Brought to you by:
Paper The following article is Open access

On the applicability of three and four parameter fits for analysis of swept embedded Langmuir probes in magnetised plasma

, , , , , , , , , , , and

Published 9 August 2022 © 2022 The Author(s). Published on behalf of IAEA by IOP Publishing Ltd
, , Citation M. Komm et al 2022 Nucl. Fusion 62 096021 DOI 10.1088/1741-4326/ac8011

0029-5515/62/9/096021

Abstract

The problem of power exhaust is one of the grand challenges of nuclear fusion research today. In order to understand the physics phenomena occurring in the scrape-off layer and the divertor regions of tokamaks, it is essential to correctly determine the divertor plasma parameters, which are often measured by swept Langmuir probes (LPs). While the construction and operation of this diagnostic can be straightforward, the data analysis using three- or four-parameter fits presents a challenge and can potentially lead to erroneous values of electron temperature and ion saturation current. In this work, we present modelling and experiments aimed at determination of conditions for proper analysis of swept LPs using these two fitting models. Particle-in-cell modelling was employed to evaluate the sheath-expansion effects for particular probe geometry and plasma conditions, yielding a semi-empirical rule capable of predicting its magnitude. Experiments with unusually wide range of swept voltage in the divertor of the COMPASS tokamak explored the magnitude of voltage range required for successful analysis with either three or four-parameter fitting. With the use of our new semi-empirical rule, it is possible to improve the four-parameter fit reliability in situations where the available voltage range is limited. In addition, we introduce the tangent method—an independent and fast method of electron temperature estimation, which allows to reliably determine the available voltage range and as such assist more complex methods of probe analysis.

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

The Langmuir probe (LP) [1] is one of the oldest plasma diagnostics, having attracted generations of experimentalists thanks to its relatively easy construction and application in various types of plasmas. The current Ip collected by a LP at given bias Vp depends on local plasma parameters (such as electron temperature Te and density ne), which in principle allows to determine such parameters with good spatial resolution. However, in practice the interpretation of probe measurements proved to be a delicate task, which requires great attention to the dominant processes in the plasma as well as treatment of the measured data. A wide range of theoretical models for different plasma regimes has been developed (see the review of these models in [2]), however, measurements in certain regimes can still fail the assumptions of the simplified models. One such regime is magnetised plasma in the scrape-off layer of tokamaks, which is the point of focus of this work.

Current–voltage (IV) characteristics of swept LPs in tokamak plasma can exhibit several deviations from the original model proposed by Langmuir [1]. One of them affects the ion current when the bias voltage is substantially lower than the floating potential. Under certain conditions, this current is not insensitive to probe bias but increases with decreasing probe voltage—dubbed the sheath expansion effect. This effect is particularly pronounced in the case of flush-mounted probes oriented at grazing angles of incidence with respect to the magnetic field lines. As presented by Matthews [3], this non-saturation of the ion current lead to over-estimation of Te when a simple three-parameter fit was applied to the IV characteristics:

Equation (1)

where Isat is the ion saturation current and Vfloat the floating potential. In response to this experimental observation a number of models [46] were proposed to account for the non-saturation of the ion current. These can be represented in a generalised form

Equation (2)

The individual models provided different theoretical derivations of the values of coefficients αi and β (the latter is typically not used as a free parameter for fitting but fixed at value either 3/4 or 1). However, these models were based on a simplified treatment of processes occurring in the Debye sheath and the magnetic pre-sheath, leading to uncertainties in their predictions. In practice, the options for validation of measured quantities are either the comparison (or even calibration) against other diagnostics or dedicated 3D3V modelling using the particle-in-cell (PIC) method [7]. In this work we extend the modelling study on flush mounted probes [7] to the case of protruding probes which are employed in the divertor of the COMPASS tokamak.

The divertor of the COMPASS tokamak [8] has been equipped with a combined probe array, which comprises 55 probe triplets, each triplet consisting of two LPs and one ball-pen probe [9]. Under normal operating conditions, one of the LPs in each triplet is biased to constant large negative voltage (−270 V) in order to measure the ion saturation current Isat. For the experiments reported in this work, the bias of this probe was swept at a frequency of 1 kHz in the range of −320 V and +50 V to acquire the IV characteristic. This extended range of sweeping allowed to detect the sheath expansion effects and to determine how large sweeping range is required to obtain valid result with either three- or four-parameter fits of the IV characteristics. This work studies the problem of sheath expansion both from the side of analysis of experimental data as well as PIC modelling of the underlying processes.

The paper is organised as follows: section 2 presents results of PIC modelling of the probes, which allowed to develop a semi-empirical scaling law for the sheath expansion magnitude. Section 3 deals with analysis of experimental data from COMPASS divertor probes: section 3.1 presents results of various fitting models, including an improved four-parameter fit which makes use of the aforementioned scaling law, section 3.2 introduces the tangent method of Te estimation, which allows to determine independently the voltage range necessary for a proper fitting and section 3.3 discusses how large this voltage range is required for different fitting models. Finally, section 3.4 compares results of the fits to the ball-pen probe technique.

2. Sheath expansion modelling

2.1. Simulation description

The ion collection of a protruding LP corresponding to those in the divertor of COMPASS was simulated using the PIC method, which can properly resolve finite ion Larmor effects and non-quasineutral plasma within the Debye sheath. The geometry of the divertor probes is inherently three dimensional and so the preferred tool for its simulation is the 3D3V PIC code SPICE3 [7, 10, 11]. However, due to the size of the probe and the shallow angle of incidence of the magnetic field with respect to divertor tile surface, a full-scale calculation with realistic plasma parameters is unfortunately too computationally demanding. Instead, a subset of simulations with feasible plasma parameters was performed and compared with simulations of the 2D3V PIC code SPICE2 [12, 13] (2D simulations are in general computationally less demanding), which assumes infinite poloidal length of the probe. The differences between 2D and 3D simulations were characterised and finally the 2D simulations were performed for realistic plasma parameters and geometry.

