326
330
(
2001
).
8.
K.
Sugiyama
,
K.
Schmid
, and
W.
Jacob
, “
Sputtering of iron, chromium and tungsten by energetic deuterium ion bombardment
,”
Nucl. Mater. Energy
8
,
1
7
(
2016
).
9.
A.
Goehlich
,
N.
Niemöller
, and
H.
Döbele
, “
Determination of angle resolved velocity distributions of sputtered tungsten atoms
,”
J. Nucl. Mater.
266–269
,
501
506
(
1999
).
10.
M. T.
Robinson
and
I. M.
Torrens
, “
Computer simulation of atomic-displacement cascades in solids in the binary-collision approximation
,”
Phys. Rev. B
9
,
5008
(
1974
).
11.
N. A.
Kubota
,
D. J.
Economou
, and
S. J.
Plimpton
, “
Molecular dynamics simulations of low-energy (25–200 ev) argon ion interactions with silicon surfaces: Sputter yields and product formation pathways
,”
J. Appl. Phys.
83
,
4055
4063
(
1998
).
12.
T.
Tanabe
,
N.
Bekris
,
P.
Coad
,
C.
Skinner
,
M.
Glugla
, and
N.
Miya
, “
Tritium retention of plasma facing components in tokamaks
,”
J. Nucl. Mater.
313-316
,
478
490
(
2003
).
13.
M.
Zlobinski
,
G.
Sergienko
,
I.
Jepu
,
C.
Rowley
,
A.
Widdowson
,
R.
Ellis
,
D.
Kos
,
I. H.
Coffey
,
M.
Fortune
,
D.
Kinna
,
M.
Beldishevski
,
A.
Krimmer
,
H. T.
Lambertz
,
A.
Terra
,
A.
Huber
,
S.
Brezinsek
,
T.
Dittmar
,
M.
Flebbe
,
R.
Yi
,
R.
Rayaprolu
,
J.
Figueiredo
,
P.
Blatchford
,
S. A.
Silburn
,
E.
Tsitrone
,
E. H.
Joffrin
,
K.
Krieger
,
Y.
Corre
,
A.
Hakola
, and
J.
Likonen
, “
First results of laser-induced desorption - quadrupole mass spectrometry (LID-QMS) at JET
,”
Nucl. Fusion
64
,
086031
(
2024
).
14.
G. S.
Maurya
,
A.
Marín-Roldán
,
P.
Veis
,
A. K.
Pathak
, and
P.
Sen
, “
A review of the LIBS analysis for the plasma-facing components diagnostics
,”
J. Nucl. Mater.
541
,
152417
(
2020
).
15.
J.
Karhunen
,
A.
Hakola
,
J.
Likonen
,
A.
Lissovski
,
M.
Laan
, and
P.
Paris
, “
Applicability of LIBS for in situ monitoring of deposition and retention on the ITER-like wall of jet – comparison to SIMS
,”
J. Nucl. Mater.
463
,
931
935
(
2015
), part of special issue PLASMA-SURFACE INTERACTIONS 21.
16.
C.
Linsmeier
,
B.
Unterberg
,
J. W.
Coenen
,
R. P.
Doerner
,
H.
Greuner
,
A.
Kreter
,
J.
Linke
, and
H.
Maier
, “
Material testing facilities and programs for plasma-facing component testing
,”
Nucl. Fusion
57
,
092012
(
2017
).
17.
A.
Kallenbach
,
M.
Bernert
,
R.
Dux
,
L.
Casali
,
T.
Eich
,
L.
Giannone
,
A.
Herrmann
,
R.
McDermott
,
A.
Mlynek
,
H. W.
Müller
,
F.
Reimold
,
J.
Schweinzer
,
M.
Sertoli
,
G.
Tardini
,
W.
Treutterer
,
E.
Viezzer
,
R.
Wenninger
, and
M.
Wischmeier
, “
Impurity seeding for tokamak power exhaust: From present devices via ITER to DEMO
,”
Plasma Phys. Controlled Fusion
55
,
124041
(
2013
).
18.
D.
Nishijima
,
M. J.
Baldwin
,
R. P.
Doerner
, and
J. H.
Yu
, “
Sputtering properties of tungsten ‘fuzzy’ surfaces
,”
J. Nucl. Mater.
415
,
S96
S99
(
2011
).
19.
J.
Jussila
,
F.
Granberg
, and
K.
Nordlund
, “
Effect of random surface orientation on W sputtering yields
,”
Nucl. Mater. Energy
17
,
113
122
(
2018
).
