Brought to you by:
Paper

Electric field in APTD in nitrogen determined by EFISH, FNS/SPS ratio, α-fitting and electrical equivalent circuit model

, , , , , , , and

Published 15 June 2023 © 2023 IOP Publishing Ltd
, , Citation Martina Mrkvičková et al 2023 Plasma Sources Sci. Technol. 32 065009 DOI 10.1088/1361-6595/acd6de

0963-0252/32/6/065009

Abstract

We investigate the electric field development in weak microseconds-lasting atmospheric pressure Townsend discharge operated in a barrier discharge arrangement in pure nitrogen. The electric field is determined using four different methods: laser-aided electric field induced second harmonics (EFISH), optical emission-based first negative/second positive systems (FNS/SPSs of molecular nitrogen) intensity ratio, electrical equivalent circuit approach and via determination of the Townsend first coefficient $\alpha(E/N)$ from the spatial optical emission profile. The resulting values of the electric field obtained by these respective methods, regardless of the differences in absolute values, lie within a reasonable range. The limitations and advantages of all methods are discussed in detail for the investigated discharge. The EFISH measurements are supported by re-computation of the effective interaction-path of the laser using an electrostatic model. The FNS/SPS method provides systematically higher values compared to other methods. We discuss in detail the potential origin of this discrepancy, as this method is at the limit of its applicability due to the low E/N values and we also consider the impossibility of full verification of the underlying assumptions. The focused discussion addresses best-practice issues and identifies possible future steps to improve each of the four methods under given conditions.

Export citation and abstract BibTeX RIS

1. Introduction

Atmospheric pressure cold plasmas are intensively studied phenomena and are broadly applied in technology in many fields, e.g., in surface treatment, flow control or plasma medicine and agriculture [1, 2]. Cold plasmas are generated in various gases and discharge types (dielectric barrier, corona, plasma jets or nanosecond pulsed) depending on the desired utilization. One of the most important plasma parameters, which determines the discharge dynamics and initiated chemistry in these plasmas, is the electric field [3]. Different experimental methods have been developed and are applied for electric field determination, yet with such a variety of discharge types, gases and other conditions, there is usually only one method matching the requirements of the studied discharge system.

Typically, for streamer or fast-pulsed plasmas in nitrogen containing gas mixtures, the optical emission spectroscopy-based method of the intensity ratio of molecular nitrogen spectral bands is used [46]. The ratio of the optical emission intensities of the first negative system (FNS) of N$_2^+$ and second positive system (SPS) of N2 is sensitive to the reduced electric field (FNS/SPS = f($E/N$)). When combined with a detection device of very high spatial and temporal resolution and triggering precision at far sub-nanosecond scales (fast PMT [7], time-correlated single photon counting [6], and picosecond cameras [8, 9]), the method delivers unique results consistent with numerical simulations [6]. Nevertheless, the method is limited by uncertainties in the fundamental kinetic data (cross-sections, collisional quenching, Boltzmann equation solver) as it is based on a kinetic model and it is not a direct method. The exact influence of the input uncertainties on the reduced electric field strength determination by this method was quantified by Obrusník and Bílek et al [1012].

Laser-aided diagnostics offer the EFISH method, the generation of the second harmonics induced by the local electric field [13, 14]. Since its recent introduction into low-temperature plasma physics by Dogariu et al [13], the method has found a broad field of applications in fast pulsed discharges in different gases and under different conditions [1418]. This in-situ direct method for electric field measurement offers high sensitivity and accuracy, yet, as with every laser-based method, it is limited in spatial resolution by the interaction length of the laser beam, its profile and its possible intrusive interaction with close surfaces [19]. Atmospheric pressure filamentary discharges of diameter of a few tens of micrometers are not easily measurable with a hundreds-of-micrometers laser beam profile. The temporal resolution of the method is limited by the laser pulse length, typically of hundreds of picoseconds or even femtoseconds, and by the accuracy of the laser synchronization with the discharge. Recently, also, a sub-nanosecond resolution was achieved by a nanosecond pulsed laser [20].

In the case of diffuse barrier discharges in the so-called Townsend regime [2123], the electrical equivalent circuit model can be applied for the average electric field determination in the discharge gap [21, 2427], too. In the atmospheric pressure Townsend discharge (APTD) in volume barrier discharge arrangement in nitrogen, the electric field is slightly ($\lt$10%) disturbed by a space charge, and therefore can be approximated as $E = U_\textrm{g}/d$, where $U_\textrm{g}$ is the voltage in the gas gap and d is the dielectrics spacing [23]. The $U_\textrm{g}$ can be obtained from the theory of an equivalent circuit, using the current and voltage measurements. Additionally, the discharge capacitances have to be determined theoretically or experimentally with satisfactory precision. The capacitance determination may be a bottleneck on the way to electric field determination in these diffuse discharges, as the electrostatic edge effects or partial discharging may play a non-negligible role [21, 24, 25, 28, 29].

Furthermore, the electron density in Townsend discharges rises exponentially from the cathode to the anode. As the optical emission of the ionizing plasma is directly proportional to the electron density, one may, under given conditions, fit the spatial emission profile by an exponential function and receive the Townsend ionization coefficient α from the fitting function. The $\alpha(E/N)$ is an unambiguous function of the electric field and thus, the electric field value can be estimated from the discharge emission [4, 30].

Additionally, the authors of the recent work [31] investigated $\mathrm{N_2}({\mathrm A}^3 \Sigma^+_\textrm{u}, v)$ metastables in APTD in pure nitrogen by phase-resolved spectroscopy and a state-resolved kinetic model. They estimated the density of the two lowest metastable $\mathrm{N_2}({\mathrm A}^3 \Sigma^+_\textrm{u}, v = 0,1)$ states to be between 1012 and 1014 cm−3 depending on the applied voltage. The knowledge of these values is crucial for the electric field determination using optical methods (α-fitting and the intensity ratio method) since these metastable states affect the SPS intensity in the post-discharge phase by producing the SPS emission through pooling reactions.

In this paper, we study the electric field development in barrier discharge APTD using all four of the above introduced methods. Such an investigation gives a unique opportunity to directly compare the results of the determined electric field. Based on the direct comparison, the discussion of method uncertainties becomes quantitative where all factors, also the previously neglected ones, have to be taken into account. The limitations and advantages of all methods are discussed in detail for the investigated discharge. The method of determining the electric field from the ratio of FNS and SPS band intensities includes an extension using the SPS bands originating from $v\geqslant0$ vibronic levels [12]. The EFISH method takes into account the recently presented results of Chng et al [32, 33]. It is also used to obtain the spatially resolved electric field development, and subsequent computation of positive ion densities in the APTD further addresses the uncertainty of the method. The possible effect of nearby dielectric surface charging is taken into account, and its contribution to the method uncertainty is quantified.

Due to the usual conventions in the barrier discharge community, the electric field is presented in Townsend units as a reduced electric field $E/N$ where 1 Td = 10−21 Vm2. For atmospheric pressure gas and 300 K, the number density N is given by the Loschmidt constant. In this manuscript, the short term 'electric field' is used for the reduced electric field E/N.

The paper is divided into the following sections. Section 2 describes the experimental setup of the APTD and the diagnostic techniques used for electric field measurement. Section 3 reports on methods used for electric field determination and presents the results. Finally, in section 4 we discuss in great detail the obtained electric field values and consider possible reasons responsible for the discrepancies between the four different methods.

2. Experimental setup

The scheme of the investigated discharge setup is shown in figure 1. A homogeneous dielectric barrier discharge was ignited between two plane-to-plane rectangular electrodes ($14\times15$ mm) deposited on the outer surfaces of two microscope slides (glass, width 1.1 mm, relative permittivity $\varepsilon_r = $ 4.7) separated by a gas gap of d = 1 mm width. The electrode setup was housed inside a closed plastic chamber, with a continuous flow of pure nitrogen at atmospheric pressure. The gas inlet and outlet were positioned at the vertical level of the inter-electrode gas gap, perpendicular to the two quartz windows (oriented at the Brewster angle) protruding from the chamber, enabling discharge optical observation and laser beam passage.

Figure 1.

