You are on page 1of 11

ETH Library

Flame stabilization with


nanosecond repetitively pulsed
discharge in a sequential
combustor

Conference Paper

Author(s):
Shcherbanev, Sergey; Dharmaputra, Bayu; Solana Pérez, Roberto ; Noiray, Nicolas

Publication date:
2022-01

Permanent link:
https://doi.org/10.3929/ethz-b-000527565

Rights / license:
In Copyright - Non-Commercial Use Permitted

Originally published in:


AIAA SciTech, https://doi.org/10.2514/6.2022-2254

Funding acknowledgement:
820091 - ThermoacOustic instabilities contRol in sequential Combustion cHambers (EC)

This page was generated automatically upon download from the ETH Zurich Research Collection.
For more information, please consult the Terms of use.
AIAA SciTech Forum 10.2514/6.2022-2254
January 3-7, 2022, San Diego, CA & Virtual
AIAA SCITECH 2022 Forum

Flame stabilization with nanosecond repetitively pulsed


discharge in a sequential combustor

Sergey A. Shcherbanev, Bayu Dharmaputra, Roberto Solana-Pérez, Nicolas Noiray


CAPS Laboratory, Department of Mechanical and Process Engineering, ETH Zurich, Zurich 8092, Switzerland

This paper describes the experimental results of the flame dynamics stabilized by nanosec-
ond repetitively pulsed discharge (NRPD) in a sequential combustor configuration. In such
configuration of the combustor, two flame regions are organized sequentially: (i) first stage
flame and (ii) second-stage sequential flame placed downstream of the first one. The sequential
flame is located downstream of the injection of the dilution air and the second stage fuel into the
hot vitiated products of the first stage flame. Several operation conditions of the second stage
Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

flame were considered in the present paper. The parameters of the first stage flame and amount
of the dilution air were fixed while the second stage fuel injection was composed of natural
gas/H2 mixture with different fractions of hydrogen. It is shown that the flame anchoring and
combustion dynamics is significantly affected by the amount of hydrogen and mean power of the
NRPD. Thermal effect of the plasma was studied with optical emission spectroscopy. Namely,
the temperature increase due to the fast gas heating was measured during applied high-voltage
pulses. It was shown that, for a mean vitiated flow temperature of 1000K, the temperature
increase due to nanosecond pulse discharge can reach up to 2100K. The coupling of the gas
flow with NRPD was identified for different frequencies and amplitudes of applied pulses. It
was found that for high pulse repetition frequency (PRF>40 kHz) the discharge propagation
and the parameter of the plasma have a cumulative effect. Spatially resolved electron density
and mean gas temperature in the plasma region were measured in order to characterize the
flow-discharge coupling in the flow direction. High speed OH∗ chemiluminescence is used for
quantitative analysis of the ignition and stabilization of the second stage flame.

I. Introduction
sequential combustor architecture is a key concept for the improvement of ground based heavy-duty gas turbines
S efficiency [1, 2]. This concept is attractive because it allows increasing operational and fuel flexibility at lower
emissions and higher combined cycle efficiency. This architecture of the combustion chamber is associated with a
substantially increased system complexity in terms of: (i) operating concept; (ii) 1st and 2nd stage flames are acoustically
coupled with each others; (iii) of combustion dynamics related to ignition, anchoring and blow-off physics. In such
sequential combustors, dilution air/fuel are injected downstream of the first stage lean premixed flame. The resulting
temperature in the mixing section of the sequential burner is low enough to allo high quality mixing before spontaneous
ignition, and therefore allows for ultra low NOx emissions, which is a case in this concept [3].
Sequential combustion chambers are thus of particular interest for clean combustion of sustainable fuels in gas
turbine applications. However, changes of fuel reactivity through variations of H2 fraction, can perturb the sequential
flame anchoring and dynamics. One of the possible ways to stabilize this sequential flame is to initiate a non-equilibrium
discharge in the mixing section where the vitiated flow from the first stage and dilution air are mixed with the second
stage fuel. It is known fact that NRPD is an efficient tool for the enhancement of the lean stability and blow-off limits [4].
Initiation of non-equilibrium plasma can serve as an additional flame holder that may prevent unwanted combustion
dynamics and strengthen the flame anchoring while changing the operation conditions.
This study is oriented on characterization of the discharge properties in the vitiated hot environment in a generic
sequential burner. The NRPD was initiated in pin-to-pin configuration in the mixing channel of the generic sequential
combustor. Two regimes of the discharge were identified: glow and spark. Electron number density and gas temperature
are measured for different discharge power. The effect of plasma initiation on flame ignition and anchoring on the
second stage of sequential combustor is studied. It is shown that the coupling between the flow and the plasma is
significantly affects the ignition dynamics of the sequential flame.