The geometry of the simulations is depicted in figure 1. Particles are injected through the top boundary, the side boundaries are periodic and the bottom boundary represents the wall (for details of this scheme see [13]). The probe geometry is based on the LPs utilised in the combined LP and ball-pen probe array in COMPASS divertor [9]. The probe has a triangular cross-section in the toroidal direction, with base length Lp = 6.5 mm and height Hp = 1.5 mm. The poloidal cross-section is rectangular, with width Wp = 2.0 mm. The poloidal spacing between neighbouring probes of the array is Dp = 2.0 mm (note that in the context of the simulations, the poloidal direction coincides with the radial direction). The dimensions of the simulated area were chosen such that in the direction normal to the wall there was enough space above the probe to resolve the magnetic pre-sheath (Lz = Hp + 10rLi [14], where rLi is the ion Larmor radius). In the poloidal direction, the width of the simulation is simply Lx = Wp + Dp. The toroidal dimension was chosen to be long enough to capture the magnetic shadow cast by the probe (see later in this section). Given the fact that the boundaries at x = 0, x = Lx , y = 0 and y = Ly are periodic, this setup corresponds to a simulation of an infinite probe array or infinite probe matrix in case of 2D and 3D simulations respectively. During the course of the simulation, the particles were introduced through the top boundary and gradually filled an initially empty simulated volume. Once the simulation achieved saturation, the normalised potential of the probe Φ = (VpVplasma)/kTe was increased in a step-wise manner from −10 to +10 by increments of 0.1. The current during each step was averaged to form a data point on the IV characteristics. First 50% of time steps during each voltage step were discarded in order to avoid possible interference of the recorded current by the reaction of the simulated particles to the sudden change of probe bias. Note that the duration of the sweep in the simulation is not representative of the experimental conditions: the entire simulation represents ∼1 μs, while the sweeping period in the experiment was ∼1 ms. The potential at the upper boundary was set to 0, which corresponded to the plasma potential. The potential of the surrounding divertor tiles (blue in figure 1) was fixed at −3 kTe, which was an approximate value of the floating potential.

Figure 1.

Figure 1. Schematic view of the simulation geometry for the divertor LP (red) embedded in the divertor tile (blue). The probe has toroidal length Lp, radial width Wp and protrudes Hp above the tile surface.

Standard image High-resolution image

Since our model involved only a small volume of plasma in the vicinity of the probe head, it could not capture long range effects which the probe may have induced in the plasma. In particular, when the probe had large positive bias (with respect to the floating potential), it may have drained electrons from the adjacent flux tube, which would result in a reduction of the electron probe current (so-called flux tube charging) [15, 16]. In order to avoid this effect, we did not consider electron current in this part of the IV in further analysis.

2.2. Benchmarking of 2D and 3D simulations

The first series of simulations represented a scan in the angle of incidence of the magnetic field (ψ in figure 1) for computationally affordable plasma conditions. The angle was scanned in the range from 10° down to 1.25°, where the lower bound is representative of conditions in the COMPASS outer divertor, where the angle of incidence is typically between 3° (far SOL) and ∼0.5° (strike point). The choice of plasma parameters for this initial scan was not motivated by a specific discharge in COMPASS but by the desire to compare both 2D and 3D simulations, while capturing the scales which were expected to be important for probe current collection.

Initially, it was assumed that the driving mechanism of the sheath expansion effect was the polarisation drift in the magnetic pre-sheath (as described in [17] for the case of flush-mounted probes), which was a consequence of a strong gradient of E field in this region. As such, the magnitude of the drift was assumed to scale with the thickness of this pre-sheath (which in turn scales with rLi [14]). The sheath expansion effect was expected to be more pronounced when the majority of the probe body was submerged in the magnetic pre-sheath, therefore we selected Ti = Te = 40 eV, which is representative of plasma conditions in the vicinity of the outer strike point [18]. The density, on the other hand, was expected to have only minor impact on the sheath expansion as it influences the E field only in the Debye sheath via the Debye length ${\lambda }_{\text{D}}\sim {n}_{\text{e}}^{-1/2}$. As such, it was chosen to significantly reduce the density to speed up the calculations. Note that results presented in section 2.5 have contradicted this initial assumption and confirmed that in the particular case of this probe geometry and plasma conditions, the role of Debye sheath and plasma density is significant, which is consistent with the view presented in [19].

The cell size in a PIC grid is required to be smaller than λD, so a low density scenario (with large λD) allows to cover the simulated volume by a smaller number of cells. Therefore, the density ne = 1017 m−3 was selected for the initial set of simulations, while the plasma density in the vicinity of the outer strike point in COMPASS is in the range of 1018 m−3.

As mentioned before, the aim of the first series of simulations was to compare the results of a simplified 2D simulation with the full 3D simulation as it was essential to cross-check whether the two codes agree in an identical simulation geometry. This was not just an exercise in algorithm benchmarking since SPICE3 calculates electric field in all three dimensions, which enables e.g. E × B drift in the yz plane (which is not included in the 2D simulation). In order to achieve such a comparison, a dedicated 3D simulation was performed with Wp equal to Lx , which resulted in a simulation of an infinitely wide probe (just like in the 2D simulation). The resulting ion current (normalized to its value at Φ = −3) is shown in figure 2(A) for ψ = 10°. A very good agreement for the infinite width of the probe is achieved between 2D and 3D simulations, confirming that both codes are well compatible. However, as expected for the 3D finite width of the probe, there is a notable difference with the 2D simulation, which can be explained by the presence of magnetic pre-sheath in x direction (between the probes of the array in the poloidal direction). Regarding the electron current (figure 2(B)), all cases match (same slope), with the infinite probe having a greater current only due to its greater collection area.

Figure 2.

