+ All Categories
Home > Documents > Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least...

Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least...

Date post: 22-Jan-2021
Category:
Upload: others
View: 3 times
Download: 0 times
Share this document with a friend
10
Combustion and Flame 206 (2019) 239–248 Contents lists available at ScienceDirect Combustion and Flame journal homepage: www.elsevier.com/locate/combustflame Counterflow diffusion flame oscillations induced by ns pulse electric discharge waveforms Yong Tang a,b , Marien Simeni Simeni a , Kraig Frederickson a , Qiang Yao b , Igor V. Adamovich a,a Department of Mechanical and Aerospace Engineering, Nonequilibrium Thermodynamics Laboratory, The Ohio State University, Columbus, OH 43210, USA b Department of Energy and Power Engineering, Key Laboratory for Thermal Science and Power Engineering of Ministry of Education, Tsinghua University, Beijing 100084, China a r t i c l e i n f o Article history: Received 21 February 2019 Revised 1 May 2019 Accepted 1 May 2019 Available online 16 May 2019 Keywords: Counterflow flame Ns pulse discharge Flame oscillations Ion wind Electric field Second harmonic generation a b s t r a c t Repetitive ns pulse, dielectric barrier discharge voltage waveforms, combined with a tail several ms long, are used to induce oscillations of a counterflow atmospheric pressure diffusion flame. A baseline ns pulse discharge operated at 10 Hz results in a relatively modest oscillatory response of the flame, which be- comes more pronounced in burst mode operation, at the same burst repetition rate of 10 Hz. This effect is most likely caused by the residual electric field after the discharge pulse, producing the electrohy- drodynamic (EHD) force (“ion wind”) on the charges generated during the discharge, although plasma chemistry and Joule heating by the discharge may also contribute. Manipulating the external circuit to add a variable duration tail to the discharge pulse, without changing the pulse shape during breakdown or the pulse repetition rate, considerably enhances the impulse of the EHD force and increases the ampli- tude of the flame oscillations. To quantify this effect, the electric field distribution between the electrodes during and after the discharge pulse is measured by ps Electric Field Induced Second Harmonic (E-FISH) diagnostic. The results show that the electric field is maintained during the voltage tail, although it is lower compared to the Laplacian field due to the charge accumulation on the dielectric sleeves covering the electrodes. The time scale of the flame oscillations at the present conditions, of the order of 10 ms, is limited by the relatively slow momentum transfer from the ions to the neutral species. The present results demonstrate feasibility of enhancing the flame control authority, by combining a high peak ion- ization fraction generated by a ns pulse discharge with the EHD force applied on a long time scale, using a single plasma generator. © 2019 Published by Elsevier Inc. on behalf of The Combustion Institute. 1. Introduction The effect of sub-breakdown AC and DC electric fields on flame stabilization has been studied extensively over the last several decades (e.g. [1,2] and references therein), as well as more re- cently [3–19]. Over the last decade, the scope of these studies has expanded to include the effect of electric discharges, primar- ily those sustained by repetitive ns duration pulses [20–33], since they generate more stable and reproducible high-pressure plasmas compared to most DC, AC, RF, and microwave discharges. Gener- ally speaking, the effect of sub-breakdown electric fields on the flame reaction zone can be described in terms of the electrohydro- dynamic (EHD) force (“ion wind”) [34–38], although its accurate quantitative description requires the development and validation of a high fidelity kinetic model incorporating detailed kinetics and Corresponding author. E-mail address: [email protected] (I.V. Adamovich). transport of positive and negative ions, as well as electrons. De- pending on the flame and flow geometry, as well as the electric field amplitude and frequency, the ion wind may displace and dis- tort the reaction zone [4,5,8,12,13,18,19], induce flame instabilities [3,6,9,11], and generate coherent flow structures [10,14–17]. Quali- tatively, the magnitude of this effect is limited by the EHD inter- action parameter (the ratio of the Coulomb force work to the flow kinetic energy) [39] η ε 0 E 2 ρ u 2 ξ en i ϕ ρ u 2 ξ , (1) where ε 0 is dielectric permeability of vacuum, n i is the ion num- ber density, E and ϕ are the electric field and the potential dif- ference across the space charge region (e.g. the reaction zone), ξ is the applied electric field duty cycle, ρ and u are the flow density and velocity. For the EHD force interaction to be significant, the interaction parameter should be of the order of one, η 1. Since chemi-ionization processes in flames generate fairly low ion den- https://doi.org/10.1016/j.combustflame.2019.05.002 0010-2180/© 2019 Published by Elsevier Inc. on behalf of The Combustion Institute.
Transcript
Page 1: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

Combustion and Flame 206 (2019) 239–248

Contents lists available at ScienceDirect

Combustion and Flame

journal homepage: www.elsevier.com/locate/combustflame

Counterflow diffusion flame oscillations induced by ns pulse electric

discharge waveforms

Yong Tang

a , b , Marien Simeni Simeni a , Kraig Frederickson

a , Qiang Yao

b , Igor V. Adamovich

a , ∗

a Department of Mechanical and Aerospace Engineering, Nonequilibrium Thermodynamics Laboratory, The Ohio State University, Columbus, OH 43210, USA b Department of Energy and Power Engineering, Key Laboratory for Thermal Science and Power Engineering of Ministry of Education, Tsinghua University,

Beijing 10 0 084, China

a r t i c l e i n f o

Article history:

Received 21 February 2019

Revised 1 May 2019

Accepted 1 May 2019

Available online 16 May 2019

Keywords:

Counterflow flame

Ns pulse discharge

Flame oscillations

Ion wind

Electric field

Second harmonic generation

a b s t r a c t

Repetitive ns pulse, dielectric barrier discharge voltage waveforms, combined with a tail several ms long,

are used to induce oscillations of a counterflow atmospheric pressure diffusion flame. A baseline ns pulse

discharge operated at 10 Hz results in a relatively modest oscillatory response of the flame, which be-

comes more pronounced in burst mode operation, at the same burst repetition rate of 10 Hz. This effect

is most likely caused by the residual electric field after the discharge pulse, producing the electrohy-

drodynamic (EHD) force (“ion wind”) on the charges generated during the discharge, although plasma

chemistry and Joule heating by the discharge may also contribute. Manipulating the external circuit to

add a variable duration tail to the discharge pulse, without changing the pulse shape during breakdown

or the pulse repetition rate, considerably enhances the impulse of the EHD force and increases the ampli-

tude of the flame oscillations. To quantify this effect, the electric field distribution between the electrodes

during and after the discharge pulse is measured by ps Electric Field Induced Second Harmonic (E-FISH)

diagnostic. The results show that the electric field is maintained during the voltage tail, although it is

lower compared to the Laplacian field due to the charge accumulation on the dielectric sleeves covering

the electrodes. The time scale of the flame oscillations at the present conditions, of the order of ∼10 ms,

is limited by the relatively slow momentum transfer from the ions to the neutral species. The present

results demonstrate feasibility of enhancing the flame control authority, by combining a high peak ion-

ization fraction generated by a ns pulse discharge with the EHD force applied on a long time scale, using

a single plasma generator.

© 2019 Published by Elsevier Inc. on behalf of The Combustion Institute.

1

s

d

c

h

i

t

c

a

fl

d

q

o

t

p

fi

t

[

t

a

k

η

w

b

f

h

0

. Introduction

The effect of sub-breakdown AC and DC electric fields on flame

tabilization has been studied extensively over the last several

ecades (e.g. [1,2] and references therein), as well as more re-

ently [3–19] . Over the last decade, the scope of these studies

as expanded to include the effect of electric discharges, primar-

ly those sustained by repetitive ns duration pulses [20–33] , since

hey generate more stable and reproducible high-pressure plasmas

ompared to most DC, AC, RF, and microwave discharges. Gener-

lly speaking, the effect of sub-breakdown electric fields on the

ame reaction zone can be described in terms of the electrohydro-

ynamic (EHD) force (“ion wind”) [34–38] , although its accurate

uantitative description requires the development and validation

f a high fidelity kinetic model incorporating detailed kinetics and

∗ Corresponding author.

E-mail address: [email protected] (I.V. Adamovich).

t

a

i

c

ttps://doi.org/10.1016/j.combustflame.2019.05.002

010-2180/© 2019 Published by Elsevier Inc. on behalf of The Combustion Institute.

ransport of positive and negative ions, as well as electrons. De-

ending on the flame and flow geometry, as well as the electric

eld amplitude and frequency, the ion wind may displace and dis-

ort the reaction zone [4,5,8,12,13,18,19] , induce flame instabilities

3,6,9,11] , and generate coherent flow structures [10,14–17] . Quali-

atively, the magnitude of this effect is limited by the EHD inter-

ction parameter (the ratio of the Coulomb force work to the flow

inetic energy) [39]

∼ ε 0 E 2

ρu

2 ξ ∼ e n i �ϕ

ρu

2 ξ , (1)

here ε0 is dielectric permeability of vacuum, n i is the ion num-

er density, E and �ϕ are the electric field and the potential dif-

erence across the space charge region (e.g. the reaction zone), ξ is

he applied electric field duty cycle, ρ and u are the flow density

nd velocity. For the EHD force interaction to be significant, the

nteraction parameter should be of the order of one, η ∼ 1. Since

hemi-ionization processes in flames generate fairly low ion den-

Page 2: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

240 Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248

c

f

O

i

t

i

a

0

s

u

c

a

i

o

i

o

b

a

t

o

w

m

u

n

l

m

t

w

T

o

t

e

d

m

e

fl

p

i

e

a

p

v

T

p

d

c

p

A

a

r

r

s

o

m

7

P

P

d

t

t

c

C

c

c

sity, n i ∼ 10 8 –10 11 cm

−3 [2] , this limits the applicability of EHD

interactions to relatively low-speed flows.

In electric discharges sustained by higher than breakdown

electric fields, the effect on the flame is complicated significantly

by the generation of excited species and radicals in the plasma,

resulting in plasma chemical reactions, as well as by the Joule

heating accelerating the rate coefficients of chemical reactions.

As discussed in numerous experimental studies, these processes

may significantly increase the burning velocity, the flame speed,

the blow-off limit, and the flammability limits [20–33] . Compared

to these effects, significant EHD interaction in highly transient,

very low duty cycle plasmas ( ξ ∼ 10 −5 –10 −3 ), such as generated

by repetitive ns pulse discharges used in most plasma assisted

combustion studies, appears unlikely, in spite of much higher

electron / ion densities, up to at least n + ∼ 10 14 –10 15 cm

−3 [40] .

However, the slowly varying residual electric field produced by

the charge accumulation on dielectric surfaces, such as occurs in

dielectric barrier discharges [41] , may well increase the effective

duty cycle and therefore the magnitude of the ion wind effect.

In Ref. [22] , the effect of an ns pulse electric discharge on the

lifted jet flame was detected only when a dielectric barrier was

used, indicating that the ion wind was the dominant factor in

flame stabilization. In this study, the effect of a dielectric barrier

ns pulse discharge on the flame lift-off height was also found to

be comparable to that of sub-breakdown DC field and AC fields.