Figure 1. Schematic of the discharge configuration. (a) Side view, (b) top view.

Standard image High-resolution image

The discharge was powered by sinusoidal AC high voltage with a frequency of 11 kHz and a typical amplitude of 7 kV, additionally modulated by a 500 Hz on-off modulation signal with a duty cycle of 45%. Specifically, a burst of 10 discharge cycles with a single-cycle period of 91 µs was followed by 1090 µs off-time. The voltage and current waveforms of a full modulation period are shown in figure 2.

Figure 2.

Figure 2. The applied voltage and current waveforms of the discharge during one modulation period. Magenta, orange and green vertical lines represent the exposure windows of the ICCD frames that will be displayed and discussed later in figure 4.

Standard image High-resolution image

An oscilloscope, a Keysight DSO-S204A with a Tektronix P6015A high-voltage probe and a Tektronix CT2 current probe were used to obtain the electrical parameters of the discharge necessary for the simplest electrical equivalent circuit method of electric field determination, as well as for continuous monitoring of the discharge behavior during all of the experiments.

Optical observations of the discharge were performed with an intensified CCD camera PIMAX-3 (Princeton Instruments). A macro lens made of fused silica was used to magnify the picture onto the camera chip. The gate width of the ICCD camera was set to 2 µs and the time step of the imaging was also set to 2 µs, i.e. there was no overlap in the exposure times of the images. The camera recordings were synchronized with electrical measurements.

The experimental setup for the electric field-induced second harmonic generation method (EFISH) is shown in figure 3. An Nd:YAG picosecond pulsed laser EKSPLA PL2231-50-TRAIN produced laser pulses of duration 30 ps, wavelength 1064 nm, typical pulse energy 4 mJ and repetition rate of 50 Hz. The laser beam was focused by a spherical lens (f = 50 cm) to the discharge reactor, resulting in a beam waist of 80 µm in the focal point and a Rayleigh length 20 mm. The EFISH signal was separated from the pumping beam using two dichroic mirrors, a dispersive prism, diaphragm, and narrow bandpass filter, and was detected by the PMT Photek PMT210. The energy of the pumping beam was monitored simultaneously by the photodiode Thorlabs DET10A/M and power meter Vega (Ophir).

Figure 3.

Figure 3. Optical setup for the EFISH measurement. 1a, 1b, lenses to focus and collimate the laser beam (f = 50 cm); 2a, 2b, dichroic mirrors (reflecting 532 nm); 3, dispersive prism; 4, iris diaphragm; 5, narrow bandpass filter 532 nm.

Standard image High-resolution image

Optical emission spectra were obtained with both an ICCD-based (phase resolution spectra) and non-intensified CCD-based spectrometers (time-averaged spectra), FHR-1000 (Horiba) and Shamrock 750 (Andor), with CCD DU940P-BU2 and ICCD DH340T-18 F-03 from Andor. Additionally, a fast photomultiplier (PMT) in combination with appropriate bandpass filters was used to obtain the intensity waveforms of selected vibronic bands. A PMT (Hamamatsu H10721) and an iHR-320 spectrometer Jobin–Yvon equipped with an Andor iStar ICCD DH740i-18U-03 camera were used to register the phase-resolved optical emission waveforms and spectra collected from the whole discharge volume.

3. Methods of electric field determination and results

The operation of the discharge in burst mode assured very stable performance without any long-term temperature drift and great repeatability between individual bursts. At the beginning of the modulation on-phase, while the voltage amplitude was still increasing to the set level, the discharge was dominated by strong filaments with high current peaks. The discharge filaments were occurring in stable positions in both time and space, as was confirmed by the electric and ICCD camera measurements. After increasing the voltage amplitude to a set value, approximately after two periods, the discharge switched to homogeneous mode, covering the whole area of the electrodes. The objective of this work, the electric field, was investigated at the end of the modulation on-phase, i.e. during seventh to ninth discharge period, where the discharge was stable, repeatable and free of any superposed current pulses. Figure 4(a) shows an example of one period of measured voltage and current during this phase of modulation burst. Well pronounced current humps are symmetrical for both polarities. Figure 4(b) presents the relative scale ICCD camera frames of the homogeneous discharge during positive and negative half-periods and filamentary discharge at the very beginning of the burst. The green, orange and magenta borders of the ICCD frames correspond to the exposure time windows of these frames highlighted in figures 2 and 4(a).

Figure 4.

Figure 4. (a) The measured voltage and current waveforms of the discharge during one period of homogeneous discharge operation. (b) Exemplary ICCD frames of the discharge appearance in homogeneous and filamentary modes. The respective exposure windows are color-coded in figures 2 and 4(a). The green and orange subfigures show the discharge in positive and negative half-periods, respectively.

Standard image High-resolution image

3.1. Equivalent circuit

In the case of APTD in symmetric volume barrier discharge, we are interested in the electric field in the gas gap, which is directly connected to the reduced electric field as $E/N = U_\mathrm{g}/(d \cdot N)$, where N is the nitrogen density. The voltage in gas gap $U_\mathrm{g}$ cannot be measured directly, however it can be calculated from the external voltage V(t) and current waveform I(t) measurements using the electrical equivalent circuit theory [25], which gives us this formula:

Equation (1)

where Q(t) is the transferred charge obtained as $Q(t) = \int_0^t I(t^{^{\prime}}) \textrm{d}t^{^{\prime}}$ (in this experiment a symmetric charge transfer is expected, with the integration constant Q0 set to meet this assumption, see [24, 25]) and $C_\mathrm{d}$ is the capacitance of the dielectric barriers, which can be determined from a QV plot. This method of electric field determination was employed for four different applied voltages: 6, 6.5, 7 and 7.5 kV. The QV plots were constructed from one average period of discharge operation for each investigated applied voltage. The QV plots are shown in figure 5(a) and it can be seen that they are almost ideal parallelograms, which allows the straightforward determination of the $C_\mathrm{d}$ as linear fits of the respective active phases of the QV plot (and the capacitance of the discharge cell $C_\mathrm{cell}$ as linear fits of the respective passive phases). The fits and resulting values of the capacitances are shown in figure 5(a) for the applied voltage amplitude of 7 kV.

Figure 5.

Figure 5. (a) The QV plots for four investigated applied voltages. Linear fits (black lines) and resulting capacitances are shown for $V_\mathrm{amp} = 7.0$ kV. (b) Measured and calculated electrical parameters for the applied voltage amplitude of 7.0 kV.

Standard image High-resolution image

The resulting overview of the electrical parameters, both measured (I and V) and calculated by this method (the displacement current $I_\mathrm{cap} = C_\mathrm{cell}\frac{\mathrm{d} V(t)}{\mathrm{d}t } $, the discharge current $I_R = \frac{1}{1-\frac{C_\mathrm{cell}}{C_\mathrm{d}}} (I-I_\mathrm{cap})$ and the voltage in the gas gap $U_\mathrm{g}$) [25], can be seen in figure 5(b). This example is shown for the applied voltage amplitude of 7 kV. The maximum $U_\mathrm{g}$ is calculated to be approximately half of the maximum V. It is also ahead of the V due to the charge deposited during the previous half-period, which adds up to V. After discharge ignition, the voltage remains almost constant (even slightly decreasing) for the duration of the discharge, even though the applied voltage keeps increasing.

Figure 6 shows the calculated electric fields during a single period of discharge operation for four different applied voltages. Since the investigated electric field is directly proportional to the gas gap voltage, compared to the sine wave applied voltage, the electric field is disturbed by the presence of the discharge. The maximum electric field is reached during the plateau section of its development, corresponding to the active phase of the discharge. The maximum electric field is at almost the same value for the three higher voltages (134 Td) and slightly lower for the lowest voltage (125 Td). The electric field is symmetric for positive and negative half-periods. The error bands (determined from the signal noise and uncertainty of the capacitance fits) are low for all voltages. The influence of different capacitance determination methods is discussed in the next section.

Figure 6.

Figure 6. Electric fields calculated using the simplest electrical equivalent circuit method for four applied voltage amplitudes.