20.
M.
Stepanova
and
S. K.
Dew
, “
Estimates of differential sputtering yields for deposition applications
,”
J. Vac. Sci. Technol. A: Vac. Surf. Films
19
,
2805
2816
(
2001
).
21.
S.
Ertmer
,
O.
Marchuk
,
M.
Sackers
,
S.
Dickheuer
,
S.
Brezinsek
,
P.
Mertens
, and
A.
Kreter
, “
Optical isolation of spectral lines emitted by sputtered tungsten in a weakly magnetized plasma
,”
J. Phys. B: At. Mol. Opt. Phys.
54
,
025401
(
2020
).
22.
M.
Sackers
,
O.
Marchuk
,
S.
Ertmer
,
S.
Brezinsek
, and
A.
Kreter
, “
Space-resolved line shape model for sputtered atoms of finite-size targets
,”
Phys. Scr.
98
,
115603
(
2023
).
23.
M.
Sackers
,
O.
Marchuk
,
D.
Dipti
,
Y.
Ralchenko
,
S.
Ertmer
,
S.
Brezinsek
, and
A.
Kreter
, “
Zeeman effect of isotopes of Kr and Xe investigated at the linear plasma device PSI-2
,”
Plasma Sources Sci. Technol.
33
,
025015
(
2024
).
24.
M.
Baldwin
and
R.
Doerner
, “
Helium induced nanoscopic morphology on tungsten under fusion relevant plasma conditions
,”
Nucl. Fusion
48
,
035001
(
2008
).
25.
A.
Kreter
,
C.
Brandt
,
A.
Huber
,
S.
Kraus
,
S.
Möller
,
M.
Reinhart
,
B.
Schweer
,
G.
Sergienko
, and
B.
Unterberg
, “
Linear plasma device PSI-2 for plasma-material interaction studies
,”
Fusion Sci. Technol.
68
,
8
14
(
2015
).
26.
D. M.
Goebel
,
G.
Campbell
, and
R. W.
Conn
, “
Plasma surface interaction experimental facility (PISCES) for materials and edge physics studies
,”
J. Nucl. Mater.
121
,
277
282
(
1984
).
27.
N.
Ohno
,
D.
Nishijima
,
S.
Takamura
,
Y.
Uesugi
,
M.
Motoyama
,
N.
Hattori
,
H.
Arakawa
,
N.
Ezumi
,
S.
Krasheninnikov
,
A.
Pigarov
, and
U.
Wenzel
, “
Static and dynamic behaviour of plasma detachment in the divertor simulator experiment NAGDIS-II
,”
Nucl. Fusion
41
,
1055
1065
(
2001
).
28.
J.
Rapp
,
W. R.
Koppers
,
H.
van Eck
,
G. J.
van Rooij
,
W. J.
Goedheer
,
B.
de Groot
,
R.
Al
,
M. F.
Graswinckel
,
M. A.
van den Berg
,
O.
Kruyt
,
P.
Smeets
,
H. J.
van der Meiden
,
W.
Vijvers
,
J.
Scholten
,
M.
van de Pol
,
S.
Brons
,
W.
Melissen
,
T.
van der Grift
,
R.
Koch
,
B.
Schweer
,
U.
Samm
,
V.
Philipps
,
R.
Engeln
,
D. C.
Schram
,
N. J.
Lopes Cardozo
, and
A. W.
Kleyn
, “
Construction of the plasma-wall experiment Magnum-PSI
,”
Fusion Eng. Des.
85
,
1455
1459
(
2010
).
29.
M.
Sackers
,
O.
Marchuk
,
S.
Ertmer
,
S.
Dickheuer
,
U.
Czarnetzki
,
T. V.
Tsankov
,
D.
Luggenhölscher
,
S.
Brezinsek
, and
A.
Kreter
, “
Zeeman-resolved TDLAS using metastable levels of Ar in the weakly magnetized plasma of the linear plasma device PSI-2
,”
J. Phys. D: Appl. Phys.
54
,
395001
(
2021
).
30.
S.
Ertmer
,
O.
Marchuk
,
S.
Dickheuer
,
M.
Rasiński
,
A.
Kreter
, and
S.
Brezinsek
, “
Light-reflection-induced changes in the line shape of sputtered atoms
,”
Phys. Scr.
T171
,
014031
(
2020
).
31.
M. W.
Thompson
, “
II. The energy spectrum of ejected atoms during the high energy sputtering of gold
,”
Philos. Mag.