Copyright © 2022 by Sergey Shcherbanev, Nicolas Noiray. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

Fig. 1 Axially staged combustor consisting of a lean premixed first stage with a 4×4 array of turbulent flames
and a secondary jet flame in vitiated cross flow. The combustor has a constant cross section and is composed of
several modules equipped with quartz windows.

II. Experimental setup


The experiments were conducted using a generic combustor of cross section of 62 × 62 mm2 equipped with a
sequential burner cross section 25 × 38 mm2 operated at atmospheric pressure (see figure. 1). The first stage is a
4 × 4 array of lean turbulent premixed jet flames. The vitiated hot flow was produced by the combustion of a lean
natural gas/air mixturein the first stage of the combustor. The distance between the wave of the first stage burner and
the injection of the dilution air is 290 mm. The first-stage equivalence ratio and thermal power were held constant at
𝜙fs = 0.8 and 35 kW respectively. The dilution air is injected in the contraction section connecting the 62 × 62 mm2 and
25 × 38 mm2 modules through the 24 jet injectors distributed over the square cross section. The temperature of the
dilution air is 300 K. Further downstream in the 25 × 38 rectangular module the mixture of H2 /natural gas is injected
through the co-flow injector as it is shown in the figure. Pin-to-pin electrode system arranged perpendicularly to the axis
of the combustor is installed downstream of the injector. All modules are equipped with four quartz windows giving full
optical access along the combustor. One of the side walls of the second module is used to accommodate the inlet of the
second stage fuel, as well as the electrode arrangement. The high speed OH chemiluminescence imaging was used to
characterize the second stage flame. High speed chemiluminescence images were acquired at 10 kHz using a LaVision
highspeed star X camera with IRO intensifier. The camera was equipped with a 45 mm F/1.8 CERCO UV Lens in
combination with OH filter centered around 307.5 nm and with 10 nm full-width at half-maximum (FWHM).

Table 1 Operational conditions of the sequential combustor. First stage conditions are NG1 =0.7 g/s, air1 =15 g/s.
Dilution air was fixed to 18 g/s for all OPs.

Operational Natural gas Hydrogen Second stage power


Point (g/s) (g/s) (kW)
OP1 0.69 0.02 36.9
OP2 0.64 0.04 36.8
OP3 0.6 0.06 37.2
OP4 0.55 0.08 37.1
OP5 0.5 0.1 37.0

For direct discharge visualization the Pi-Max4 ICCD camera is used. The camera is equipped with 337 nm narrow
band filter in order to capture the morphology of the N∗2 . Optical emission spectroscopy techniques is used for the
analysis of the heat release in the discharge region. The temperature is measured from the emission spectra of the second
positive system of molecular nitrogen. The electron density was measured with the Stark broadening of the H 𝛼 line. For
spatially resolved measurements of the discharge temperature and electron density the discharge region was projected to

2
the entrance slit of the PI IsoPlane-320 spectrometer equipped with Pi-Max4 camera (are not shown in the figure).
The effect of the discharge initiation on the combustion of the second stage flame was characterized for different
operational points (OP). All considered experimental conditions are listed in a test matrix presented in table 1. The
parameters of the first stage flame as well as the amount of dilution air were fixed for all OPs. The equivalence ratio of
the first stage flame is set to 0.8 by injecting 0.7 g/s of the natural gas and 15 g/s of the air. These parameters result to
the 35 kW of thermal power of the first stage flame. The thermal power of the second stage flame was nearly equal to
37 kW for all OPs. Mixtures of natural gas/hydrogen with different fractions of hydrogen is used as fuel on the second
stage of the combustor.