Standard image High-resolution image

3.2. Townsend coefficient

This method relies on the Townsend avalanche theory, which tells us that the electron density in the Townsend discharges rises exponentially from the cathode to the anode due to a gas ionization process where free electrons are accelerated by an electric field, collide with gas molecules, and consequently produce secondary electrons. Here, we assume no losses of electrons due to recombination, which seems reasonable in such a weakly ionized gas of APTD. The electron density n grows exponentially toward the anode and in any place inside the gas gap can be described by the well known formula [34]

Equation (2)

where n0 is the initial electron density, x is a distance from the cathode surface toward the anode and $\alpha (E/N)$ is the first Townsend ionization coefficient, giving the number of free electrons generated per unit length by one electron moving from cathode to anode.

When the excited states are populated exclusively by the electron impact excitation, the relative intensity of the optical emission of the plasma is directly proportional to the electron density. This can be advantageously used to obtain the Townsend ionization coefficient directly from an experiment without the need for any absolute calibration. The spatially resolved discharge emissions were recorded by the ICCD camera with a gate width of 2 µs and temporal step of 2 µs, scanning through the whole discharge period in synchronization with the electrical measurements, resulting in a series of frames, such as the examples shown in figure 4. Each frame was then processed to obtain α corresponding to a given time, as represented in figure 7(a). First, the most active portion of the discharge was selected from the entire frame to minimize any impact of slight non-uniformity of the discharge in the horizontal direction caused by the gas flow. The pixel intensities of each row were then averaged to suppress the noise, providing a 1D spatial relative emission intensity profile. The section of the profile corresponding to the gas gap area was fitted, and the Townsend coefficient α was obtained from the fit. For better fit reliability, the actual fitting was done in logarithmic scale by a linear function.

Figure 7.

Figure 7. (a) The ICCD frame processing steps; the cropped portion of the entire frame, the averaged intensity of each row of pixels and the intensity profile with inter-electrode region fitted by exponential function (the apparent emission outside of this region is due to the transparency of the glass electrodes). The letters A and C denote anode and cathode, respectively. (b) The dependency of $\alpha(E/N)$ on reduced electric field [35].

Standard image High-resolution image

The first Townsend coefficient is an unambiguous function of the reduced electric field. Several databases provide us with this dependency for nitrogen gas. For the conversion, we have used the Biagi database with the BOLSIG+ solver in the LXCat webpage [3537]. The obtained dependency, shown in figure 7(b), was then interpolated and directly used to calculate the reduced electric field from the α coefficient.

The results of this procedure are shown in figure 8 for four different applied voltages (the same amplitudes as in section 3.1). There is a clearly visible increase in the electric field during both polarities, temporally coinciding with the active discharge phases. The electric field evolution is the same for both polarities. The maximum reduced electric field value increases slightly and gradually with applied voltage amplitude, from 158 Td to 163 Td. The minimum electric field does not seem to drop below 110 Td, with only one data-point exception. The error bars, calculated based on both the uncertainty of the fits and the uncertainty of the reproducibility, are also the largest for the lowest $E/N$ values. In addition to the two previous uncertainties, a non-negligible uncertainty springs from the uncertainty of the α($E/N$) relation, which is different when considering different cross-sections databases or experimental swarm parameters. This uncertainty is dominant for the highest $E/N$ values and reaches about 5% for 160 Td.

Figure 8.

Figure 8. Electric fields calculated using the Townsend coefficient method for four applied voltage amplitudes.

Standard image High-resolution image

Apparently, as the light intensity significantly decreases between the active phases of the Townsend discharge, the results of the $E/N$ data become less and less accurate. Moreover, in the post-discharge, the emission may originate from mechanisms other than electron-impact excitation, such as, for example, the action of metastable species. This is further elaborated in the discussion in section 4.

3.3. Electric field induced second harmonics—EFISH

3.3.1. EFISH methodology.

The principle of the EFISH method lies in the focusing of a strong laser on the investigated electric field area, which leads to the generation of the second harmonic radiation parallel to the primary laser beam.

Following [32] and [33], the power of EFISH signal for a focused probe beam is given by

Equation (3)

Equation (4)

Equation (5)

where $\alpha^{(3)}$ is the third-order nonlinear hyperpolarizability of the gas, N is the gas number density, $P_0^{(\omega)}$ is the power of the probe beam (1064 nm), zR is the Rayleigh length of the focused laser beam, $E_{\mathrm{ext}}(z)$ is the externally applied electric field distribution along the beam propagation axis z, E0 is the electric field strength at z = 0, $E^{^{\prime}}_{\mathrm{ext}}(z^{^{\prime}})$ is a dimensionless electric field profile, and $\Delta k = [2k^{(\omega)}-k^{(2\omega)}]$ is the wave-vector mismatch.

Assuming that neither gas composition nor the laser shape and the electric field distribution along the z-axis do not change during all of the measurements, nor calibration, the majority of formula (3) can be reduced to one calibration constant A,

Equation (6)

For the calibration, we can apply the external electric voltage V to the same electrode setup, but with the voltage low enough not to cause a discharge. Under these conditions, the Laplacian electric field in the interelectrode gap can be calculated as

Equation (7)

where $d_{\mathrm{diel}}$ is the sum of widths of both dielectric barriers, $d_{\mathrm{gap}}$ is the distance between them, $\varepsilon_{r,\mathrm{diel}}$ is the relative permittivity of the dielectrics and $E_{\mathrm{surf}}$ is the electric field caused by the surface charge remaining on the dielectric barriers. Plotting the quadratic dependence of $P^{(2\omega)}$ on V in conditions without discharge allows us to obtain the parameter A needed for the calibration of the subsequent measurements.

In our experiment, we detected the EFISH signal with a PMT and recorded it with an oscilloscope. All waveforms were acquired from 16 laser shots. For the evaluation, the value $P_p^{(2\omega)} = p \cdot P^{(2\omega)}$ was taken as the depth of the (negative) signal peak. Similarly, the probe beam power $P_0^{(\omega)}$ was monitored by the photodiode and its peak height $P_{0q}^{(\omega)} = q \cdot P_0^{(\omega)}$ was used for evaluation. Constants p and q are included in the calibration constant A. The shape of both the PMT and photodiode signal remained the same during all the measurements. In all of the following figures, the term 'EFISH signal' denotes the signal already divided by the measured probe power,

Equation (8)

The calibration was performed before the main measurement. As the first step, we slid the dielectric surface with a conductive mesh to remove all the surface charge possibly remaining from the previous discharges. Because of this, the term $E_{\mathrm{surf}}$ should become zero. We then applied the AC voltage of the sub-breakdown value and measured the dependence of the EFISH signal on the voltage phase, i.e., on the instantaneous voltage value. Figure 9 shows the resulting dependencies for various voltage amplitudes. For amplitudes lower than 5 kV, all the measurements are in good agreement, and the dependence is purely quadratic. However, the peaks of all the parabolic dependencies are slightly shifted from the zero voltage. This shift indicates the presence of some unknown parasitic signal. Fortunately, its value was constant for all the measurements. Therefore, we suggested extending the calibration formula (6) by the term B describing the parasitic signal from the constant electric field:

Equation (9)

Figure 9.

Figure 9. Dependence of the EFISH signal on the instantaneous applied voltage for various amplitudes of applied voltage. For amplitudes up to 5 kV, the electric field was Laplacian and all the curves follow the same parabolic shape. For 6 kV and above, the discharge was ignited, resulting in the disturbance of the electric field by the surface charge.

Standard image High-resolution image

Both parameters A and B were used to calibrate the subsequent measurements. For the voltage amplitude 6 kV, the discharge ignited, resulting in the deposition of the surface charge on the dielectric surface during the active discharge phases and the hysteretic behavior of the measured EFISH signal during the applied voltage period.

To check the validity of this calibration method, the electrode setup was replaced by plan-parallel conductive plates of length 14 mm generating a Laplacian electric field of the same shape, without being affected by the presence of the dielectric. The shift of parabolic dependencies remained unchanged. The resulting calibration parameters $A_\mathrm{p}$ and $B_\mathrm{p}$ led to an increase in the calibrated values of the measured electric field by 10% compared to the first calibration. This deviation is expected to be caused by the uncertainty of the dimensions of the electrodes, dielectrics and the inter-electrode gap, or the uncertainty of the dielectric permittivity. Uncertainty of the EFISH method is further studied in the next section.