The magnitude of the EHD interaction may be increased sig-

nificantly by combining the pulsed discharge, which generates a

much higher electron/ion density compared to chemi-ionization

in a flame, with a sub-breakdown DC or AC electric field, with a

much higher duty cycle. Therefore this approach may produce a

greater time-integrated impulse of the EHD force, limited by the

peak electron density generated during the discharge pulse, the

rate of electron-ion and ion-ion recombination, and the breakdown

threshold between the discharge pulses.

Quantifying the enhanced plasma/EHD effect on the flame

requires measurements of the electric field distribution in the

plasma, since the externally applied field may be perturbed signifi-

cantly by the charge separation and charge accumulation on the di-

electric surfaces. In the present work, we are using ps Electric Field

Induced Second Harmonic (E-FISH) diagnostic [42,43] , which has

also been employed in our previous experiments in ns pulse dis-

charge plasmas sustained in air and in a hydrogen diffusion flame

[44,45] . At the present conditions, the electric field is put on the

absolute scale by measuring a Laplacian field during the voltage

pulse ∼100 ns duration, before breakdown.

The objectives of this work are to demonstrate that adding a

long (ms time scale) “tail” to the ns pulse discharge waveform

may enhance significantly the EHD effect on the flame, and to

quantify the flame forcing by measuring the electric field distri-

bution during the discharge pulse and in the afterglow, in simple

geometry. For this, we monitor the amplitude of the oscillations

of the reaction zone of a non-premixed counterflow flame, which

is known to be sensitive to the electrohydrodynamic forcing

[14–16] . These proof-of-concept measurements may contribute

to the development of an effective combustion stabiliza-

tion/flameholding method, as well to the development and

validation of a predictive kinetic modeling of the electric-discharge

enhanced EHD forcing of the flame. Such a model needs to incor-

porate electron impact ionization processes and plasma chemical

reactions, Joule heating, electron/ion transport, and their coupling

to the neutral flow, in a realistic geometry.

2. Experimental

Figure 1 shows a schematic of the burner, the flame, the dis-

charge electrode assembly, and the position of the laser beam. A

ustom-made burner is used to sustain a laminar counterflow dif-

usion flame in a mixture of 14% CH 4 /86% Ar (bottom) and 42%

2 /58% Ar (top), with the nitrogen co-flow. The diameters of the

nner and co-flow nozzles are D 1 = 10 mm and D 2 = 16 mm, respec-

ively, and the gap between the fuel and oxidizer nozzle exits

s �1 = 15 mm. The flow rates of the fuel and oxidizer mixtures

re 1.25 slm each, and the co-flow rates are 2.25 slm (top) and

.75 slm (bottom). The fuel and oxidizer mixture compositions are

et up to match the momenta of the two flows, based on the vol-

me flow rates and molecular weights of the components. The N 2

o-flow rates are chosen to help balance the effect of buoyancy

nd to reduce the flame distortion when the electrodes are put

n place. The estimated flow velocity at the exit of the fuel and

xidizer nozzles is U 0 = 26 cm/s, and the estimated co-flow veloc-

ty is U 1t = 30 cm/s (top) and U 1b = 10 cm/s (bottom). The momenta

f the fuel and oxidizer flows are balanced. The Reynolds number

ased on the nozzle exit diameter is Re D1 ≈ 130. The flame di-

meter is approximately 25 mm. As expected, the flame is not en-

irely flat, with the peripheral part slightly lifted by the buoyancy

f the combustion products. Temperature distributions in the flame

ith and without the effect of buoyancy, predicted by 2-D axisym-

etric simulations of a non-premixed counterflow laminar flame

sing OpenFOAM software with GRI-Mech 1.2 combustion mecha-

ism [46] demonstrate that incorporating buoyancy results in the

ifting of the outer edge of the flame, consistent with the experi-

ental observations.

The discharge is generated between two parallel brass rod elec-

rodes d 1 = 0.6 mm in diameter, covered by alumina ceramic tubes

ith the outside diameter of d 2 = 1.5 mm, as shown in Fig. 1 .

he overlap between the electrodes is L 0 = 24 mm, with the non-

verlapping length for each electrode of L 1 = 6 mm, and the elec-

rode gap of �2 = 12 mm. The laser beam is directed parallel to the

lectrodes. The burner/electrode assembly is mounted on a three-

imensional translation stage, such that the laser beam can be

oved relative to the flame and the electrodes, in the plane of the

lectrodes. Putting the electrodes in place somewhat distorts the

ame and displaces it approximately 1–2 mm above the horizontal

lane of symmetry (see Fig. 1 ). In the present work, two parallel

nsulated rod electrodes are used, rather than two parallel mesh

lectrodes, to prevent the discharge filamentation and to generate

more diffuse plasma.

The electrodes are powered by a custom-made high-voltage

ulse generator producing alternating polarity pulses with peak

oltage of up to U peak = 16 kV and pulse repetition rate of 20 Hz.

he alternating polarity pulse train is converted to a positive-

olarity, 10 Hz pulse train, essentially without the pulse waveform

istortion, by using ultra-fast high-voltage diodes (UF1007-T)

onnected in series between the high-voltage terminal of the

ulse generator and the actuator electrodes, as shown in Fig. 1 .

dditional diodes are placed between the high-voltage electrode

nd the 10 k � buffer resistor (see Fig. 1 ), to slow down the voltage

eduction on the electrode by several orders of magnitude. This

esults in adding a gradually decaying tail to the voltage pulse

hape, several ms long, used to enhance the electric field effect

n the flame. The discharge voltage and current waveforms are

easured by the Tektronix P-6015 high voltage probe (bandwidth

5 MHz) and Pearson 2877 current probe (bandwidth 200 MHz).

lasma emission images are taken by the Princeton Instruments

I-Max 3 ICCD camera with a UV lens. Figure 2 plots the baseline

ischarge pulse voltage and current waveforms measured without

he additional diodes between the high-voltage electrode and

he buffer resistor, plotted together with the capacitive current,

alculated using the measured stray capacitance of the electrodes,

stray = 0.44 pF. The flame has a negligible effect on the stray

apacitance. It can be seen that the discharge current exceeds the

apacitive current by almost an order of magnitude. The labels

Page 3: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248 241

Fig. 1. Schematic of the counterflow burner, double dielectric barrier discharge electrode assembly, electric circuit, and the laser beam (a,b). Photos of the burner with the

electrodes in place and the flame, (c) axial view and (d) side view.