III. Experimental results

A. Initiation of combustion with NRPD


The nanosecond repetitively pulsed discharge was initiated between the first and second stage flames downstream
of the dilution air and secondary fuel injection. For the analysis of the initiation of the combustion with nanosecond
Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

discharges the conditions with low fraction of hydrogen were selected. The second stage flame without initiation
of NRPD was not igniting for operation points OP1 and OP2 shown in table 1. For other conditions the flame was
autoignited. With increase of the hydrogen fraction in the fuel mixture the flame becomes more stable and shifts toward
the wake of the combustion chamber. A backward facing step configuration (from 25x38mm2 to 62x62mm2 ), enables
strong anchoring of the sequential flame due to strong recirculation regions. In order to shift the flame downstream of
the combustion chamber a temperature and equivalence ratio of the gas mixture on the second stage, 18 g/s of dilution
air were injected that resulted in mean flow temperature about 950 K and equivalence ratio 0.57-0.6.
Figure 2 shows the high speed chemiluminescence images of the flame while initiation of the NRPD. The relative
position of the electrode system, fuel injector are shown the combustion chamber is shown in the figure. First column
demonstrated the dynamics of the ignition kernels and the initiation of the flame for the OP2 conditions. It can be seen
that there is no flame in the combustion chamber when the NRPD is turned off. While the discharge is switched on the
first distinct ignition kernels take place in the combustion chamber that further expand and the well developed flame is
stabilized in the downstream part of the chamber. The discharge was initiated during 0.1 s at 50 kHz pulse repetition
frequency (PRF) and 13.5 kV of applied voltage.
The second column in figure 2 corresponds to the OP5 conditions. Even without plasma the flame is autoignited and
stabilized in the combustion chamber. The flame has asymmetrical structure. This can be caused by a non-symmetric
injector geometry. In the case where the discharge is switched off, the flame is distributed along the entire length of the
combustion chamber. During the discharge initiation the distinct spontaneous ignition kernels can be partially observed
at the outlet of the mixing channel. These kernels propagate downstream, merge with the flame and stabilize the flame
on both (upper and lower) parts of the combustion chamber wake. The flame becomes anchored and more symmetrical
with respect to the axis of the combustor. The overall length of the flame is reduced and for a particular case of OP5
occupies only a half of the volume of the combustion chamber.
On the averaged OH∗ chemiluminescence the mean flame profiles can be seen. For the OP2 the flame was initiated
and stabilized with NRPD. Once the discharge is turned off some remaining flame can be still observed in the chamber
during the following 0.02-0.05 s. At later time the flame is blown out. For the OP5 the flame profile has rather symmetric
structure and well anchored to the outlet of the mixing channel. When the plasma is off the flame is detached and
stabilized downstream of the chamber wake.
The time averaged profiles of the OH∗ can be seen in figure 3 for all OPs and fixed parameters of the applied HV
pulses: 𝑉 = 19.5 kHz and 𝑃𝑅𝐹 = 50 kHz. For the lowest hydrogen content in the fuel mixture (OP1), the unstable
pulsed ignition with plasma was observed. The ignition take place well downstream near the outlet of the combustion
chamber. Ignition kernels were instantly extinguished once the plasma discharge is turned off. For the voltage amplitudes
lower than 16 kV any ignition events was observed for the OP1. For the OP2 the similar behaviour to the one explained
above can be observed with the only difference that the flame is shifted upstream because the higher amplitude of the
applied voltage. For the case of OP3 the ignition kernels in the mixing channel can be observed more regularly because
of higher reactivity of the mixture.
For even higher amount of hydrogen in the gas mixture (OP4 and OP5) the flame without plasma is anchored from
the one side of the combustion chamber inlet. However, once the NRPD is on, the ignition takes place in the vicinity
of the electrode system and well developed flame can be observed at the outlet of the mixing channel. It should be
mentioned that the flame in the mixing channel is initiated from the side of the cathode electrode. The energy density in

3
Instantaneous OH chemiluminescence
100 Mixing channel Combustion chamber
X, (mm)
Injector Electrodes
50

t=1.5 ms t=1.9 ms
0
100
X, (mm)

50

t=2.1 ms t=2.7 ms
0
100
X, (mm)