Figure 2. Comparison of ion (A) and electron (B) current in 2D and 3D simulations for discrete width of the probe and 2D-like continuous geometry for ψ = 10° and the scan in the width of the 2D simulations for ψ = 1.25° (C) using either reduced or realistic value of μ = mi/me.

Standard image High-resolution image

The scan was extended by further simulations in order to assess the effect of the mass ratio μ = mi/me. In order to accelerate the simulations, it is common to employ a reduced value of μ = 200. This setup generally should not affect the potential profile in the sheath nor the ion current. In some of the previous simulations, it was however observed that it can affect the electron probe current. Simulations with realistic μ = 3670 (corresponding to deuterium plasma) were therefore performed. Fortunately, in the case of this particular setup the role of reduced μ is negligible both on ion (figure 2(A)) and electron (figure 2(B)) currents. Nevertheless, we will employ simulations with realistic μ for the analysis of the electron current whenever possible.

Another important check is the length of the simulation in the toroidal direction (y axis). At grazing angle of incidence ψ the probe casts a long magnetic shadow and if the simulation is not sufficiently large, the probe may start shadowing itself. As shown in figure 2(C), the ion current is insensitive to the length for Ly ⩾ 1.5 Lshadow, where Lshadow = Hp/tan ψ is the length of the magnetic shadow cast by the probe. All the following simulations were designed to fulfil this criteria.

2.3. Parameters influencing the sheath expansion

2.3.1. Density, temperature and B field incidence

The variation of the ion current with the angle of incidence and density for both 2D and 3D simulations is shown in figure 3. The presence of the sheath expansion effect is clearly manifested by a slope on the ion current for Vp < Vfloat and is most pronounced for cases with low density and low angle of incidence. There is a notable difference between 2D and 3D simulations—the 3D simulations in general exhibit stronger sheath expansion (as it was already observed in section 2.2). Note that the variation of ion current is not limited to bias below the floating potential: for Φ > −3 the ion current is strongly reduced. In such case, Φ is higher than the potential of neighbouring tiles (which are set to −3 kTe in the simulations) and it is the sheath expansion above these tiles which is responsible for this effect. The associated polarization drift is pushing ions onto the surface of these tiles, preventing them from reaching the probe. When performing analysis of the total probe current obtained in experiment, the reduction of the ion current produces the same effect as if the electron current was increased—therefore acting against the reduction of electron current due to e.g. flux tube charging. As such, the reduced ion current contributes to the formation of the local minimum of the Te,fit(Vcut,upper) dependence (which is discussed in section 3.1).

Figure 3.

Figure 3. Ion current obtained in 3D (full lines) and 2D (dashed lines) simulations for varying density and angle of B field incidence ψ. Vertical dashed green lines denote Vp = −3 kTe and allow determination of Δi.

Standard image High-resolution image

The values of the ion saturation current obtained in these simulations can be compared with its theoretical value calculated as

Equation (3)

where Ti is the ion temperature, ns.e. is the electron density at the sheath entrance, mi the ion mass and Aeff is the effective probe collecting area for ions, which was assumed in previous works to be equal to 2.8 mm2, based on the geometrical cross-section of the probe [9] (the real probes have rounded edges, and as such slightly reduced cross-section). Note that in the simulations the nominal value of Isat is assumed to be the value of ion current at Φ = −3, which is the approximate location of the floating potential. The variation of the ion saturation current (figure 4(A)) shows a sensitivity both on angle and density. The variation with angle is, however, reduced at small angles and, especially in case of 3D simulations, the absolute value is close to Isat,theor (assuming Te = Ti). These observations also confirm that the value of Aeff assumed in previous work is correct.

Figure 4.

Figure 4. Variation of the ion saturation current Isat+ (A) and sheath expansion parameter Δi (B) with angle of incidence ψ and density for 2D and 3D simulations.

Standard image High-resolution image

As a further step to characterise the sheath expansion, we introduce its normalised magnitude Δi as

Equation (4)

where $\bar{\frac{\mathrm{d}{I}_{\text{i}}}{\mathrm{d}{\Phi}}}$ is the slope of the ion current for Φ < −3 (obtained using linear fit of the current in this bias range, thus assuming β = 1 in equation (2)). This parameter essentially describes how large is the relative increment of the ion current per 1 kTe when the bias decreases below the floating potential. Thus, it presents a convenient normalisation of the parameter αi introduced in equation (2), which can be expressed as:

Equation (5)

Figure 4(B) shows the variation of Δi in the performed simulations. Despite original expectations, the variation with plasma density is significant. However, the difference in the magnitude of Δi between 2D and 3D simulations can be reconciled using a fixed constant pre-factor 1.65, allowing us to employ 2D simulations to study the variation of Δi at high densities.

In addition to the simulations with varying density and angle of incidence, a scan in Te was executed for ne = 1 × 1018 m−3 and ψ = 3°. The resulting scaling (shown in figure 8(A)) is well characterised by a square root dependence.

2.3.2.  B field strength

COMPASS operates with a range of toroidal magnetic field BT of 0.9–1.5 T (at the magnetic axis) and so it was important to assess the influence of BT variation on the sheath expansion. A series of simulations for ne = 5 × 1017 m−3 and ψ = 2.5° was performed for equivalent values of BT at the location of the outer strike point (1.0–1.6 T). The results showed that the effect on Δi was negligible—less than 5% of change with respect to the value at BT = 1.0 T. This is probably caused by the narrow range of BT, which does not modify important scale orderings between the size of the probe, rLi and λD.

2.3.3. Probe geometry

So far all the simulations targeted the geometry of the LPs in the combined probe array. For completeness, we have extended the study to include geometry of the swept probe array, which was installed in COMPASS already during its operation in Culham [20] and was retained after re-installation in Prague. Originally, the probes in the swept array had an approximately circular cross-sections protruding 2.0 mm above the divertor tile with a probe length equal to 7.0 mm and probe width 3.0 mm. It was suspected that after many years of operation, their geometry may have been affected by the repeated impact of plasma particles, reducing their height. This was confirmed by a post-mortem analysis using an optical 3D microscope (Leica DVM6, Leica Microsystems, Switzerland), revealing a range of probe heights between 0.5 mm (in the vicinity of the strike points) and 1.75 mm at the HFS, as shown in figure 5(A).