Fig. 2. Pulse voltage and current waveforms measured without additional diodes

between the high-voltage electrode and the buffer resistor, plotted together with

the capacitive current. “Breakdown”, “peak”, and “current decay” indicate the mo-

ments when the electric field distributions along the discharge gap were measured

(see Fig. 9 (b)).

w

m

d

t

h

a

c

p

t

l

t

t

τ

e

s

i

t

i

t

P

e

c

b

b

r

a

m

i

F

t

t

b

t

s

p

t

t

o

a

b

k

e

T

i

e

t

s

3

e

l

breakdown”, “peak”, and “current decay” indicate the moments

hen the electric field distributions along the discharge gap were

easured (see the discussion of Fig. 9 below).

Figure 3 compares the pulse voltage waveforms measured with

ifferent number of diodes between the high-voltage electrode and

he buffer resistor ( n = 0–6). As expected, adding the extra diodes

as essentially no effect on the pulse shape during the voltage rise

nd even during the voltage reduction, as long as the discharge

urrent after breakdown remains significant. Basically, the voltage

ulse shape changes only during the voltage reduction, after

he conductivity of the discharge gap is reduced significantly be-

ow the peak value during the discharge pulse. At these conditions,

he effective buffer resistance becomes extremely high (estimated

o be R b ∼ 1 GOhm), resulting in a very long voltage decay time,

decay ∼ R b C stray ∼ 1 ms. Thus, adding the diodes increases the

ffective buffer resistance after the conduction current decay by

everal orders of magnitudes, compared to the baseline case, and

ncreases the voltage fall time from a few hundred nanoseconds

o several milliseconds.

The schematic of ps E-FISH laser diagnostics, shown in Fig. 4 ,

s essentially the same as in our previous work [44,45] . Briefly,

he fundamental, vertically polarized output beam of an Ekspla

L2143A Nd:YAG laser, with a pulse duration of 30 ps and pulse

nergy of 10 mJ, operating at 10 Hz, is focused into the dis-

harge/flame region, using a 100 cm focal distance lens. The laser

eam diameter at the focal point, measured by traversing a razor

lade across the beam, is approximately 200 μm, with the Rayleigh

ange of about z R ≈ 3 cm. The focal point of the beam is placed

t the center of the discharge electrode assembly. The second har-

onic signal beam is separated from the fundamental beam us-

ng a pair of dichroic mirrors and a dispersion prism, as shown in

ig. 4 . The signal beam is recollimated and focused onto the en-

rance slit of a monochromator, followed by a narrowband pass fil-

er (50% transmission at 532 nm, 10 nm band pass), and detected

y a photomultiplier tube (PMT). A polarizer mounted on a rota-

ion stage before the focusing lens isolates the second harmonic

ignals generated by the vertical and horizontal electric field com-

onents, since the signal polarization is parallel to the direction of

he field. At the present conditions, the root mean square value of

he electric field, averaged nearly uniformly over the span of the

verlapping electrodes [45] (see Fig. 1 ), is measured. The timing

nd the pulse energy of the fundamental laser beam are monitored

y a photodiode.

Absolute calibration of the measurements is obtained from the

nown electrostatic electric field generated halfway between the

lectrodes during the ns pulse voltage rise, before breakdown.

he electrostatic electric field distribution is calculated by solv-

ng the Laplace equation for the electric potential, for the given

lectrode geometry. During the calibration and measurements in

he discharge, the PMT voltage is kept sufficiently low to avoid its

aturation at high electric fields [44] .

. Results and discussion

Figure 5 shows a set of single-shot, 10-ns camera gate plasma

mission images taken during the discharge pulse, as well as the

ong gate images (400 ns showing the entire discharge pulse, and

Page 4: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

242 Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248

Fig. 3. Pulse voltage and current waveforms measured with different number of diodes between the high-voltage electrode and the ballast resistor ( n = 0–6). Adding the

diodes increases the effective buffer resistance by several orders of magnitudes, compared to the baseline case, and increases the voltage fall time.

Fig. 4. Schematic of picosecond Electric Field Induced Second Harmonic (E-FISH) diagnostic.

Fig. 5. Single-shot, 10-ns camera gate plasma emission images taken during the discharge pulse and long gate images (400 ns showing the entire discharge pulse, and 100 μs

showing the flame) in the positive polarity “baseline” discharge ( n = 0) sustained across the flame plane, axial view (left) and side view (right). The locations of the electrodes

are indicated with dashed lines.

i

fi

F

c

t

i

n

d

i

c

100 μs showing the flame). In Fig. 5 and all subsequent figures,

t = 0 represents the moment when the applied voltage peaks, as

indicated in Fig. 2 , and the time stamps indicate the moments

when the camera gate opens. These images are taken in the posi-

tive polarity “baseline” discharge sustained across the flame plane,

without the extra diodes between the high voltage electrode and

the buffer resistor (i.e. at n = 0). Both the axial view (along the

electrodes) and the side view are shown. The locations of the

electrodes in the images are indicated with dashed lines. It can

be seen that the discharge generates a relatively diffuse plasma

n the plane of the electrodes and across the flame, although the

lamentary structure on the bottom (fuel) side is readily apparent.

igure 6 shows a set of images taken in a positive polarity dis-

harge with four diodes between the high voltage electrode and

he buffer resistor ( n = 4). Comparison of the two sets of plasma

mages ( n = 0 and n = 4) shows that the plasma produced in an

s pulse discharge with a tail several ms long appears more

iffuse, with no sign of well-defined individual filaments detected

n the baseline ns pulse discharge. This is confirmed by further

omparison of the emission images taken with a long camera gate,

Page 5: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248 243

Fig. 6. Single-shot, 10-ns gate plasma emission images, and long gate images

(400 ns showing the entire discharge pulse, and 100 μs showing the flame) in a

positive polarity discharge with n = 4.