18
,
377
414
(
1968
).
32.
M.
Sackers
,
C.
Busch
,
T. V.
Tsankov
,
U.
Czarnetzki
,
P.
Mertens
, and
O.
Marchuk
, “
Plasma parameters and tungsten sputter rates in a high-frequency CCP
,”
Phys. Plasmas
29
,
043511
(
2022
).
33.
Y. V.
Martynenko
,
A. V.
Rogov
, and
V. I.
Shul'ga
, “
Angular distribution of atoms during the magnetron sputtering of polycrystalline targets
,”
Tech. Phys.
57
,
439
444
(
2012
).
34.
A.
Eksaeva
,
E.
Marenkov
,
D.
Borodin
,
A.
Kreter
,
M.
Reinhart
,
A.
Kirschner
,
J.
Romazanov
,
A.
Terra
,
S.
Brezinsek
, and
K.
Nordlund
, “
ERO modelling of tungsten erosion in the linear plasma device PSI-2
,”
Nucl. Mater. Energy
12
,
253
260
(
2017
).
35.
S.
Dickheuer
,
O.
Marchuk
,
Y.
Krasikov
,
P.
Mertens
,
C.
Brandt
,
S.
Ertmer
,
P.
Börner
,
D.
Reiter
,
B.
Göths
,
K.
von Bovert
, and
A.
Kreter
, “
Polarization by light reflection at metallic surfaces observed in the shape of the Balmer- α line of low density plasmas
,”
Phys. Plasmas
26
,
073513
(
2019
).
36.
Y.
Yamamura
,
T.
Takiguchi
, and
M.
Ishida
, “
Energy and angular distributions of sputtered atoms at normal incidence
,”
Radiat. Eff. Defects Solids
118
,
237
261
(
1991
).
37.
S.
Ertmer
, “
High-resolution spectroscopy studies on sputtered atoms in the linear plasma device PSI-2
,” Doctoral thesis (
Ruhr-Universität Bochum, Universitätsbibliothek
,
2022
).
38.
J.
Völkening
,
M.
Köppe
, and
K. G.
Heumann
, “
Tungsten isotope ratio determinations by negative thermal ionization mass spectrometry
,”
Int. J. Mass Spectrom. Ion Processes
107
,
361
368
(
1991
).
39.
K. J. R.
Rosman
and
P. D. P.
Taylor
, “
Isotopic compositions of the elements 1997 (Technical Report)
,”
Pure Appl. Chem.
70
,
217
235
(
1998
).
40.
H.
Kopfermann
and
E. E.
Schneider
,
Nuclear Moments, Pure and Applied Physics
(
Academic Press
,
New York
,
1958
), Vol.
2
.
41.
R.-J.
Champeau
and
M.
Miladi
, “
Déplacement isotopique relatif dans le spectre d'arc du tungstène. structure hyperfine du niveau 5d5 6s 7s3 de WI
,”
J. Phys.
35
,
105
111
(
1974
).
42.
A.
Kramida
,
Y.
Ralchenko
,
J.
Reader
, and
NIST ASD Team
. “
NIST Atomic Spectra Database, NIST Standard Reference Database 78
” (
NIST Atomic Spectra Database
,
1999
).
43.
A.
Corney
,
Atomic and Laser Spectroscopy
, 1st ed., Oxford Science Publications (
Clarendon Press
,
Oxford
,
1979
).
44.
N. J.
Stone
, “
Table of recommended nuclear magnetic dipole moments: Part I - Long-lived states
” (
International Atomic Energy Agency
,
2019
).
45.
D. D.
Laun
and
C. H.
Corliss
, “
The first spectrum of Tungsten (W I)
,”
J. Res. Natl. Bur. Stand. Sect. A
72A
,
609
755
(
1968
).
46.
W. F.
Meggers
,
C. H.
Corliss
, and
B. F.
Scribner
,
Tables of Spectral-Line Intensities
(
National Bureau of Standards
,
1975
).
47.
R. D.
Cowan
,
The Theory of Atomic Structure and Spectra
, Los Alamos series in basic and applied sciences (
University of California Press
,
Berkeley, Calif
.,
1981
), Vol.
3
.
48.
H.-J.
Kunze
,
Introduction to Plasma Spectroscopy
(
Springer Science & Business Media
,
2009
), Vol.
56
.
49.
R.
Behrisch
, ed.,
Sputtering by Particle Bombardment: Experiments and Computer Calculations from Threshold to MeV Energies
, Topics in Applied Physics (
Springer-Verlag GmbH
,
Berlin Heidelberg
,
2007
), Vol.