Figure 5.

Figure 5. Probe height measurements by the 3D optical microscope (A) and effective collecting area Aeff calculated using the PIC simulations for the LPs of the swept probe array (B).

Standard image High-resolution image

In order to assess whether results obtained in section 2 are applicable also to these probes, an additional set of 2D simulations was performed for circular probes extending 2.0 mm, 1.5 mm and 0.5 mm above the divertor tile respectively for realistic plasma conditions (Te = 40 eV, ne = 5 × 1018 m−3), as shown in figure 6. The results for these plasma conditions show a strong insensitivity to the particular shape of the probe for fixed Hp = 1.5 mm. However, Δi is sensitive to the probe height, which was expected. In addition, the absolute values of the ion current show that Aeff of the swept probe array is 7.15, 4.68 and 2.49 mm2, respectively (for Hp = 2.0, 1.5 and 0.5 mm), including the factor 1.4 derived from the comparison between 2D and 3D simulations in section 2.3.1. A linear interpolation was used to derive collecting areas for each LP, as shown in figure 5(B). Note that these collecting areas are for most of the probes lower than the value 6.52 mm2 used in the experimental analysis [21], which has a direct impact on the deduced ion current density.

Figure 6.

Figure 6. Sheath expansion of LPs in the combined probe array and the swept probe array with original and eroded geometries.

Standard image High-resolution image

2.4. Correction for non-ambipolar conditions

So far we have assumed that the potential of the neighbouring divertor tiles is close to the floating potential—meaning that the SOL currents passing through the tiles can be neglected. However, such currents are routinely observed in COMPASS [22] and it is common for the floating potential of divertor LPs to be distinctly different from the machine ground (which is the potential of the divertor tiles). Since the sheath expansion of the probe is in competition with the sheath expansion of the neighbouring tiles, modification of tile potential could produce a visible effect on Δi. Additional 2D simulations for selected subset of plasma parameters were performed, in which the potential of the tiles VT was scanned. Note that simulations with VT > Vfloat can suffer from formation of an artificial sheath near the injection plane when large fraction of the impacting electron current is absorbed by the tile. For this reason the scan was limited to VTVfloat + 1 kTe, where the magnitude of such artificial sheath was insignificant (below 0.1 kTe).

Figure 7 summarises the results. In general, when the potential of the tiles is above the floating potential (as measured by the probe), Δi is reduced with respect to the ambipolar case. This is mainly due to an increase of ion current to the probe, which is used as normalisation parameter in equation (4). In the first order approximation, this trend can be characterised by a linear function:

Equation (6)

Figure 7.

Figure 7. Correction of Δi for the case of nonambipolar divertor tiles.

Standard image High-resolution image

In case of a divertor tile which is ∼1 kTe above the local floating potential (common case in the vicinity of the outer strike point), Δi is reduced by $\sim 20\%$.

2.5. Semi-empirical scaling of sheath expansion

Combining all the simulations corresponding to the geometry of LPs in the combined probe array together, we derived an empirical scaling for Δi:

Equation (7)

This scaling provides a decent fit to the data obtained in simulations, as demonstrated in figure 8(B) (only simulations with VT = −3 kTe shown here). It predicts an increase of sheath expansion magnitude for small angles of incidence, however the overall magnitude of Δi is suppressed for high density and/or low temperature.

Figure 8.

Figure 8. Dependence of Δi on Te (A) and the proposed semi-empirical scaling of Δi (B). Green point indicates conditions relevant to divertor probes in COMPASS, which are studied in section 3.

Standard image High-resolution image

Note that unlike in earlier works on this topic (e.g. [4, 19]) scaling parameters and their functional dependencies in equation (7) were not guided by simplified models of the sheath but derived solely empirically, using the output of simulations. This approach allowed us to avoid possible biases due to imperfections of such simplified models and was largely enabled thanks to the advancements of high performance computing, which allowed us to perform series of simulations for realistic plasma conditions. Nevertheless e.g. the scaling with angle of incidence ∼ sin−1/3ψ is reminiscent of the dependence sin−1/2ψ, which appeared also in model for sheath expansion parameter for the flush-mounted probe by Bergmann [4], which was confirmed by Gunn [17]. More detailed discussion on this parameter can be found in [7]. The square root dependence on Te is a surprising result since this means that αi is practically independent on Te (since ${\alpha }_{\text{i}}\sim 1/\sqrt{{T}_{\text{e}}}$). The derived scaling is far from being completely generic, for example the pre-factor 0.0041 was derived for the particular shape of the probe used in COMPASS. It was outside the scope of this work to characterise the influence of all possible probe shapes which are used in tokamaks.

The scaling of Δi with Te and ne is somewhat unfortunate, because it depends on quantities which should be obtained by the fit of IV characteristic itself. A solution of this problem can be the use of tangent estimate method, which is described in section 3.2.

2.6. Effect of sheath expansion on fitted temperature

Equation (7) suggests that Δi should be rather small for realistic plasma conditions and B field orientation in the COMPASS divertor (⩽0.1 per 1 kTe). In order to verify whether the sheath expansion is small enough to be completely neglected in the analysis of swept IV characteristics, it was important to assess how does its value influence the final value of Te obtained by a fit to the total probe current (consisting of contributions of both ions and electrons). This is achieved by fitting the probe current for Vp < Vfloat +1 kTe using the three-parameter fit. Unlike the analysis of the experimental data, in case of simulations the deviation of Te,fit can be easily quantified, since the temperature of electrons which are injected in the simulation Te,orig is known. This comparison is presented in figure 9 and confirms the initial findings by Matthews [3]—the stronger sheath expansion (larger Δi), the large over-estimation of temperature. The dependence can be characterised by a linear dependence and suggests that, for realistic plasma conditions, the expected over-estimation is ⩽25%.