3.3.2. Uncertainty quantification of the EFISH measurement.

Recent publications by Chng et al [32, 33] offered a detailed insight into the origin of the EFISH signal and provided expressions that can be directly used to obtain the theoretical value of the measured signal. Experimental measurements were used to validate the theory (albeit the agreement was only qualitative) and to confirm that the EFISH signal changes dramatically and non-monotonically with the lateral distribution of the electric field (i.e., electric field intensity along the beam path). Specifically, the authors illustrated that even for parallel-plate electrodes of varying length, the measured EFISH response shows an oscillatory behavior.

In our work, most of the effects described in [32] should not apply because we used an identical geometry for EFISH calibration and EFISH measurement. However, we use these expressions to quantify the uncertainty in our EFISH measurements because even a subtle change in the interaction path of the beam can change the result rather dramatically.

There are two reasons why the interaction path could change in our geometry:

  • (i)  
    Edge effects: in the bulk of the plasma, it is reasonable to expect that the E vector is perfectly perpendicular to the electrode surface (as it would be in the calibration case without plasma). However, at the edges, the presence of the plasma may affect the intensity of the electric field near the edges of the electrodes, e.g., due to accumulated surface charge on the adjacent glass.
  • (ii)  
    Laser misalignment: a small change in the laser angle changes the interaction length.

To quantify the uncertainty caused by edge effects, we created three synthetic profiles of the electric field. These electric field profiles are shown in figure 10(a). First, the electric field is constant between the electrodes and zero everywhere else. In the second and third profiles, the electric field is allowed to decrease gradually (exponential decrease), reaching the half-maximum value at 0.1 mm and 0.4 mm from the electrode edge, respectively. Although the profile is synthetic, it is a reasonable analogy to the case when surface charge would accumulate on the glass adjacent to the electrode. The expression

Equation (10)

from equation (3) was integrated numerically in order to obtain the synthetic EFISH signal. Figure 10(b) compares the resulting EFISH signal for these three cases. It is apparent that when the electric field extends outwards to the electrodes only slightly, the effect on the EFISH signal is minuscule (less than 2% deviation). However, if the electric field is allowed to extend further beyond the electrode, the theoretical EFISH signal can change by around 16% for our electrode dimensions. This uncertainty, therefore, clearly contributes to the overall uncertainty of the EFISH measurement. Moreover, a similar situation could have arisen in the case of our second calibration, where the conductive plates might have dimensions a few percent different from the original setup.

Figure 10.

Figure 10. Three electric field profiles considered in the uncertainty assessment (a). The theoretical EFISH signal $\Lambda^{^{\prime}} /z_R$ for these three profiles depending on electrode length 2L, (b).

Standard image High-resolution image

Second, the effect of laser misalignment effectively increases the interaction length of the beam with the plasma. However, even if we consider that the angle of the laser would change by 10 between the calibration experiment and the plasma measurement, the error caused is of the order of 2%. This information is contained within plot 10(b) because the misalignment of the laser effectively increases the interaction length of the beam. From this, we conclude that the method is not overly sensitive to spatial misalignment of the laser beam.

3.3.3. Parasitic signals.

In this section, we will discuss the asymmetry of the measured signal-voltage dependencies shown in figure 9. The axes of symmetry of all the measured datasets were shifted from the zero voltage by the same value B, corresponding to the electric field caused by the applied voltage of 250 V. Due to the quadratic dependence of the signal on the electric field, such a parasitic signal was invisible in the case of zero voltage, but it caused up to a 10% change of the signal in the case of the highest electric field measured for 6 kV of applied voltage. Specifically, it caused a 10% increase for the positive voltage half-period and a 10% decrease for the negative one. If not treated properly by the term B in the calibration formula (9), that might cause a significant systematic error in the obtained results.

Next, we will try to investigate possible sources of such a parasitic signal. First, it is hypothesized that the EFISH signal could be influenced by structural holders placed in close proximity to the plasma cell (5 mm from it). The schema of the holder is depicted in figure 11. If these holders accumulate sufficient floating potentials by capturing free charges escaping from the active discharge region, they could create a parasitic electric field. Such an electric field would be parallel to the electric field in between the electrodes in the positive half of the period (thereby increasing the EFISH signal) and anti-parallel in the negative half of the period (thereby decreasing the signal).

Figure 11.

Figure 11. Schema of the electrode setup concerning the plastic holder causing the parasitic EFISH signal.

Standard image High-resolution image

To quantify the effect and estimate the magnitude of this parasitic electric field, we produced synthetic electric field profiles shown in figure 12(a). It is apparent that we added two peaks of electric field at 2 cm distance from the plasma cell axis, i.e., at the position at which the dielectric holders are placed. In one case, the electric field of the holders is parallel to the electric field in the plasma, in the other case, it is antiparallel. In both cases, the intensity of the parasitic electric field was set to 10% of the electric field intensity inside the plasma cell. Figure 12(b) confirms that a parasitic electric field of such magnitude would certainly influence the EFISH signal in a measurable way. In particular, the 25% increase in the signal in the positive half-period and 25% decrease in the negative half-period of the applied voltage can be expected in the presented case.

Figure 12.

Figure 12. (a) Electric field profiles approximating the situation when dielectric holders near the plasma accumulate substantial surface charge. (b) The impact of the charged holders onto the theoretically measured EFISH signal $\Lambda^{^{\prime}} /z_R$ in the positive and negative half-periods, respectively.

Standard image High-resolution image

We could conclude that the asymmetry presented in figure 9 might be caused by such a parasitic field. To demonstrate this experimentally, we covered the structural holders with a conductive layer and connected it to the ground. However, the parasitic signal remained nearly unchanged, indicating that the problem is elsewhere.

The second candidate and the final culprit was a parasitic signal originating in the input and output windows of the reactor chamber. These windows can generate a significant amount of signal, especially in the case when they are situated close to the laser focal point. To prove this, we performed three nearly identical calibration measurements of the EFISH signal dependence on instantaneous applied voltage in sub-breakdown conditions in ambient air. The first measurement was performed in the closed chamber; in the second and the third case, we removed the input and output windows of the reactor. Figure 13 shows the perfect disappearance of the signal asymmetry in the case without any contact of the laser beam with the reactor windows.

Figure 13.

Figure 13. Elimination of the parasitic signal by the removal of the input and output windows of the reactor chamber.

Standard image High-resolution image

This section should show the breadth of possibilities as to where parasitic signals can arise. Fortunately, the effect of a constant parasitic signal can be fully eliminated by the proposed calibration method.

3.3.4. Results.

The EFISH results of the temporal evolution of the electric field in the middle of the inter-electrode gap during one AC period are shown in figure 14. The measurements were performed for applied voltage amplitudes ranging from 1 kV to 8 kV. For voltages of up to 5 kV, there was no breakdown and the measured signal followed the shape of the quadrate of the applied voltage. For applied voltage amplitudes from 6 kV to 8 kV, the breakdown occurred and the discharge burned diffusely, as can be seen in the current waveforms. The evolution of the EFISH signal during the applied voltage period reveals a strong asymmetry for the positive and the negative half-periods, supposedly caused by the parasitic signal. This asymmetry fully disappeared after calibration of the signal by formula (9), resulting in values of the electric field that are the same for both half-periods. The maximal electric field of 150 Td was reached for 6 kV, while higher applied voltages led to an earlier breakdown at slightly lower values of the electric field.

Figure 14.

Figure 14. Temporal evolution of the electric field during one discharge period in the middle of the inter-electrode gap, determined by EFISH, together with waveforms of the applied voltage, current and measured signal. The 10% error-bars are not shown in the graph.

Standard image High-resolution image

The uncertainty of the measured electric field is assumed to be 10%, considering the uncertainty of the determination of parameters A and B from the fit, the uncertainty of the dimensions of the electrode setup, the dielectric permittivity and the resulting mismatch between two different calibration methods.