50

t=2.4 ms t=2.9 ms
0
100
X, (mm)
Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

50

t=2.7 ms t=3.2 ms
0
100
X, (mm)

50

t=3.3 ms t=3.7 ms
0
100
(mm)
X, (mm)

50
X,

0
t=3.9 ms t=4.1 ms
100
X, (mm)

50

t=4.5 ms t=4.8 ms
0

Z, (mm) Z, (mm)
Averaged OH chemiluminescence
Plasma ON Plasma ON
X, (mm)

Plasma OFF Plasma OFF


X, (mm)

Z, (mm) Z, (mm)

Fig. 2 Time-resolved OH∗ chemiluminescence images for OP2 (left column) and OP5 (right column). Discharge
parameters: PRF=50 kHz, V=13.5 kV.

the cathode region is higher than that near the anode, the further spectroscopy analysis confirmed that issue.

B. Fast-gas heating in hot vitiated environment


It was shown for the preheated air that the nanosecond spark discharge initiates a fast gas heating processes [5].
Experimentally it was shown that the temperature increase rate during the discharge phase and the near afterglow can be
as high 30 − 40 K/ns. Later it was shown in [6] by means on kinetic modelling that the main mechanisms of such a high
heat release rate is explained by a huge portions of energy release from the excited state of molecular nitrogen that
instantly goes to has heating (e-T energy transfer). This energy release from internal degrees of freedom is carried out

4
Mixing channel Combustion chamber
OP 1 Electrodes OP 2 OP 3

Injector plasma on Discharge emission plasma on plasma on

plasma off plasma off plasma off

OP 4 OP 5

Ignition plasma on Ignition plasma on


Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

plasma off plasma off

Fig. 3 Time averaged images of the OH∗ chemiluminescence. OP - stands for operational point listed in the
table. Applied voltage V=19.5 kV, PRF = 50 kHz.

Fig. 4 Dynamic of gas (left) and vibrational (right) temperature during the discharge for different applied
voltages.

within collisional quenching of N∗2 with O2 molecules leading to the dissociation of the latter and energy release as high
as several electronvolts.
This mechanism of fast gas heating is well studied in air [7]. The kinetic mechanism of the fast gas heating in more
complex gas mixtures is not known nowadays. The energy balance and transfer between different degrees of freedom
might be significantly complicated by the presence of 3- or higher atomic molecules.
Here we present the temperature dynamics in the discharge region for the plasma initiated in vitiated hot flow gas.
For further analysis the discharge was initiated at atmospheric pressure for a OP5 conditions. Figure 4 shows the
dynamics of the gas and vibrational temperature measured from the optical emission of N2 (C-B)(Δ𝜈 = 1) transition
of second positive system. The presented temperature curves are shown together with the profile of the applied HV
pulse. The emission spectra are acquired with the system described in Experimental setup section. Camera gate for the
acquisition of the optical emission spectra was set to 2 ns with 100 on-CCD accumulations. Experimental emission
spectra were fitted with the SpecAir [8] software in order to obtain rotational and vibrational temperatures in the plasma
region. Temperatures were measured for 4 different amplitudes of applied voltages: 10.5, 12.0, 22.5 and 27.5 kV. As
one can see in the left plot of figure 4, there is practically no temperature increase for 10.5 kV case. With increase of
amplitude of applied voltage heat release rate gradually increases. For 27.0 kV amplitude the temperature is increased

5
0.01

Energy deposition per pulse, J


0.001

a) b)
1E-4
10 12 14 16 18 20 22 24 26 28
Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

Applied voltage, kV

Fig. 5 a) Signal from back current shunt; b) deposited energy per pulse for different applied voltages.

from 1000 K to 2200 K during 14 ns which is equivalent to the increase rate of 85 K/ns. This value of the gas heating
rate exceeds the values given in the literature for measuring the temperature dynamics in the air.
Figure 4 also shows the dynamic of vibrational temperatures during the applied HV pulse. It can be clearly seen that
the vibrational temperature values start at about 4000 K and then decreases during 6-8 ns for all voltages. For the lowest
voltage, this trend continues until the end of the applied pulse. However, with increase of voltage amplitude the value
of vibrational temperature has an infimum. After this value the vibrational temperature starts increasing. The point
of minimal vibrational temperature is shifted towards the leading edge of the pulse with increase of the amplitude of
applied voltage.