Figure 9.

Figure 9. Dependence of fitted temperature using three-parameter fit on Δi.

Standard image High-resolution image

2.7. Conclusion of PIC simulations

Dedicated PIC simulations have demonstrated that for the particular geometry of the divertor LPs and the plasma conditions relevant to the COMPASS divertor, the sheath expansion effects should not have large influence on the analysis of experimental data. While there appears to be a systematic difference in the ion current obtained in 2D and 3D simulations, this variation can be reconciled by a simple constant pre-factor, meaning that relatively inexpensive 2D simulations can be ran instead of more computationally costly 3D runs. In addition, in case of the combined probe array, for realistic angles and plasma parameters the effective ion collecting area Aeff matches quite remarkably the value assumed in experimental analysis (i.e. 2.8 mm2), although this was previously deduced using only a simple geometrical consideration. For probes in the swept probe array, the Aeff used for experimental analysis appears to be overestimated by 50% or more, depending on the degree of probe erosion.

3. Fitting of experimental data

Dedicated experiments at COMPASS, in which swept voltage was applied to a selected LP in the combined probe array, were performed to verify the predictions of sheath expansion modelling presented in the previous section. The sweeping frequency was set to 1 kHz and the voltage was varied from −320 V up to +50 V using a triangular waveform. This unusually large range of applied voltage allowed to investigate the effects of sheath expansion and also the impact of the available voltage range on the quality of the fit, as will be shown in the following subsections.

3.1. Minimum Te method with stationary bootstrap

We have developed a fully automatised routine to analyse measurements of the swept LPs by adopting the minimum Te method [23, 24] and the stationary bootstrap [25], which is described in detail in this section. The workflow of data analysis is illustrated here for a single Ohmic low confinement discharge which is run at the beginning of each experimental day on COMPASS (Ip = −180 kA, BT = −1.15 T, ne = 4 × 1019 m−3). Based on previous observations, the area of interest was the vicinity of the outer strike point, where largest Te and smallest angle of incidence were expected (which should result in largest sheath expansion effects, as shown in the previous section). Therefore, probe #LPA36 (R = 510 mm, located 33 mm outside the outer strike point) was used. The measurements during the discharge flat-top were analysed using the three- and four-parameter fits (assuming β = 1 in equation (2)).

For each IV characteristic, a series of fits was performed with a varying upper boundary (Vcut,high) of the fitted voltage range: from ${V}_{\text{float}}+\frac{1}{4}\enspace {\text{kT}}_{\text{e,}\;\text{tangent}}$ up to ${V}_{\text{float}}+\frac{3}{2}\enspace {\text{kT}}_{\text{e,}\;\text{tangent}}$, where Te,tangent was an estimate of Te obtained by the tangent estimate method described in section 3.2. From this series of fits, the one with the minimum resulting Te was selected, as illustrated in figure 10. This method [23] allows to maximise the precision of the fit (by including some data at Vp > Vfloat) and it avoids over-estimation of Te due to distortion of electron current in the magnetised plasmas [15]. However, due to fluctuations of the probe current, some of the fits may under-estimate the real temperature and by choosing the minimum Te one could introduce a systematic bias into the analysis. In order to mitigate this effect and also to achieve realistic errorbars of the measurement, the stationary bootstrap method [25] was incorporated into the analysis toolchain. This method is based on the construction of a large number of virtual datasets based on random portions of the original experimental data (in our case each data point consist of a pair of current and voltage measurements), fitting these datasets and obtaining the distribution of resulting Te, Isat and Vfloat. If these distributions have Gaussian shape, the precision of the fit can be approximated as the standard deviation of the distribution and the final value as its mean.

Figure 10.

Figure 10. Example IV characteristic (A) with the dependence of fitted temperature on the upper boundary of voltage range used for the three-parameter fit (B).

Standard image High-resolution image

In the stationary bootstrap method, the new datasets are constructed by choosing blocks of data at random location in the original dataset. The length of the block is drafted from an exponential distribution with the decay parameter being equal to the autocorrelation time. This time, as deduced from the measurements of ion current at negative probe voltage, is taken as ∼5 μs. New blocks are selected until the length of the new virtual dataset is equal to the length of the original dataset. Since the blocks are selected randomly, some data points may be selected multiple times, while other data points may be completely missing. This allows to assess whether some outlying data points (e.g. due to impacting filaments) are affecting the resulting Te (especially when the minimum Te method is applied). Typically, ∼100 datasets are generated for each IV characteristic, while ∼50 fits are needed for the minimum Te method. In total, ∼5000 fits are required to obtain a reliable result from a single IV characteristic.

Figure 10 shows an example of IV and the scan for minimum Te using a three-parameter fit. As can be seen, sheath expansion effect is not pronounced, which is in agreement with PIC modelling for this range of plasma parameters. The semi-empirical formula introduced in equation (7) predicts Δi = 0.07 per kTe (Isat = 60 mA corresponds to density ne = 2 × 1018 m−3, angle of incidence ψ ∼ 1.5°). Note that no binning of the experimental data was applied for the analysis, instead all available data points were used for the fit. Binning is sometimes used to speed up fit convergence and for visual inspection of the fidelity of the fit, however, it can introduce bias in the data in situations where fluctuations of the probe current are of non-Gaussian nature, which is a common observation with probe current [26]. In such a case, the mean value is not a good representation of the underlying distribution of measurements.