Fig. 7. 400 ns gate plasma emission images, showing the entire discharge pulse, and

100 μs gate images showing the flame, in positive polarity discharges with n = 0–6.

4

d

i

t

n

i

s

d

f

t

t

l

o

e

t

t

m

w

fi

i

t

s

v

o

t

i

b

n

v

a

T

fi

i

s

f

c

t

fi

t

m

b

a

s

t

v

c

t

o

e

t

a

f

o

c

t

t

d

i

d

t

c

2

(

d

i

o

e

n

p

r

c

e

d

c

i

F

F

00 ns (showing the plasma emission integrated over the entire

ischarge pulse), and 100 μs (also showing the flame), for n = 0–6,

llustrated in Fig. 7 . Since adding the extra diodes does not change

he voltage waveform during the discharge pulse, this behavior is

ot completely understood, but it is likely that the electron and

on transport by the electric field in the pulse tail, on an ms time

cale, helps reduce the nonuniformity of the residual electron

ensity distribution between the pulses, precluding the filament

ormation by the subsequent voltage pulse.

Figure 8 (a) plots the calculated Laplacian field distribution be-

ween two parallel rod electrodes in alumina ceramic sleeves for

he applied voltage of 14.5 kV, which was used for the abso-

ute calibration of the measurements. For a lower peak voltage

f 8 kV, when no plasma is generated in the gap, the measured

lectric field follows the Laplacian electric field during the en-

ire voltage pulse, as illustrated in Fig. 8 (b). Figure 9 (a) shows the

ime-dependent Laplacian field, calculated by multiplying the nor-

alized solution of the Laplace equation by the applied voltage

aveform, and the time-resolved vertical component of the electric

eld in the baseline ns pulse discharge ( n = 0) sustained at 10 Hz

n the counterflow flame, plotted at three different locations be-

ween the electrodes, indicated schematically in the inset. It can be

een that the use of the Laplacian field for calibration during the

oltage rise, before breakdown, is justified since the electric field

ffset before the discharge pulse is typically small, with the excep-

ion of the measurements near the grounded electrode, where it

ntroduces an estimated 10% uncertainty in the calibration. After

reakdown, the electric field at all three locations is reduced sig-

ificantly due to the plasma self-shielding, as observed in our pre-

ious measurements of the electric field in ns pulse discharges in

ir and hydrogen, as well as in a hydrogen diffusion flame [44,45] .

he data taken before and during the voltage rise (specifically, the

eld behavior at z = 1 mm, −200 ns < t < −50 ns, see Fig. 9 (a))

ndicate the electric field reversal between the discharge pulses,

uch as has been detected in our previous measurements in a sur-

ace plasma flow actuator [44] . The field reversal is caused by the

harge accumulation on the dielectric surfaces covering the elec-

rodes, which may persist for a long time after the applied electric

eld is removed [44] . At the present conditions, this effect is rela-

ively weak, except near the grounded electrode.

Figure 9 (b) shows the vertical component of the electric field

easured at different locations across the discharge gap near the

reakdown moment at z = 0 (at t = −20 ns), as well as at peak volt-

ge (at t = 0 ns) and after the discharge current decay (at t = 20 ns,

ee Fig. 2 ), at the conditions of Fig. 9 (a). The spatial distribution of

he vertical component of the Laplacian field for the peak applied

oltage of U peak = 14.5 kV used in these measurements is shown for

omparison. As expected, the electric field near the grounded elec-

rode remains significantly higher compared to that in the middle

f the gap, although at all locations across the discharge gap, the

lectric field after breakdown is significantly lower compared to

he “nominal” Laplacian field corresponding to the applied voltage

t that moment of time.

Comparison of the temporal variation of the electric field at dif-

erent spatial locations, shown in Fig. 9 (a), and the spatial variation

f the field at different moments of time, plotted in Fig. 9 (b), indi-

ates that the electric field varies rapidly in time during and after

he discharge pulse, on the time scale of tens to hundreds of ns. On

he other hand, the electric field distribution in the gap during the

ischarge pulse remains smooth and does not exhibit well-defined

solated maxima indicative of ionization waves propagating in the

ischarge gap, such as detected in our previous work [47] .

Figure 10 plots the time-dependent Laplacian field and the

ime-resolved electric field measured in the baseline ns pulse dis-

harge operated in burst mode, at the pulse repetition rate of

50 Hz, burst repetition rate of 10 Hz, and with 10 pulses in a burst

burst duration 40 ms). The data shown in the figure are measured

uring the pulse #6, i.e. at t = 20 ms. Although these data are qual-

tatively similar to the results obtained in the baseline discharge

perated at 10 Hz, it is readily apparent that the deviation of the

lectric field after breakdown from the Laplacian field is less pro-

ounced. This appears somewhat counterintuitive, both since the

ulse repetition rate during the burst remains low (such that the

esidual ionization from the previous pulse is likely to be insignifi-

ant), and since higher residual ionization would only increase the

ffect of plasma self-shielding [48] . This nature of this effect is un-

erstood better when the flame oscillations induced by the dis-

harge are taken into account, as discussed below.