C. Deposited energy and electron density


Nanosecond discharges in open electrode configurations are characterized by several regime of operation: corona,
glow, spark [9] and thermal spark [10]. Each of the regime can be distinguished by analyzing the discharge specific
energy deposition, electron density and/or ionization degree, and heat release rate. This section presents the deposited
energy and electron density for different peak voltages.
The energy deposition into the discharge was measured by acquiring the voltage waveforms with and without plasma
with calibrated back-current shunt implemented in the transmission line connecting the high voltage generator and
electrode system. Figure 5a) shows the characteristic waveforms of the incident and reflected from the electrode system
pulses with and without discharge initiation. The deposited energy is calculated from the difference between the energies
stored into the reflected pulse with and without plasma. It allows to take into account all electrical losses such that
attenuation of the signal in the transmission line, charging of the electrical capacity of the electrode system and the
displacement current.
Figure 5b) shows the deposited energy values for different peak voltages of the applied pulses. Sharp overshoot of
the energy value about the order of magnitude can be seen while increasing the voltage from 10.5 to 12 kV. Further
increase of the voltage amplitude changes the value of deposited energy from 3 to 15 mJ while increasing the peak
voltage from 12 to 28 kV. The lowest energy value that is observed for 10.5 kV corresponds to the glow discharge type.
This is confirmed by the weak heat release that was shown in previous section.
Electron density measurements are carried out by evaluating the stark broadening of the H 𝛼 line at 656 nm. The
apparatus slit function is taken into account during the post-processing of the emission spectra. The details of the
electron density measurement method can be found elsewhere [10]. The sharp increase in electron density also takes
place with increase in peak voltage. Electron density changes from 3.5·1014 cm−3 at 10.5 kV to the 4.6·1015 cm−3 at
12 kV. Typical electron number density for the streamer and glow discharges in ambient air is ∼ 1014 cm−3 [11] where
as the spark discharge is characterized by the 𝑛𝑒 ∼ 1015 − 1017 cm−3 [12]. Such a sharp jump in the electron density
corresponds to a change in the plasma regime from a glow discharge to a spark.

6
1E16

Electron density, cm-3 1E15


Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

1E14
10 12 14 16 18 20 22 24
Applied voltage, kV

Fig. 6 Electron density as a function of peak voltage.

D. Spatially resolved ionization degree


With increase of the PRF and applied voltage amplitude, the discharge-vitiated flow coupling becomes substantial.
In our experiments the channel morphology remained unchanged from pulse to pulse for PRF< 50 kHz. For frequencies
10-40 kHz and any voltages, the plasma channel remained straight between the two pin electrodes. With the increase of
PRF up to 50 kHz, depending on the amplitude of applied voltage, the plasma channel remained straight or curved in the
flow direction. To demonstrate the strong effect of the plasma-flow coupling, we selected the following parameters:
OP5, PRF=100 kHz, V=19.5 kV.
Figure 7a) shows the discharge morphology averaged over 10 subsequent pulses. It is clearly seen that the majority
of time discharge is on a state curved in flow direction. The mechanism of plasma/flow coupling is the following: (i)
the first pulse generates the spark discharge with rather intense heating due to the fast gas heating mechanism; (ii) the
heated region is convected downstream by vitiated flow; (iii) for the second pulse, the plasma tends to propagate to the
region of higher reduced electric field (E/N), which is higher within preheated by previous pulse convected gas region.
(iv) the second pulse increases the temperature of the convected region more (v) when subsequent pulse of the discharge
is initiated, the step (iii) is repeated until the breakdown by the shortest way becomes more favorable.
In order to measure spatially resolved electron density and gas temperature, the optical system equipped with an
image rotator (Dove prism) was used in order to focus the middle line (white dashed line in figure 7a)) on the entrance
slit of the spectrometer. The central line of the inter-electrode gap along the flow direction was used for further analysis
of the plasma parameters. Characteristic spatially resolved H 𝛼 line is shown in figures 7b) and c). The electron density
was determined from the Lorenz FWHM by fitting the spectral line with voigt profile. The spatially resolved emission
spectrum of the second positive system that was used for temperature measurements along the flow direction is shown in
figure 7d). The orientation of the electrode system with respect to the entrance slit is designated with white color in the
same figure.
Figure 7e) shows the temperature and electron density as a function of distance from the electrode axis. It is clearly
seen that the temperature increases from ∼ 1000 K at 0 mm position until 4000 K at the position 8 mm from the axis
of the electrode system. It can be noticed that the sharpest temperature increase takes place between 2 and 5 mm.
At distances exceeding 5 mm, the gas temperature reaches saturation and does not rise to high values. This happens
because of the intense heat release with the flow. Meanwhile, it can be seen that the maximum electron density is
achieved at 3.5 mm from the electrodes. At higher distances the electron density gradually decreases.
The ionization degree was calculated as:
𝑛𝑒
𝛼= · 100% (1)
𝑛𝑒 + 𝑛𝑔𝑎𝑠
where 𝑛𝑒 and 𝑛𝑔𝑎𝑠 are electron and gas number density respectively. Gas number density was calculated for a local