The distributions of Te, Isat, Vfloat and the upper boundary of the fit Vcut,max are shown in figure 11. Except for Vcut,max all of them have approximately Gaussian distribution of values, which allows to use their mean and standard deviation as representations of the most probable value and confidence interval. The distribution of the upper voltage boundary of the fit shows multiple peaks, which is caused by the presence of probe current fluctuations in the vicinity of the upper boundary (forming multiple local minima in figure 10(B)). The scan in number of samples constructed for bootstrapping shows that estimation of mean values and standard deviations is possible already for 100 samples, however at least 1000 samples is required for analysis of the exact shape of the distributions. Figure 12 demonstrates correlations between the fitted parameters. There appears to be a significant correlation between the obtained Te and Isat, which is not surprising since ${I}_{\text{sat}}\sim \sqrt{{T}_{\text{e}}}$. An important result is the weak dependence only of the fitted Te on the choice of the upper voltage boundary for the fit. Note that the distribution of Vfloat in figure 11 does exhibit a bimodal distribution, however the scatter of the values is generally quite small.

Figure 11.

Figure 11. Histograms of obtained Te, Isat, Vfloat and Vcut,max using the stationary bootstrap method for varying number of samples. Fitting was performed for #20474, t = 1124 ms, probe #LPA36.

Standard image High-resolution image
Figure 12.

Figure 12. 2D histograms of Te, Isat, Vfloat and the upper cut off voltage, as calculated using the stationary bootstrap method for #20474, t = 1124 ms, probe #LPA36.

Standard image High-resolution image

3.2. Tangent method

For the purpose of further analysis of swept probes, we introduce a new method of Te estimation, which uses only data in the vicinity of the floating potential, which we call the tangent method. The method works as follows:

  • (a)  
    The IV data points are sorted according to voltage and then smoothed using running mean filter. The smoothing window size Nwindow is calculated as Nwindow = Sw N100, where Sw is typically equal to 0.4 and N100 is the number of data points with −100 $< {V}_{\text{p}}< $ 0 V. The optimum value of Sw depends on the level and character of noise of current measurements and as such needs to be tailored to a specific data acquisition system.
  • (b)  
    An estimate of Vfloat is found by iterating through the IV characteristic and looking for voltage where current of smoothed IV characteristic crosses 0. In order to make this procedure robust, the algorithm looks for interception of zero from left and right side independently. The estimate of Vfloat is then obtained as an average of these two values. This helps to mitigate problems in case of distorted IVs with multiple crossings of zero.
  • (c)  
    A tangent to the IV characteristic is found at Vp = Vfloat. The slope is obtained by linear fit of the smoothed IV characteristic in the interval of Vfloat ± 5 V. Note that the slope S of this tangent is equal to −Isat/Te (when sheath expansion effects are neglected).
  • (d)  
    The tangent is extended towards negative bias as Itangent. For each datapoint of the smoothed IV (Ij , Vj ) having Vj < Vfloat, the following ratio is calculated:
    Equation (8)
    Such point satisfies the following equation
    Equation (9)
    This equation is transcendental but a solution for Te can be expressed using the Lambert function W [28]:
    Equation (10)
    In practice the calculation of Lambert function is performed using a fast algorithm developed by Fukushima [29]. For example, for a data point with corresponding tangent current being twice the value of the smoothed measured current (r = 2) the solution reads Te = −(Vj Vfloat)/1.59. The final Te,tangent estimate is obtained as a mean value of all obtained values of Te with 0.6 < r < 3. Extending the procedure towards higher values of r would bring risk of influence of the sheath-expansion effects (which are small in the vicinity of the floating potential since these scale as ∼(VVfloat)).

The method is illustrated in figure 13(A). The magenta dashed line shows the tangent to the smoothed IV, the vertical dashed red line shows obtained value of Vj . The reliability of this method was tested on the dataset of swept LPs in the combined probe array as shown in figure 13(B) (including measurements with large sweeping range). While there is noticeable scatter in the data, the method works reasonably well for a wide range of temperatures, especially for Te < 40 eV. As the tangent method does not require any iterative fitting, it is notably faster than the method described in section 3.1 and appears to be a good candidate for real-time applications. Note that the tangent method is distinctly different from the double probe technique, which can also make use of the first derivative of the IV characteristic [2].

Figure 13.

Figure 13. Example of the application of tangent estimate method using data from swept LP LPA36 of combined probe array in #20474 (A). Comparison of the estimated temperature with result of the three-parameter fit (B). Colours in (B) indicate different discharges.

Standard image High-resolution image

It is instructive to use synthetic probe data to inspect this method. One important question is its sensitivity to sheath expansion—since this method does not use current measurements at large negative voltage, it should be insensitive to Δi. We applied two approaches when dealing with synthetic probe data: (i) we generated smooth probe current for a pre-set of constant values of Te, Isat, Vfloat and αi (labelled synthetic) and (ii) we used fast measurements of the combined probe array and employed time-dependent values of Te, Isat, Vfloat (using data from probe triplet #35 in #20474). The value of Δi was calculated using equation (7), where ψ was artificially manipulated to obtain a scan in sheath expansion. In both cases, the probe current was generated for a synthetic voltage waveform, which had the same range and frequency of sweeping as in the experiment.

The resulting dependence of Te,tangent on Δi is shown in figure 14(A). Both approaches yield very similar trends, while using the time-dependent values of plasma parameters resulted in significantly larger and asymmetric errorbars (here represented by 10% and 90% quantiles of dataset measured over a flat top of #20474). The method is reasonably stable for Δi < 0.2. As such, it provides an independent mean to cross-check the validity of three- or four-parameter fits (which is presented in next sections).

Figure 14.

Figure 14. Sensitivity of the tangent method on sheath expansion (A), error in determination of the slope (B) and error in determination of Vfloat (C).