Figure 11 shows the results of the electric field measurements

n a 10 Hz ns pulse discharge with n = 4, with a long tail (2 ms

WHM) at the trailing edge of the voltage pulse waveform (see

ig. 3 ). At these conditions, the deviation of the measured field

Page 6: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

244 Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248

Fig. 8. (a) Calculated Laplacian field distribution between two parallel rod electrodes in dielectric sleeves for the applied voltage of 14.5 kV, used for absolute calibration and

(b) time-dependent Laplacian field (curve) and measured time-resolved electric field (symbols) halfway between the electrodes for a lower peak applied voltage of 8 kV.

Fig. 9. (a) Time-dependent Laplacian field (curves) and measured time-resolved electric field (symbols) in the baseline ns pulse discharge sustained at 10 Hz in the counter-

flow flame, at three different locations between the electrodes indicated in the inset and (b) Vertical electric field measured at different locations along the discharge gap

near breakdown at z = 0 ( t = −20 ns), at peak voltage ( t = 0 ns), and after discharge current decay ( t = 20 ns, see Fig. 2 ). Spatial distribution of the vertical component of the

Laplacian field for peak applied voltage of U peak = 14.5 kV is shown for comparison. A schematic of the discharge pulse train is also shown.

A

d

1

i

t

t

a

o

t

F

t

w

t

s

o

t

z

t

from the Laplacian field is even less pronounced, which suggests

that the electron density in the plasma may be lower compared

to the baseline 10 Hz pulse discharge. This effect may be due to

suppression of the filaments detected in the baseline discharge

plasma, when the extra diodes are added to the external circuit

to produce the long tail in the voltage waveform, as illustrated in

Fig. 7 . It can also be seen that adding the tail generates a fairly

significant electric field after the discharge pulse, in the range of

2–5 kV/cm at the measurement locations, on the time scale of sev-

eral hundred ns.

Figure 12 compares the time-dependent Laplacian field with the

absolute value of the electric field measured in the gradually de-

caying tail of the applied voltage pulse measured at different loca-

tions in the discharge gap at the conditions of Fig. 11 ( n = 4), on

a longer time scale, up to 1 ms. Although the electric field in the

gap is significantly lower compared to the Laplacian field, it is still

detectable up to at least 1 ms, and is consistently higher compared

to the field after the baseline ns pulse discharge without the tail.

lso, the field near the grounded electrode, ≈ 1 kV/cm, is detected

uring the entire time period between the discharge pulses, 0.1–

00 ms. Note that since the field reversal moment is difficult to

dentify from the relatively sparse data between the pulses, only

he absolute value of the electric field is plotted in Fig. 12 . To de-

ermine whether the electric field of this magnitude may produce

n effect on the flame, sub-breakdown DC voltage in the range

f 0.5–3.0 kV/cm was applied to the electrodes, resulting in a de-

ectable displacement of the flame above ∼1 kV/cm. The data in

igs. 9–12 illustrate and quantify the difference between the elec-

ric field in the gap after the discharge pulse, measured with and

ithout the tail in the voltage waveform. The detection limit of

he present diagnostics is approximately 0.5 kV/cm, limited by the

tray second harmonic signal generated in the components of the

ptical system, which was subtracted from the signal generated in

he plasma. As expected, the highest electric field is measured at

= 5 mm, near the grounded electrode. At this location, the elec-

ric field above the sensitivity limit, approximately 1–3 kV/cm, is

Page 7: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248 245

Fig. 10. Time-dependent Laplacian field (curves) and measured time-resolved elec-

tric field (symbols) in the baseline ns pulse discharge operated in burst mode (pulse

repetition rate 250 Hz, 10 pulses in a burst, burst repetition rate 10 Hz, data shown

after a pulse #6). A schematic of the discharge pulse train is also shown.

Fig. 11. Time-dependent Laplacian field (curves) and measured time-resolved elec-

tric field (symbols), in an ns pulse discharge with n = 4 at 10 Hz. A schematic of the

discharge pulse train is also shown.

d

u

v

l

fl

a

i

c

p

c

Fig. 12. Time-dependent Laplacian field (curves) and the absolute value of the

time-resolved electric field (symbols) in the gradually decaying tail of the applied

voltage waveform at different locations along the centerline. The conditions are the

same as in Fig. 11 ( n = 4).

Fig. 13. Flame emission images illustrating flame oscillations excited by the base-

line, positive polarity, ns pulse discharge operated at 10 Hz (at the conditions of

Fig. 9 ). Electrode gap 12 mm, camera gate 1 ms.

Fig. 14. Flame emission images illustrating flame oscillations excited by the base-

line, positive polarity, ns pulse discharge operated in burst mode, at the conditions

of Fig. 10 (pulse repetition rate 250 Hz, 10 pulses in a burst, burst repetition rate

10 Hz). Electrode gap 12 mm, camera gate 1 ms.

s

2

r

t

t

t

c

etected during the entire interval between the discharge pulses,

p to t = 100 ms (e.g. see Figs. 11,12 ), when the applied voltage is

ery low, indicating the effect of residual surface charge accumu-

ation on the ceramic sleeve covering the grounded electrode.