Only one of the presented methods, EFISH, was able to obtain a spatially resolved electric field in the gap between the dielectric barriers. We varied the vertical positions of the discharge reactor with a step of 0.1 mm. For each position, the EFISH signal development during one discharge period was recorded. The calibration procedure described in section 3.3.1 was performed for each position individually. The resulting spatiotemporal map of the reduced electric field is shown in figure 15. The measurements reveal a slight inhomogeneity of the electric field during the active discharge phase. The absolute value of the electric field measured in the positive half-period of the applied voltage gradually increases when we scan the inter-electrode gap from the powered electrode (anode) to the grounded electrode (cathode), as is expected for Townsend discharge. The same trend is apparent also for the negative half-period. The increase up to 10% is just at the border of the APTD conditions, which assume just slight distortion (${\lt}10\%$) of the external electric field caused by the space charge [22, 23].

Figure 15.

Figure 15. Temporal development of the electric field as measured by the EFISH method in various positions in the inter-electrode gap. The subfigures show (a) the voltage and current waveforms of the investigated discharge period, with vertical lines indicating times displayed in subfigures (c) and (d); (b) temporal development of the electric field measured in fixed positions; spatial map of the electric field at chosen times in the positive (c) and negative (d) half-period; and (e) the full spatiotemporal development.

Standard image High-resolution image

3.3.5. Ion density.

Supposing that the spatial profile of the reduced electric field is correct, we have a possibility to estimate the $[\textrm{N}_4^+]$ density, which are dominant ions in the discharge and the increase in their density is responsible for the spatial variation of the electric field:

Equation (11)

where ε0 denotes the vacuum permittivity (in this case it approximates well the molecular nitrogen permittivity) and the $q_\textrm{e}$ denotes elementary charge. The electron density in the case of the APTD is several orders of magnitude lower than the ion density and therefore can be neglected. Thus, the ion density is given as:

Equation (12)

The most pronounced drop of the electric field from cathode to anode is observed during the active phase of the positive (8, 12, 16 µs) and negative (57, 61 µs) half periods, as shown in figures 15(c) and (d). This drop corresponds to the value of $\frac{E(x_2) - E(x_1)}{x_2 - x_1} \sim 10-20 $ Td/mm, which when converted into SI units gives approximately 0.25–0.5 GV m−2. Such a gradient of the electric field leads to an ion density of 1.5–3.0 × 1010 cm−3, which is in good agreement with the ion densities calculated in the APTD models of [22, 23].

3.4. Optical emission spectrometry

3.4.1. Methodology.

The FNS/SPS intensity ratio method is an alternative tool, based on the detection of UV emission in the 300–400 nm spectral range, and has been used to estimate $E/N$ since the 1970s [38]. The sensitivity of the method is based on the difference of energy thresholds and cross section profiles for electron-impact excitation and ionization for the N2(C$^3\Pi_\textrm{u}$) and N$_2^+$(B$^2\Sigma_\textrm{u}^+$) states (SPS and FNS upper states, respectively), which are excited from the ground electronic state N2(X$^1\Sigma_\textrm{g}^+$, v = 0). Recently, the method has been critically updated in [10, 11], and subsequently significantly extended in [12] by taking into account SPS band emissions from $v \geqslant 1$ levels of the N2(C$^3\Pi_\textrm{u}$) state. The originally used single FNS(0,0)/SPS(0,0) ratio can now be replaced by five independent ratios: FNS(0,0)/SPS(ξ), where $\xi\in$ {(0,3); (1,4); (2,5); (3,6); (4,7)}. The advantage of this extension lies in the more thorough use of spectrometric data and five different intensity ratios will be used for the determination of the electric field within this work.

However, under the present discharge setup ($E/N \lt$ 300 Td, repetitive frequency ∼10 kHz, microseconds lasting discharge), the basic conditions of validity and applicability of the FNS/SPS intensity ratio method must be carefully examined. The estimation of the reduced electric field is based on the conclusions of the work [12], where it is supposed that:

  • (i)  
    $T_\textrm{rot}$ and the gas temperature $T_\textrm{g}$ are not very different from room temperature, so the rate constants for the collisional quenching of the involved radiative states (determined at $T_\textrm{g}$ = 300 K) can be used.
  • (ii)  
    The SPS and FNS upper states excitation is governed by the direct electron-impact excitation/ionization from the v = 0 level of the $\mathrm{N_2}({\mathrm X}^1 \Sigma^+_\textrm{g})$ state.

In the case of the APTD, the $T_\textrm{rot}$ and $T_\textrm{gas}$ do not increase over 400 K, which is due to the low electron density (lower than 108 cm−3), and therefore condition (i) is fulfilled. Fulfillment of condition (ii) can be verified by studying the vibration distribution function (VDF) of $\mathrm{N_2}(\textrm{C}^3\Pi_\textrm{u},v)$ and $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v)$. As shown next, the VDF of $\mathrm{N_2}(\textrm{C}^3\Pi_\textrm{u},v)$ is apparently driven by direct electron impact excitation during the active phase of the discharge. To test the VDF of $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v)$ state is much more challenging due to the very low densities of $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v \gt 0)$ and unfortunately, this could not be done within this study. Thus, the validity of condition (ii) for FNS cannot be directly verified, and the use of the FNS/SPS intensity ratio method is therefore potentially subject to systematic error due to other processes that might affect the measured FNS intensity.

The dependence $R(E/N)$ relating the intensity ratio and reduced electric field is [12]:

Equation (13)

where $I_\textrm{FNS, SPS}$ are the measured intensities of the respective spectral bands and $\tau_{\mathrm{eff}}^{\mathrm{\scriptscriptstyle FNS, SPS}}$ are their effective lifetimes. The $g_\textrm{FNS, SPS}$ are factors denoting the ratio of the bandhead intensity and the integral band intensity of the respective FNS and SPS bands. These factors depend on the $T_\textrm{rot}$ and instrumental function parameter $\Delta_\textrm{1/2}$ (half-width of triangular instrumental function of a spectrometer, see [12]). In our case $\Delta_\textrm{1/2}$ = 0.175 nm and the $\frac{g_{\mathrm{\scriptscriptstyle SPS}}}{g_{\mathrm{\scriptscriptstyle FNS}}}$ reaches values of 0.75, 0.69, 0.64 for 300, 350 and 400 K, respectively. We used the value for 350 K, which was measured also in our conditions. In the specific case of the APTD, the steady state conditions for emission are fulfilled, since the largest changes of the intensity with time are at the initiation of the active discharge and these changes are in the order of microseconds:

Equation (14)

where effective lifetimes at atmospheric pressure nitrogen are $\tau_{\mathrm{eff}}^{\mathrm{\scriptscriptstyle FNS}}$ = ($0.052\pm0.010$) ns and $\tau_{\mathrm{eff}}^{\mathrm{\scriptscriptstyle SPS}}$ = ($3.15\pm0.65$) ns for zero vibrational levels of the FNS and SPS upper states; the values and their uncertainties are defined and thoroughly discussed in [12]. Therefore, the time derivative terms in equation (13) can be omitted, and formula (13) reduces simply to:

Equation (15)

The function ${R}\left(E/N\right)$ can be obtained either experimentally in a Townsend discharge [39] or using a theoretical ${R}\left(E/N\right)$ determination constructing a well-justified kinetic model and providing the ${R}\left(E/N\right)$ with its confidence band [1012]. In this section, we will use the latter, which enables us to determine the electric field, also with its confidence band.