7
+ HV electrode

Slit
orientation

Flow

- HV electrode a)

Slit orientation
Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

Gap axis

b) d)

Exp. data 5-6 mm


1-2 mm
1.0 1.0
Voigt fit 3-4 mm
5-6 mm
8-9 mm
Normalized intensity

0.8 0.5

0.6 0.0

655 656 657


Wavelength, nm

0.4

0.2

c) e)
0.0

652 654 656 658 660


Wavelength, nm

Fig. 7 Space resolved measurements of the electron density and gas temperature in the plasma region in the
middle of the inter-electrode gap. (a) and (c) spatially resolved emission spectra of the H 𝛼 line and second
positive system of N2 respectively; (b) H 𝛼 line profile for different distances from the electrode axis; (d) Spatially
resolved gas temperature, electron density and ionization degree. Operational conditions: 100 kHz, V=19.5 kV,
OP5

temperature value T and atmospheric pressure as 𝑛𝑔𝑎𝑠 = 𝑁 𝐿 · (𝑇0 /𝑇), where 𝑁 𝐿 = 2.68678 · 1019 cm−3 - is Loschmidt
constant. The ionization degree gives information about the specific energy deposition. As can be seen, the highest
specific energy is correlated with the highest electron number density region.
According to the above-mentioned descriptions, 3 regions of the plasma-flow coupling can be distinguished:
• Pre-heating region, where the energy is progressively delivered and increased from pulse to pulse. In this region,
the discharge heats up the gas by several hundred degrees producing active species and metastable states. They
are convected further downstream allowing initiation of the discharge with higher ionization degree.
• The region with the highest efficiency of the energy transfer.
• Thermal maintenance region, where the balance of energy is established between the gas heating by subsequent

8
pulses and heat transfer to the surrounding flow.

IV. Conclusion
An experimental study was carried out to investigate the effects of nanosecond repetitively pulsed discharge (NRPD)
on the stabilization of the flame in a generic sequential combustor. By means of fast OH∗ chemiluminescence imaging, it
was shown that the effect of NRPD on the flame anchoring is significantly depend on the composition and the reactivity
of the gas mixture. The bursting ignition events can be observed for NRPD initiation at a low fraction of hydrogen
(OP1), but the flame was not stabilized in the combustion chamber. For a moderate amount of hydrogen (OP2), the
ignition efficiency is significantly affected by the mean power of the NRPD, which can be controlled either by voltage
or by PRF of applied pulses. For a higher content of the hydrogen in second stage fuel mixtures (OP3-5), the flame
was auto-ignited without plasma. However, the NRPD showed a significant effect on flame anchoring and stabilisation
of flickering flames on the second stage of the sequential combustor by both glow and spark NRPD regimes. The
temperature dynamics during a single pulse was measured for different deposited energies. It was shown that the fast gas
heating leads to a significant temperature increase rate of up to 85 K/ns. The analysis of the energy deposition and the
Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