Figure 13 shows a set of flame emission images, illustrating

ame oscillations excited by the baseline ns pulse discharge oper-

ted at 10 Hz (at the conditions of Fig. 9 ), taken during the 100 ms

nterval between the discharge pulses. It can be seen that the dis-

harge induces the flame distortion and oscillations, with the am-

litude of approximately 1–2 mm, on the time scale much longer

ompared to the discharge pulse duration ( ∼10 ms vs. ∼100 ns). As

hown in Fig. 14 , operating the discharge in 10 Hz burst mode, at

50 Hz pulse repetition rate and 10 pulses per burst (burst du-

ation rate of 40 ms), i.e. at the conditions of Fig. 10 , enhances

he effect on the flame considerably. In this case, it is apparent

hat the flame becomes strongly distorted near the location of

he plasma during the discharge burst (at t = 0–40 ms), with the

entral part moving toward the grounded electrode. As discussed

Page 8: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

246 Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248

Fig. 15. Flame emission images comparing flame oscillations excited by an ns pulse discharge operated at 10 Hz, with and without tail added to the voltage waveform: (a)

n = 0; (b) n = 2; (c) n = 4 and (d) n = 6. Electrode gap 12 mm, camera gate 1 ms.

d

e

F

c

F

r

b

b

i

t

o

e

s

t

o

in Refs. [15,38] , the body force induced by the applied electric

field, enhanced by the ionization generated during the discharge

at the present conditions, affects the bi-directional transport (dif-

fusion and convection) of the fuel and oxidizer species in the dis-

charge gap, resulting in the displacement of the counterflow dif-

fusion flame. Since the electric field and electron density distribu-

tions between the parallel rod electrodes is not uniform, the cen-

tral part of the flame, overlapping with the plasma, is moving to-

ward the grounded electrode during the discharge burst (at t = 0–

40 ms), such that the flame becomes distorted.

After the discharge burst, the central part of the flame returns

to near original position (at 60 ms), while the peripheral region

of the flame moves out of phase with the central part (at 70–

80 ms), resulting in a “flapping” motion (see Fig. 14 ). The flame

displacement from the stationary position (without the discharge)

is likely responsible for a somewhat weaker apparent effect of the

ischarge on the electric field after breakdown for burst mode op-

ration, at t = 20 ms, compared to the 10 Hz discharge (compare

igs. 9 (a) and 10 )). In this case, the electric field during the dis-

harge pulses in the middle of the burst (such as pulse #6 in

ig. 10 ) is measured in a lower temperature fuel (CH 4 –Ar) mixture,

ather than in the higher temperature reaction zone, such that the

reakdown voltage is higher and the electron density is likely to

e lower.

Finally, Fig. 15 compares several sets of flame emission images

n an ns pulse discharge operated at 10 Hz, with and without long

ail added to the voltage waveform, taken at the same moments

f time. The results in Fig. 15 are shown for different number of

xtra diodes between the high-voltage electrode and the buffer re-

istor, n = 0 (baseline case, conditions of Fig. 9 ), n = 2, n = 4 (condi-

ions of Fig. 11 ), and n = 6. It is readily apparent that the amplitude

f the flame oscillations for n = 4 and n = 6 is significantly higher

Page 9: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248 247

c

f

o

t

l

s

b

d

r

s

a

d

t

n

b

2

g

v

c

i

c

t

μ

o

fl

t

d

fl

t

i

4

c

t

d

e

t

o

f

b

r

o

a

i

p

t

c

o

m

i

s

t

w

a

f

r

b

w

s

f

e

m

fi

m

l

T

d

f

I

d

d

[

t

a

e

r

t

a

t

a

a

t

p

d

a

j

d

e

p

A

D

u

S

e

R

ompared with that at n = 0 and n = 2, demonstrating a strong ef-

ect of the residual electric field during the voltage pulse tail. The

scillation amplitudes at n = 4 and n = 6 are close, indicating that

he effect is near saturation, such that further increase of the tail

ength would not result in an additional increase of the amplitude.

In the present work, the frequency of the flapping motion is the

ame as the forcing frequency (discharge pulse repetition rate or

urst repetition rate of 10 Hz in Figs. 13–15 ), at all operating con-

itions. This is readily apparent both from the videos of the flame

esponse to the forcing by the discharge, and from ICCD images

uch as shown in Figs. 13–15 . Specifically, the flame images taken

t a fixed time delay after the discharge pulse have excellent repro-

ucibility, which illustrates that the flame oscillation frequency is

he same as the forcing frequency. This fact is confirmed by the dy-

amic response of the flame between the discharge pulses (pulse

ursts), illustrated in Figs. 13–15 .

Increasing the discharge pulse repetition rate from 10 Hz to

0 Hz, when the forcing frequency becomes comparable with the

lobal stretch rate of the flame, estimated as the ratio of the flow

elocity and the nozzle gap, a = U 0 / �1 = 17.7 1/s, reduces the os-

illation amplitude. Note that adding the extra diodes and extend-

ng the voltage waveform tail does not change the pulse energy

oupled to the plasma. The estimated upper bound difference in

he energy coupled with and without the tail is several tens of

J/pulse, much smaller compared to the baseline coupled energy

f 0.5 mJ/pulse. Therefore at these conditions the effect on the

ame is almost certainly due to the electrohydrodynamic interac-

ion (“ion wind”). Basically, no additional ionization is generated

uring the voltage waveform tail, and the enhanced effect of the

ame is produced by the transport of the ions generated during

he discharge pulse by the electric field on a longer time scale, thus

ncreasing the impulse of the EHD force.

. Summary

In the present work, oscillations of an atmospheric pressure,

ounterflow CH 4 –O 2 –Ar flame are induced by repetitive, ns dura-

ion, high voltage pulses combined with a gradually decaying, ms

uration tail. The electric field during the ns pulses, which gen-

rate a plasma between the electrodes, and in the afterglow be-

ween the pulses, is measured by ps Electric Field Induced Sec-

nd Harmonic (E-FISH) diagnostic. Absolute calibration is obtained

rom the second harmonic signal measured during the voltage rise,

efore breakdown, when the electric field between the electrodes

emains Laplacian. The results show that ns discharge pulses with-

ut the tail produce low-amplitude flame oscillations when oper-

ted at a repetition rate of 10 Hz. The amplitude of the oscillations

ncreases when the discharge is operated in burst mode, with 10

ulses per burst and the same burst repetition rate of 10 Hz. Al-

hough the electric field between the discharge pulses at these

onditions is low, near the detection limit, the long time scale

f the flame oscillations, ∼10 ms, suggests that they are induced

ainly by the residual EHD force on the charged species generated

n the discharge, rather than by the chemical reactions of radical

pecies generated in the plasma or the Joule heat produced during

he discharge.