3.4.2. Results.

The basic emission signatures obtained by the phase-averaged as well as phase-resolved emission spectroscopy are shown in figures 16 and 17. Figure 16(a) shows the emission spectra obtained at two selected spectral intervals and integrated during the AC cycle. The UV part of the spectrum is dominated by the vibronic bands of the SPS. The FNS(0,0) emission with the band head at 391 nm is very weak compared to the neighboring SPS bands (see insert in the figure) and indicates a low $E/N$ average [12, 31]. The NIR portion of the spectrum, as shown in figure 16(b), reveals a blend of two N2 systems, the first positive (FPS, $\mathrm{N_2}({\mathrm B}^3\Pi_\textrm{g}) \rightarrow \mathrm{N_2}({\mathrm A}^3\Sigma_\textrm{u}^+)$) and the Herman infrared (HIR, $\mathrm{N_2}({\mathrm C^{^{\prime\prime}}}^5 \Pi_\textrm{u}) \rightarrow \mathrm{N_2}({\mathrm A^{^{\prime}}}^5 \Sigma^+_\textrm{g})$). While the observed FPS emission is at atmospheric pressure caused mostly by the electron impact excitation, the HIR emission occurs due to the $\mathrm{N_2}({\mathrm A}^3 \Sigma^+_\textrm{u}, v) + \mathrm{N_2}({\mathrm A}^3 \Sigma^+_\textrm{u}, v^{^{\prime}})$ energy pooling and proves a certain density of $\mathrm{N_2}({\mathrm A}^3 \Sigma^+_\textrm{u}, v = 0, 1)$ metastable species [40].

Figure 16.

Figure 16. Typical emission spectra acquired in two selected wavelength intervals and integrated throughout the AC cycle, i.e., phase averaged. The UV part (a) of the spectrum is dominated by the vibronic bands of the N2-SPS. Weak emission of the N$_2^+$-FNS through the (0,0) band at 391 nm is shown in the insert. The NIR part of the spectrum (b) reveals a blend of the two systems, the N2-FPS and N2-HIR.

Standard image High-resolution image
Figure 17.

Figure 17. Characteristic variations in the intensities of selected SPS (a), (b), FNS (b) and NO (c) bands during the AC cycle, i.e., phase resolved spectra. The observed asymmetry in intensities between the positive and the negative half-period is most probably caused by the non-symmetric alignment of the focusing lenses.

Standard image High-resolution image

Figures 17(a) and (b) show the phase-resolved ICCD spectra captured during one AC cycle. The variations of the strongest SPS(0,0) band with the band head at 337 nm clearly indicate the time intervals in which the active discharge is present. This is also confirmed in figure 17(b), which captures the time variation of FNS(0,0) band intensity in the spectral range 389–393 nm. Note that the FNS(0,0) band intensity is lower compared to the SPS(3,6) band intensity and much lower compared to the SPS(2,5) band intensity. Consequently, the FNS(0,0) band is significantly overlapped by the tail of the SPS(2,5) band, which implies the need for careful correction when evaluating the ratios between the FNS(0,0) and SPS($v^{^{\prime}},v^{^{\prime\prime}}$) band intensities.

We also inspected ultraviolet radiation below 300 nm for the presence of NO-γ bands. Figure 17(c) shows the characteristic profile of the NO-γ(0,2) band, which also shows changes in intensity during the AC cycle, but never completely disappears. Persisting post-discharge emission of the NO-γ bands is maintained by long-lived metastable species $\mathrm{N_2}({\mathrm A}^3 \Sigma^+_\textrm{u}, v)$ through the resonant energy transfer process NO(X$^2\Pi_\textrm{r}$) + N2(${\mathrm A}^3 \Sigma^+_\textrm{u}$) $\rightarrow$ NO(A$^2\Sigma^+$) + N2(X$^1\Sigma_\textrm{g}^+$) [41].

Analysis of the intensities of SPS bands belonging to the sequences $\Delta v = 0$, −2 and −3 by means of synthetic models [12, 42] allows the obtaining of the vibrational distributions of the N2(C$^3\Pi_\textrm{u}$, $v = 0-$4) emitting states. The procedure is based on individual synthetic SPS band profiles forming the aforementioned SPS sequences, taking into account the instrumental function of the used spectrometer (given by the combination of the entrance slit and the 1200 G mm−1 diffraction grating, approximated by a triangular function). Next, the SPS(2,5) synthetic band profile was used to subtract the signal in the 390–393 nm spectral interval containing the FNS(0,0) band, thereby revealing the true FNS(0,0) intensity. This correction is quite important in this work, because the FNS intensity is relatively low (due to low $E/N$) and the FNS(0,0) band is always overlapped by the tail of the SPS(2,5) band formed by rotation lines (rotational quantum number $J \geqslant$ 19) belonging mainly to the R-branches [43]. After receiving the corrected FNS(0,0) peak intensity, we used the approach developed in our previous work [12] and estimated the $E/N$ values from several FNS(0,0)/SPS(vʹ,vʹʹ) ratios. In the case of emission spectra averaged over the discharge current pulse, the VDF normalized to v = 0 reads 1 : 0.195±0.007 : 0.045±0.005 : 0.0125±0.001 : 0.0025±0.0007 and is equal for both AC half-cycles ($V_\textrm{amp} \geqslant 6.5$ kV). As mentioned above, the shape of the VDF correlates to the mechanisms for the $\mathrm{N_2}(\textrm{C}^3\Pi_\textrm{u})$ population. Kinetic models of [31, 44] reporting on the VDF of the $\mathrm{N_2}(\textrm{C}^3\Pi_\textrm{u})$ state, caused by electron impact excitation at atmospheric pressure, give similar VDF ([31]—1 : 0.195 : 0.042 : 0.01 : 0.004; [44] at $T_\textrm{e} = 3$ eV—1 : 0.187 : 0.044 : 0.008 : 0.002) as the experimentally-obtained VDF, which confirms the $\mathrm{N_2}(\textrm{C}^3\Pi_\textrm{u})$ population dominantly driven by the electron impact excitation. This is valid during the active discharge phase of the APTD.

A very important insight is provided by the development of the VDF within one discharge period. Figure 18 shows the time evolution of the [N2(C$^3\Pi_\textrm{u}$, v = 1) ]/[N2(C$^3\Pi_\textrm{u}$, $v^{^{\prime}} = 0$)] and [N2(C$^3\Pi_\textrm{u}$, v = 2) ]/[N2(C$^3\Pi_\textrm{u}$, v = 0)] ratios (the index denotes the vibrational number). During the active discharge phase (0–20 µs), the SPS emission is caused only by electron impact excitation from the $\mathrm{N_2}({\mathrm X}^1 \Sigma^+_\textrm{g}, v = 0)$:

Equation (16)

Figure 18.

Figure 18. Characteristic SPS band ratios evaluated using the ICCD kinetic series obtained at $V_\textrm{amp} = 6.5$ and 7 kV. The green points in the figure demonstrate the profile of the FNS/SPS intensity ratio. The applicability interval of the method is denoted by the dashed red line rectangle.

Standard image High-resolution image

During the post-discharge phase, the N2(C$^3\Pi_\textrm{u}$, $v = 0-4$) states are predominantly populated by the pooling reaction from the metastable states,

Equation (17)

which affects the shape of the VDF and causes the increase in the [$\mathrm{N_2}(\textrm{C}^3\Pi_\textrm{u},v^{^{\prime}} = 1)$]/[$\mathrm{N_2}(\textrm{C}^3\Pi_\textrm{u},v^{^{\prime\prime}} = 0)$] ratio observed in $t \gt 15\,\mu$s. This increase is indicative of a growing population of N2($\textrm{A}^3\Sigma_\textrm{u}^+,v = 0$–1) metastables, as described in detail in [31]. The FNS/SPS intensity ratio, included in figure 18, illustrates the profile of the intensity ratio in the electron impact excitation-dominated region, where the FNS/SPS intensity ratio method can be applied.

The emission intensities of the FNS, SPS, and HIR bands were also monitored using a fast PMT in combination with appropriate interference filters, as shown in figure 19. It turns out that the FNS intensity waveform is non-zero only during the active discharge phase (see figure 19), while the SPS and HIR waveforms never reach zero level. This is consistent with the fact that the post-discharge emission in pure nitrogen atmospheric-pressure afterglow is sustained by $\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+)$ metastable species [42]. The remarkable phase shift of the peak of the HIR emission relative to the FNS and SPS maxima of about 15 µs confirms the initial build-up of metastable species during the active discharge phase. This initial build-up (controlled by electron-impact excitation) is replaced at the end of the discharge phase by a relaxation mechanism within the $\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+, v)$ vibrational manifold, which results in the formation of local post-discharge maxima in the $\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+,v = 0-1)$ populations [45]. The time interval between the local HIR maxima and the beginning of the next active discharge phase is suitable for a quantitative analysis of $\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+,v = 0$-1) metastables [31].