Standard image High-resolution image

Another important check is to evaluate the stability of the resulting Te,tangent with respect to the small errors in determination of the slope of the tangent S = dIp/dVp(Vp = Vfloat) and Vfloat, which are shown in figures 14(B) and (C). In both cases, we have introduced a random error in the process of its determination (with Gaussian probability function) and observed the effect on Te,tangent. This to some extent mimics the real-world case where the experimental data are acquired with some level of noise and fluctuations. In case of the slope S, the error in its determination gradually increases the scatter of Te,tangent. However, it does not introduce a significant bias in the results. Error in 10% of determination of S yields $\sim 20\%$ scatter in Te,tangent. In case of Vfloat the situation is more complicated, and having δVfloat/Te > 0.2 leads to an over-estimation of the obtained temperature estimate. However, note that the scale of the scan is rather extreme compared to the properties of the experimental data. Fitted IV characteristics typically show match of Vfloat with precision of ∼1 V, which for Te ∼ 40 eV yields δVfloat/Te ∼ 0.05.

3.3. The role of voltage range

3.3.1. Scan in Vcut,low

The generous range of swept voltage applied to the probe allowed us to assess the role of the voltage range Vrange = −(VminVfloat)/Te on precision of fits of the probe data, especially with respect to the measured Te. This can be done easily by setting different limits of the lower voltage boundary Vcut,low in the analysis, thus excluding parts of the IV characteristics from the fitting procedure. Figure 15 shows an example of results obtained with both three-parameter and four-parameter fits from the IV characteristic presented in figure 10. It can be seen that for the largest available voltage range the results for three- and four-parameter fits are very similar. As the voltage range is reduced, the three-parameter fit shows good stability in terms of Te,fit for −320 $< \ {V}_{\text{cut,}\;\text{low}}\ < $ 200 V, which suggests that the fitting is valid. On the other hand, the four-parameter fit starts to yield lower and lower temperatures. For Vcut,low ⩾ −250 V (Vrange ∼ 5 kTe) the resulting temperatures are in strong disagreement with the values obtained with maximum available voltage range. This bias towards low temperatures can be explained by an interference between the two fitting parameters: Te and αi. In vicinity of Vfloat, the exponential term in equation (2) can be linearized and for β ∼ 1 the fit becomes ill-defined:

Equation (11)

Figure 15.

Figure 15. Fitted temperature (A), sheath expansion magnitude αi (B) and voltage range Vrange (C) as a function of the lower boundary of voltage range for three-parameter fit, four-parameter fit, two step four-parameter fit and two step four-parameter fit using predicted value of αi in equation (7). The parameter Vrange in (C) was obtained using either Te resulting from fits (solid lines) or estimated using the tangent method (dashed line).

Standard image High-resolution image

Both the second and the third term scales linearly with VpVfloat, it is impossible to reliably determine both Te and αi. This problem is demonstrated in figure 15(B), which shows the dependence of αi on Vcut,low (and thus on available voltage range). In principle, its value should be constant (and close to zero in this respective case), however, there is a clear trend of increased magnitude of αi with reduced voltage range.

As a possible way to circumvent this problem, the four-parameter fit can be applied in two steps: (i) the slope of the linear term is determined using data obtained at the lowest probe bias and (ii) the extra current due to sheath expansion is subtracted from the IV characteristic, which can be then treated using the three-parameter fit. We have attempted to employ this two-step technique (green points in figure 15) but it did not yield more reliable results. On the contrary, this method is prone to fluctuations in the probe current (which can hamper determination of the linear term), which results in larger scatter of the resulting parameters. These fluctuations can be caused e.g. by MHD modes with characteristic frequency close to the probe sweeping frequency.

The two-step fitting technique can be significantly improved if the coefficient αi is not determined from the slope of the IV characteristic but if it is calculated using equations (7) and (5). This is demonstrated by the yellow points in figure 15. For the full range of swept voltage, the prediction of the semi-empirical law yields larger αi than obtained by the four-parameter fit, which leads to a certain under-estimation of Te in this situation (26.5 eV vs 37 eV given by the three-parameter fit). However, the agreement improves when the voltage range is reduced and the match is generally better than for a regular four-parameter fit. Since the prediction of αi is based on the tangent method, it is rather insensitive to Vrange (as shown in figure 15(B)).

In case of the three-parameter fit, the reduction of obtained temperatures is much less pronounced, however, having Vcut,low > −200 V (Vrange ∼ 4 kTe) also leads to a reduction of the fitted temperature. This effect may be caused by the IV characteristic not strictly following an exponential dependence, as predicted e.g. by Rozhanski [27].

A practical counter-measure against using the four-parameter fit in case of insufficient voltage range is to rely on the value of Vrange and apply a four-parameter fit only if Vrange is above a certain threshold [7]. However, the above analysis shows that this technique will fail if one uses Te originating from the four-parameter fit itself. When the range is not sufficient, Te will be under-estimated, leading to an over-estimation of Vrange. The same problem, albeit less pronounced, can also occur for the three-parameter fit (e.g. Vrange based on Te from three-parameter fit in figure 15(C) never drops below 4). The tangent method allows to circumvent this problem, since it is not using the part of IV characteristic most affected by the sheath expansion.

3.3.2. Criteria for application of the fits

The boundaries for safe application of IV characteristic fitting presented in figure 15 may be to some extent accidental, demonstrated only for a single particular case. In order to achieve more robust criteria, we have repeated the scan in Vrange over a larger data set measured during diverted phases of selected L-mode and ELM-free H-mode discharges (each data point represents a value obtained using a 10 ms time window). In each case we noted at which value of Vrange does Te drop by 20 and 50 percent, respectively, with respect to the value obtained at full voltage range. The resulting histograms are presented in figure 16. Note that we have used Te,tangent to determine Vrange, as it provides an independent reference point which only weakly depends on the actual voltage range. Vertical dashed lines indicate median values of the distributions. It can be seen that indeed the four-parameter fit requires at least 4 kTe of voltage range to achieve a systematic error better than 20%. If the voltage range is less then 3 kTe, then the resulting temperature is under-estimated by factor of two or more. For the three-parameter fit, these boundaries are 2.9 and 1.4 kTe, respectively. Figure 16(C) demonstrates that when prediction of the sheath expansion based on PIC simulations (equation (7)) is used in the two step technique, the reliability of the fitting procedure is, to large extent, recovered: Vrange of at least 2.5 kTe is needed to achieve precision of the fit of 20% and for Vrange smaller than 1.6 kTe the resulting temperature is under-estimated by at least 50%.