electron density as a function of the amplitude of applied voltage demonstrated the presence of two regimes of discharge:
glow and spark. In a glow regime, no significant gas heating was observed, and the electron density did not exceed
5·1014 cm−3 at 10.5 kV. Whereas, for a spark regime, the single event of the discharge leads to the rapid heat release and
the electron number densities 1-2 orders of magnitude higher than that in the glow regime. Spatially resolved emission
spectroscopy measurements were carried out for electron number density and temperature profiles along the gas flow
direction to characterize the coupling between the plasma and hot flow. It was shown that for considered operation
points, the NRPD with pulse repetition frequencies 𝑃𝑅𝐹 ≤ 40 kHz is not coupled with the flow for the peak amplitudes
of applied voltage 𝑉 ≤ 27 kV. For spark NRPD with 𝑃𝑅𝐹 ≤ 40 kHz, the plasma parameters are significantly affected by
the preceding applied pulses and have a cumulative character of plasma properties.

Acknowledgments
This study was supported by the European Research Council under the ERC Consolidator Grant (No: 820091) TORCH
(2019-2024).

References
[1] Guthe, F., Hellat, J., and Flohr, P., “The Reheat Concept: The Proven Pathway to Ultra-Low Emissions and High Efficiency and
Flexibility,” Turbo Expo: Power for Land, Sea, and Air, Vol. 47918, 2007, pp. 641–648.

[2] Ruedel, U., Stefanis, V., Ramaglia, A., and Florjancic, S., “Development of the new Ansaldo energia gas turbine technology
generation,” Turbo Expo: Power for Land, Sea, and Air, Vol. 50831, American Society of Mechanical Engineers, 2017, p.
V003T08A006.

[3] Ciani, A., Bothien, M., Bunkute, B., Wood, J., and Früchtel, G., “Superior fuel and operational flexibility of sequential
combustion in Ansaldo Energia gas turbines,” Journal of the Global Power and Propulsion Society, Vol. 3, 2019, pp. 630–638.

[4] Di Sabatino, F., and Lacoste, D. A., “Enhancement of the lean stability and blow-off limits of methane-air swirl flames at
elevated pressures by nanosecond repetitively pulsed discharges,” Journal of Physics D: Applied Physics, Vol. 53, No. 35, 2020,
p. 355201.

[5] Rusterholtz, D., Lacoste, D., Stancu, G., Pai, D., and Laux, C., “Ultrafast heating and oxygen dissociation in atmospheric
pressure air by nanosecond repetitively pulsed discharges,” Journal of Physics D: Applied Physics, Vol. 46, No. 46, 2013, p.
464010.

[6] Popov, N., “Fast gas heating initiated by pulsed nanosecond discharge in atmospheric pressure air,” 51st AIAA Aerospace
Sciences Meeting including the New Horizons Forum and Aerospace Exposition, 2013, p. 1052.

[7] Popov, N., “Fast gas heating in a nitrogen–oxygen discharge plasma: I. Kinetic mechanism,” Journal of Physics D: Applied
Physics, Vol. 44, No. 28, 2011, p. 285201.

[8] Laux, C., “Optical diagnostics and collisional-radiative models,” VKI LS Course on hypersonic entry and cruise vehicles, Palo
Alto, California, USA, 2008.

9
[9] Pai, D. Z., Lacoste, D. A., and Laux, C. O., “Transitions between corona, glow, and spark regimes of nanosecond repetitively
pulsed discharges in air at atmospheric pressure,” Journal of Applied Physics, Vol. 107, No. 9, 2010, p. 093303.

[10] Minesi, N., Stepanyan, S., Mariotto, P., Stancu, G. D., and Laux, C., “Fully ionized nanosecond discharges in air: the thermal
spark,” Plasma Sources Science and Technology, Vol. 29, No. 8, 2020, p. 085003.

[11] Packan, D., Repetitive nanosecond glow discharge in atmospheric pressure air, Stanford University, 2004.

[12] Sainct, F. P., Urabe, K., Pannier, E., Lacoste, D. A., and Laux, C. O., “Electron number density measurements in nanosecond
repetitively pulsed discharges in water vapor at atmospheric pressure,” Plasma Sources Science and Technology, Vol. 29, No. 2,
2020, p. 025017.
Downloaded by Sergey Shcherbanev on December 31, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2022-2254

10

You might also like