Figure 19.

Figure 19. Characteristic (a) voltage-current and (b) monochromatic photomultiplier waveforms obtained at $V_\textrm{amp} = 7$ kV. FNS and SPS waveforms were used to evaluate the FNS/SPS ratio.

Standard image High-resolution image

We evaluated the FNS/SPS ratio for all SPS bands belonging to the $\Delta v = -3$ sequence, see figures 20(a)–(e) and table 1. The green bands in figures 20(a)–(e) represent the uncertainty band in the reduced electric field determination from 80 to 500 Td. The band originates from the uncertainty in the kinetic data (cross sections, effective lifetimes, etc, see [11, 12]). The lower and upper bounds were fitted using the following formula:

Equation (18)

Figure 20.

Figure 20. Characteristic FNS/SPS ratios and corresponding $E/N$ intervals evaluated from FNS(0,0) and SPS($v,v+3$) bands using ICCD spectra obtained at $V_\textrm{amp} = 7$ and 7.5 kV.

Standard image High-resolution image

Table 1. The measured intensity ratios for the positive and negative half period of the discharge and the corresponding reduced electric field interval, for $V_\textrm{amp} = 7$ and 7.5 kV.

  7 kV  
 positive $E/N$ negative $E/N$
band R [Td] R [Td]
FNS(0,0)/SPS(0,3)0.0028(180; 300)0.0027(170; 300)
FNS(0,0)/SPS(1,4)0.0062(180; 320)0.0061(180; 320)
FNS(0,0)/SPS(2,5)0.022(180; 320)0.019(170; 300)
FNS(0,0)/SPS(3,6)0.095(170; 330)0.079(160; 310)
FNS(0,0)/SPS(4,7)0.54(150; 310)0.46(150; 290)
  7.5 kV  
 positive $E/N$ negative $E/N$
band R [Td] R [Td]
FNS(0,0)/SPS(0,3)0.0027(170; 300)0.0026(170; 290)
FNS(0,0)/SPS(1,4)0.0055(170; 310)0.0057(180; 310)
FNS(0,0)/SPS(2,5)0.022(180; 320)0.021(170; 310)
FNS(0,0)/SPS(3,6)0.087(160; 320)0.088(160; 330)
FNS(0,0)/SPS(4,7)0.49(150; 300)0.44(150; 290)

which was found to well reproduce the ${R}\left(E/N\right)$ dependency [39]. The reduced electric field determination based on the FNS/SPS ratios, as presented in figures 20(a)–(e), can be used only in the active discharge phase, where the emission of the FNS and the SPS is the strongest, and the SPS emission is still not affected by the energy pooling from metastable states.

The horizontal red dashed lines in figures 20(a)–(e) represent the measured values of intensity ratios and the corresponding blue vertical lines define the reduced electric field range obtained due to the uncertainty in the kinetic data (cross sections, effective lifetimes, etc). The measured values originate from the ICCD spectra integrated around the peak of the current pulse during a given AC half-cycle. The ranges of reduced electric field determined for external voltage $V_\textrm{amp} = 7.0$ and 7.5 kV are listed in table 1.

The intensity ratios obtained at 7 kV give slightly higher values at the positive half-period when compared to the negative half-period. However, the intensity ratios at 7.5 kV give comparable values for the negative and positive half periods. Nevertheless, we estimated the quite high uncertainty in the determination of the intensity ratios to be 10%, 10%, 15%, 15%, 25% for FNS(0,0)/SPS(ξ), with $\xi\in$ {(0,3); (1,4); (2,5); (3,6); (4,7)}, respectively. Intensity ratios for AC amplitudes of 7 and 7.5 kV are shown in figure 20(f) and prove that the ratios within experimental error depend neither on the amplitude nor on the instantaneous polarity of HV AC. Note that the emission spectra are integrated over the whole discharge gap and therefore the method in the current setup is not able to distinguish the spatial variations of the electric field. We conclude that the reduced electric field obtained by the FNS/SPS ratio method under the basic assumptions detailed in section 3.4.1 and evaluated from five independent pairs of FNS and SPS bands at the moment of peak discharge current is $240\pm90$ Td.

The temporal profile of the reduced electric field obtained from the FNS(0,0)/SPS(0,3) intensity ratio is shown in figure 21. To calculate the reduced electric field, we used single curve $R(E/N)$ from the center of the uncertainty interval of figure 20(a). The obtained values are systematically higher compared to the reduced electric field obtained by all other methods.

Figure 21.

Figure 21. Temporal profile of the reduced electric field obtained by the FNS(0,0)/SPS(0,3) intensity ratio method.

Standard image High-resolution image

4. Discussion

A comparison of all of the methods for the determination of the electric field is shown in figure 22. The best agreement is found between the EFISH and the equivalent circuit methods, their confidence bands overlap during the whole discharge period. The remaining two methods give systematically higher values. Moreover, they are not reliable at the post-discharge phases due to emissions defined by pooling reactions and no longer by electron impact excitation, which is the assumption for both methods.

Figure 22.

Figure 22. Comparison of the reduced electric field obtained by the proposed methods for the discharge with an amplitude of applied voltage of 7 kV.

Standard image High-resolution image

EFISH was the only method that provided a spatially refined electric field. Obviously, the intensity ratio method could also provide such information; however, the FNS signal is too low for such a spatial measurement, requiring the summation over the whole discharge gap. The equivalent circuit method is based on external voltage and current measurements, which are macroscopic parameters, and therefore does not have the possibility to reveal the spatial profile of the electric field along the gap. The Townsend method assumes a constant reduced electric field within the whole gap, which enables fitting of the intensity profile by a single exponential.

Next, we discuss the possible sources of systematic errors and the applicability of all four applied methods.

4.1. Electrical equivalent circuit

The proper measurements of the electrical parameters (current, charge and voltage) are a good basis for the $U_\mathrm{g}$ determination from the electrical equivalent circuit. The APTD features a slowly changing current with current peak maxima within a well measurable range, and as a result, the methodology used seems to be appropriate and no advanced approach is necessary, as e.g. in [46].

The important point of the electrical equivalent circuit method for averaged electric field determination in the gap is the selection of the values for the effective discharge capacitances [25, 47], or rather the selection of the approach as to how to obtain these values. Basically, there are two methods, theoretical and experimental. The capacitances can be analytically computed if the electrode and dielectrics arrangement is simple enough or even in slightly more complex setups [48]. An electrostatic simulation can also be undertaken to obtain these values if the design is reconstructed for a given voltage and geometry in 3D. These approaches, however, do not take into account any imperfections in the manufactured setup, edge effects, partial discharging (as described in [28, 29]) or local charging affecting the electric field distribution and thus directly influencing the charge transfer and also the capacitance. On the other hand, the experimental determination of capacitances from the QV-plot is directly related to the effective charge transfer, taking into account the distribution of the electric field in the particular experiment, although the determination of capacitances from the plot may suffer from over-simplification of the problem, e.g. insisting on the distribution of capacitance among the dielectric and the electrode gap. In our case, we have faced expected inconsistencies when comparing the results of the two methods described above. A more detailed analysis is required to properly quantify the difference, which is beyond the scope of this manuscript. There are additional effects such as stray capacitances addressed in [21, 49, 50].

The approach of taking the partial discharge, following the method of Peeters and van de Sanden [29], was also applied in our evaluation. The effect of partial discharging can be understood based on the QV-plot analysis. The resulting electric field values varied up to 10% with respect to those shown in figure 6. For the data set of the 7 kV applied voltage amplitude, as compared in figure 22 with other methods, the variation of the results obtained from the equivalent circuit was approximately 2% for our experiment.

4.2. Townsend coefficient: applicability

The main problem with the Townsend coefficient method is that the method cannot be used in the post-discharge phase, since the excited states are no longer populated dominantly by the electron impact excitation. As was demonstrated in the previous subsection, in the post-discharge phase the main mechanism for production of the optical emission is the pooling reaction producing upper states of FPS, SPS and HIR:

Equation (19)

Therefore, the optical emission is not proportional to $e^{\alpha(E/N) x}$, but rather to $e^{2 \alpha(E/N) x}$, since the spatial distribution of the $\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+,v = 0-1)$ states matches with the electron density profile, which has $e^{\alpha(E/N) x}$ dependency. Moreover, if the metastable species survive to the following half-period, they diffuse, which may further disturb the aimed result.