110
.
50.
S.
Siepa
,
S.
Danko
,
T. V.
Tsankov
,
T.
Mussenbrock
, and
U.
Czarnetzki
, “
On the OES line-ratio technique in argon and argon-containing plasmas
,”
J. Phys. D: Appl. Phys.
47
,
445201
(
2014
).
51.
O.
Schmitz
,
I. L.
Beigman
,
L. A.
Vainshtein
,
B.
Schweer
,
M.
Kantor
,
A.
Pospieszczyk
,
Y.
Xu
,
M.
Krychowiak
,
M.
Lehnen
,
U.
Samm
, and
B.
Unterberg
, “
Status of electron temperature and density measurement with beam emission spectroscopy on thermal helium at TEXTOR
,”
Plasma Phys. Controlled Fusion
50
,
115004
(
2008
).
52.
K. N. M.
Ghaly
and
R. S.
Averback
, “
Molecular dynamics investigations of surface damage produced by kiloelectronvolt self-bombardment of solids
,”
Philos. Mag. A
79
,
795
820
(
1999
).
53.
K.
Nordlund
,
M.
Ghaly
,
R. S.
Averback
,
M.
Caturla
,
T.
Diaz de la Rubia
, and
J.
Tarus
, “
Defect production in collision cascades in elemental semiconductors and fcc metals
,”
Phys. Rev. B
57
,
7556
7570
(
1998
).
54.
K.
Nordlund
, “
Molecular dynamics simulation of ion ranges in the 1–100 keV energy range
,”
Comput. Mater. Sci.
3
,
448
456
(
1995
).
55.
D. R.
Mason
,
D.
Nguyen-Manh
,
V. W.
Lindblad
,
F.
Granberg
, and
M. Yu.
Lavrentiev
, “
An empirical potential for simulating hydrogen isotope retention in highly irradiated tungsten
,”
J. Phys. Condens. Matter
35
,
495901
(
2023
).
56.
J. F.
Ziegler
and
J. P.
Biersack
, “
The stopping and range of ions in matter
,” in
Treatise on Heavy-Ion Science: Volume 6: Astrophysics, Chemistry, and Condensed Matter
, edited by
D. A.
Bromley
(
Springer US
,
Boston, MA
,
1985
), pp.
93
129
.
57.
F.
Granberg
,
A.
Litnovsky
, and
K.
Nordlund
, “
Low energy sputtering of Mo surfaces
,”
J. Nucl. Mater.
539
,
152274
(
2020
).
58.
H. J. C.
Berendsen
,
J. P. M.
Postma
,
W. F.
van Gunsteren
,
A.
DiNola
, and
J. R.
Haak
, “
Molecular dynamics with coupling to an external bath
,”
J. Chem. Phys.
81
,
3684
3690
(
1984
).
59.
M.
Minissale
,
C.
Pardanaud
,
R.
Bisson
, and
L.
Gallais
, “
The temperature dependence of optical properties of tungsten in the visible and near-infrared domains: An experimental and theoretical study
,”
J. Phys. D: Appl. Phys.
50
,
455601
(
2017
).
60.
S.
Dickheuer
,
O.
Marchuk
,
C.
Brandt
,
A.
Pospieszczyk
,
A.
Goriaev
,
M.
Ialovega
,
B.
Göths
,
Y.
Krasikov
,
A.
Krimmer
,
P.
Mertens
, and
A.
Kreter
, “
In situ measurements of the spectral reflectance of metallic mirrors at the Hα line in a low density Ar–H plasma
,”
Rev. Sci. Instrum.
89
,
063112
(
2018
).
61.
L.
Vainshtein
,
I.
Beigman
,
P.
Mertens
,
S.
Brezinsek
,
A.
Pospieszczyk
, and
D.
Borodin
, “
Ionization of w atoms and w+ ions by electrons
,”
J. Phys. B: At. Mol. Opt. Phys.
44
,
125201
(
2011
).
62.
T.
Schlummer
,
S.
Ertmer
,
S.
Brezinsek
,
A.
Eksaeva
,
A.
Pospieszczyk
,
D.
Borodin
,
A.
Kirschner
,
A.
Kreter
,
C.
Linsmeier
,
G.
Sergienko
, and
B.
Unterberg
, “
Experimental data on low energy electron impact ionisation of W
,”
Phys. Scr.
T170
,
014075
(
2017
).