Figure 16.

Figure 16. Histograms of Vrange (calculated from the tangent method) at which Te drops by 20 (red) or 50 (blue) percent with respect to fit at full voltage range for three (A), four (B) parameter fit, two step four-parameter fit using predicted magnitude of sheath expansion based on equation (7) (C) and the tangent method (D). Vertical dashed lines indicate median values of the distributions.

Standard image High-resolution image

Finally, figure 16(D) shows the results of the same analysis for the tangent method, which requires Vrange ⩾ 1.7 kTe and ⩾1.1 kTe to achieve precision of 20% and 50% respectively. The good stability of this method with respect to reduced Vrange is caused by the fact that it uses data points in a close vicinity of the floating potential.

The usefulness of Vrange derived using the tangent method is further demonstrated when αi obtained by four-parameter fits are compared to the semi-empirical scaling formula equation (7). In particular, the sheath expansion parameter αi obtained using the four-parameter fit can be confronted with the prediction of Δi presented in equation (7) (here the parameters needed for prediction of Δi were obtained using the tangent method). Such comparison (as shown in figure 17(A)) demonstrates that the semi-empirical scaling serves as good estimate of experimental Δi for most of the measurements. However, measurements with Vrange < 4 kTe show strong deviation from this trend. The same group of points exhibits disagreement between four-parameter fit and the tangent method, as shown in figure 17(B). This indicates that for small voltage range, the four-parameter fit yields incorrect results—over-estimated αi and under-estimated Te. This is the result of interference between fitting parameters described earlier.

Figure 17.

Figure 17. Comparison of parameter Δi obtained by four-parameter fit (y axis) with predictions of the semi-empirical scaling formula equation (7) (x axis) (A) and cross-check between temperature Te,fit obtained by the four-parameter fit and the Te,tangent estimate obtained by the tangent method (B). Colours of the data points represent Vrange calculated using Te,tangent.

Standard image High-resolution image

3.4. Comparison with ball-pen probe measurements

During regular operation, the combined probe array determines Te using the difference in floating potentials of LP and ball-pen probe measurements as

Equation (12)

where α is assumed to have a fixed value of 1.4 [9]. This measurement remained available even in the experiments reported here, which provided a straightforward way to cross-check the results obtained using swept probes.

Figure 18 shows results of such a comparison for our dataset of swept probe measurements. Data points with Vrange > 3 kTe are plotted in colour, measurements with insufficient Vrange are plotted as grey points (with grey outline).

Figure 18.

Figure 18. Comparison of swept LPs in the combined array with LP + BPP measurements at the same radial location. Colours indicate different discharges, stars and diamonds represent outliers with Vfloat > 0 V.

Standard image High-resolution image

The agreement between the two methods is satisfactory in most cases. Peak temperatures at the outer target in the range of 50 eV are consistently observed by both methods (see an example of such measurements in figure  19(c)), which also agrees with older measurements performed at COMPASS-D [20]. This indicates that the LFS SOL is in the sheath-limited regime with almost constant Te along the magnetic field lines. Further analysis of the outliers revealed that when the floating potential was positive—either in L-mode (discharges #12416 and #12401, represented by stars) or following a transition into an ELM-free H-mode (#20504 and #20505, represented by diamonds), the LP + BPP method yields very low (figure 19 (b)) or even negative (figure 19 (a)) temperatures, while the temperatures derived from the swept LP remain above 20 eV. The detailed cause of this behaviour is not understood at this moment and will be a subject of future investigations. Note that it is often observed that swept LPs fail to deliver temperatures below a certain threshold (∼5 eV) even if other diagnostics suggest colder plasmas [30, 31]. Unfortunately, we have not yet devised a method to determine this threshold value of Te.

4. Conclusions

We have used first principle PIC modelling and analysis of experimental results from COMPASS tokamak to determine conditions under which it is appropriate to employ three or four parameter fit for analysis of the swept LP measurements. Series of simulations allowed us to determine a semi-empirical predictive formula for the magnitude of sheath expansion for given Te, ne, B field inclination in the particular geometry of the LPs of COMPASS.

Figure 19.

Figure 19. Comparison of Te measurements performed by swept probe with LP + BPP method for ELM-free H-mode where Vfloat becomes positive (a), Ohmic discharge with positive Vfloat (b) and Ohmic discharge with negative Vfloat  (c).

Standard image High-resolution image

In case of four-parameter fit a good agreement between modelling and experimental results was achieved, indicating that at least 4.4 kTe of voltage range are needed to obtain reliable results (with systematic error less than 20%).

Most measurements obtained by sweeping the voltage on LPs of the combined probe array at the outer target can be safely analysed using three-parameter fit, since the sheath expansion effects are small (Δi < 10%) and have only limited effect on the fitted Te ($\leqslant 25\%$ over-estimation). Interestingly, this method also requires a minimal voltage range to yield solid results, however the threshold value of Vrange is significantly smaller than in case of the four-parameter fit (about 2.7 kTe).

A novel tangent method was developed to estimate Te in a manner which is insensitive to the sheath expansion effects and can provide an unbiased estimate of the actual voltage range. Coincidentally, this method allows fast computation of a Te estimate, which can have potential applications in real-time control systems.

A comparison of Te measurements with results of the ball-pen probe method confirmed existence of high Te ∼ 50 eV in the vicinity of the outer target, which suggests that the SOL plasma is in the sheath-limited regime with only small temperature gradient along the magnetic field lines.

Acknowledgments

This work was supported by the Czech Science Foundation Projects GA22-03950S and GA20-28161S and by MEYS Projects #LM2018117 and CZ.02.1.01/0.0/0.0/16_017/0002248.

Please wait… references are loading.
10.1088/1741-4326/ac8011