From an experimental point of view, this method is fast, easy to set up and robust. However care must be taken with the proper alignment of the camera to the discharge axis to avoid slanted observation. Also, with the use of macro-lenses, the depth of field is often very shallow, requiring a more closed aperture, i.e., resulting in a compromise between the amount of captured light signal and the in-focus area.

4.3. EFISH: parasitic signals

The EFISH method is straightforward and robust, with the possibility to obtain the values of the electric field with high spatial and temporal resolution given by the parameters of the laser and detectors used. The bottleneck of the EFISH method is the proper measurement calibration, considering the geometry of the measured field and changes in its cross-section with the laser beam during measurement and calibration. The presented simulation showed that even slight changes in the shape of the electric field could lead to a change in the obtained signal by tens of percent. Moreover, the measured dependencies can be easily disturbed by unknown parasitic signals, which may be invisible when the electric field is low but can cause another tens of percent decrease or increase in the signal for higher electric fields. When the parasitic field shifts the measured signal by a constant value, it can be easily treated by the proposed approach.

Under the conditions of Townsend discharge ignited between two parallel plates, we were able to perform two different calibration methods under similar geometric conditions. Their results differed by only 10%, which might indicate the reliability of all the input conditions: the proposed calibration approach and the treatment of the parasitic signal, or the value of the permittivity of the dielectrics, used in parallel in the evaluation of the equivalent circuit measurements.

4.4. Intensity ratio: overestimation of the results

The values of the reduced electric field obtained by the intensity ratio method are systematically higher compared to the reduced electric field obtained by the other methods (equivalent circuit, Townsend coefficient fitting, EFISH). Thus, a discussion concerning other processes potentially influencing the intensity ratio is needed. Since the VDF of $\mathrm{N_2}(\textrm{C}^3\Pi_\textrm{u})$ state was carefully checked, the systematic error, with regard to the other methods, may be caused by the omission of additional processes populating the $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+)$, whose VDF could not be checked.

  • The first potential candidate to discuss is the population of $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v^{^{\prime}} = 0)$ by electron impact excitation from molecular nitrogen ions:
    Equation (20)
    However, our calculations show that this source would be dominant over the ionization/excitation from $\mathrm{N_2}({\mathrm X}^1 \Sigma^+_\textrm{g},$ $v = 0)$ state:
    Equation (21)
    only at ionization fractions ($\frac{[\textrm{N}_2^+]}{[\textrm{N}_2]}$) higher than 10−9. However, in the case of APTD, we can expect just the values $\frac{[\textrm{N}_2^+]}{[\textrm{N}_2]} = \frac{10^{8}}{10^{19}} = 10^{-11}$, since N$_2^+$ ions are quickly converted to N$_4^{+}$ ions in these conditions.
  • The second candidate is the population of $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v^{^{\prime}} = 0)$ state from molecular nitrogen ions, involving metastable states, which was proposed in [51]:
    Equation (22)
    The authors suggest that this process could have a rate constant ($k_\textrm{A}$) of 2·10−10 cm3s−1. The study [31], which focused on the determination of $\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+)$, enables us to compare the possible source term caused by this reaction, defined as $k_\textrm{A}\cdot[\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+)]\cdot[\textrm{N}_2^+]$, with the source term of the electron impact excitation at the active discharge phase, defined as $k_\textrm{B}(E/N = 150\,\textrm{Td})\cdot [\mathrm e]\cdot [\textrm{N}_2]$. The $k_\textrm{B}(E/N = 150\,\textrm{Td})$ is the rate of the electron impact excitation of the $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+)_{v^{^{\prime}} = 0}$ state at the typical $E/N$ at active phase of the APTD. Supposing that $[\textrm{N}_2^+] \approx [\mathrm e]$, considering $k_\textrm{B}(E/N = 150\,\textrm{Td})$ = 5·10−11cm3s−1 and $[\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+)] \approx 10^{14}\,\textrm{cm}^{-3}$, we obtain values $k_\textrm{A} \cdot [\mathrm{N_2}(\textrm{A}^3\Sigma_\textrm{u}^+)] = 2\cdot10^4\,s^{-1}$ and $k_\textrm{B}(E/N = 150\,\textrm{Td})\cdot [\textrm{N}_2] = 1.3\cdot10^9\,s^{-1}$, which points out that the influence of the process (22) is, in our case, insignificant.
  • A third possible candidate is vibrational relaxation of $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v\gt0)$ into the $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v = 0)$ state. This would occur through higher vibrational states excitation:
    Equation (23)
    and subsequent vibrational relaxation into the v = 0:
    Equation (24)
    The cross sections for the excitation of $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v\gt0)$ states were measured in a several works (listed in table III of Tohyama and Nagata [52]) and they pointed out that $\frac{\sigma_\textrm{B1}}{\sigma_\textrm{B0}} \sim 0.1$, which is too low to significantly contribute to the $\mathrm{N^{+}_2}({\mathrm B}^2 \Sigma_\textrm{u}^+, v^{^{\prime}} = 0)$ state population.
  • The APTD generally contains high densities of $\mathrm{N^{+}_4}$ ion (∼1010 cm−3), which are produced by $\mathrm{N^{+}_2}$ collisions with molecular nitrogen. It was found by [53] that the $\mathrm{N^{+}_4}$ ion can be converted back to $\mathrm{N^{+}_2}$ ion electronic states by photodissociation through photons with wavelengths shorter than 1000 nm:
    Equation (25)
    This process could be responsible for the overproduction of N$_2^+$(X) species during the active discharge phase due to the strong SPS emission (especially (0,0) band at 337 nm) and consequently also for the systematic increase of the FNS/SPS ratio.

Indeed, further studies are needed to explain the systematically higher values estimated by the FNS/SPS intensity ratio method in this case.

5. Summary and conclusion

We have investigated the electric field development in weak microseconds-lasting APTD operated in a barrier discharge arrangement in pure nitrogen. The discharge mechanism under given conditions enabled a unique opportunity for determination of the electric field using four different methods. We used the laser-aided EFISH method, the optical emission-based FNS/SPS intensity ratio method, the electrical equivalent circuit approach and the determination via the first Townsend coefficient $\alpha(E/N)$ from the optical emission profile.

Comparing the results of four different methods, we could discuss in closer detail the possible uncertainties and mutual deviations in the resulting electric field values. It was found that the results of the equivalent circuit and the EFISH methods are in the closest agreement, and they are considered the most reliable. For the other two experimental methods, the probable sources of positive systematic error were evaluated and discussed. The method of electric field determination via the first Townsend coefficient is affected by the presence of nitrogen metastable states. The method of FNS/SPS ratio, while successfully utilized for streamer discharges [6, 54], is here at the limit of its applicability, apparently due to the impossibility of fully verifying all the underlying assumptions and the low electric field, which needs to be measured. Nevertheless, the overestimation of the electric field from the line ratio is comparable to the accuracy of the method estimated by the analysis of the molecular data (cross sections, lifetimes etc). The discrepancy between the ratio method and the other methods does not rule out that the assumptions of the ratio method are valid, but the real dependence of the intensity ratio $R(E/N)$ should be searched at the upper boundary of the confidence bands shown in figure  20. This would confirm the complexity of the uncertainty analysis of the intensity ratio method unraveled in [12].

Acknowledgments

This work was supported by the Czech Science Foundation (Project No. 15-04023S) and by the Strategy AV21 project. We acknowledge the support of the Project LM2018097 funded by the Ministry of Education, Youth, and Sports of the Czech Republic. TH acknowledges the support by Czech Science Foundation Project under Contract No. 21-16391S. This research has been also supported by the Project LM2023039 funded by the Ministry of Education, Youth and Sports of the Czech Republic.

Data availability statement

All data that support the findings of this study are included within the article (and any supplementary files).

Please wait… references are loading.