Adding a variable, ms duration, tail to the ns pulse voltage

aveform increases the flame oscillation amplitude considerably,

s the tail duration is increased. Since the discharge pulse wave-

orm, during the period when the conduction current is detected,

emains essentially the same, this effect is almost certainly caused

y the electric field in the tail, producing the EHD force (“ion

ind”) on the charges generated during the discharge pulse. Ba-

ically, adding the tail greatly increases the impulse of the EHD

orce, without producing additional charges or increasing the en-

rgy input in the plasma. This is consistent with the electric field

easurements after the discharge pulse, showing that the electric

eld in the plasma persists during the voltage tail, although it re-

ains lower compared to the Laplacian field, due to the accumu-

ation of charges on the dielectric sleeves covering the electrodes.

he characteristic time scale of the flame oscillations is of the or-

er of ∼10 ms, controlled by the relatively slow momentum trans-

er from the ions in the decaying plasma to the neutral species.

t may be reduced by increasing the pulse peak voltage while re-

ucing the pulse duration, which would increase the peak electron

ensity in the plasma without inducing the discharge instability

49] , and by operating the ns pulse discharge in burst mode. At

he present conditions, generating higher frequency oscillations is

lso limited by the low flame stretch rate. However, this factor is

xpected to be less restrictive at a higher flow velocity through the

eaction zone.

The sensitivity of the experimental observations (in particular

he enhancement of the flame oscillations amplitude vs. the volt-

ge pulse duration) to the equivalence ratio remains an open ques-

ion, since it may strongly depend on the peak electron density

nd the rate of electron recombination at the location of the flame,

nd will be the subject of the future study. The electrode geome-

ry can be optimized to enhance the residual electric field com-

onent in the desired direction. Based on the experimental results

emonstrating the effect of dielectric barrier ns pulse discharges,

s well as sub-breakdown DC and AC fields, on stability of a lifted

et flame [22] , as well as on the present data, combining a ns pulse

ischarge voltage waveform with a long tail enhancing the EHD

ffect may be used for efficient flameholding in high-speed, high-

ressure flows.

cknowledgments

This material is based in part upon work supported by the

epartment of Energy , National Nuclear Security Administration ,

nder Award Number DE-NA0 0 02374 . The support of the China

cholarship Council (No. 201706210253 ) is gratefully acknowl-

dged.

eferences

[1] J. Lawton , F.J. Weinberg , Electrical aspects of combustion, Oxford University

Press, 1970 . [2] A.B. Fialkov , Investigations on ions in flames, Prog. Energy Combust. Sci. 23

(1997) 399–528 .

[3] S.M. Lee , C.S. Park , M. Cha , S. Chung , Effect of electric fields on the liftoffof non-premixed turbulent jet flames, IEEE Trans. Plasma Sci. 33 (2005)

1703–1709 . [4] S. Marcum , B. Ganguly , Electric-field-induced flame speed modification, Com-

bust. Flame 143 (2005) 27–36 . [5] D. Wisman , S. Marcum , B. Ganguly , Electrical control of the thermodiffusive in-

stability in premixed propane-air flames, Combust. Flame 151 (2007) 639–648 .

[6] M.K. Kim , S.K. Ryu , S.H. Won , S.H. Chung , Electric fields effect on liftoff andblowoff of non-premixed laminar jet flames in a coflow, Combust. Flame 157

(2010) 17–24 . [7] A. Cessou , E. Varea , K. Criner , G. Godard , P. Vervisch , Simultaneous measure-

ments of OH, mixture fraction and velocity fields to investigate flame stabi-lization enhancement by electric field, Exp. Fluids 52 (2012) 905–917 .

[8] A. Drews , L. Cademartiri , M. Chemama , M. Brenner , G. Whitesides , K. Bishop ,

AC electric fields drive steady flows in flames, Phys. Rev. E 86 (2012) 036314 . [9] E.N. Volkov , V.N. Kornilov , L.P.H. de Goey , Experimental evaluation of DC elec-

tric field effect on the thermoacoustic behaviour of flat premixed flames, Proc.Combust. Inst. 34 (2013) 955–962 .

[10] J. Kuhl , G. Jovicic , L. Zigan , A. Leipertz , Transient electric field response of lam-inar premixed flames, Proc. Combust. Inst. 34 (2013) 3303–3310 .

[11] J. Kuhl , G. Jovicic , L. Zigan , S. Will , A. Leipertz , Influence of electric fieldson premixed laminar flames: visualization of perturbations and potential for

suppression of thermoacoustic oscillations, Proc. Combust. Inst. 33 (2015)

3521–3528 . [12] S. Karnani , D. Dunn-Rankin , Detailed characterization of dc electric field effects

on small non-premixed flames, Combust. Flame 162 (2015) 2865–2872 . [13] C. Guerra-Garcia , M. Martinez-Sanchez , Counterflow nonpremixed flame DC

displacement under AC electric field, Combust. Flame 162 (2015) 4254–4263 .

Page 10: Combustion and Flame - Non-Equilibrium Thermodynamics ......electron / ion densities, up to at least n + 1b∼ 10 14 –10 15 cm −3 [40]. However, the slowly varying residual electric

248 Y. Tang, M. Simeni Simeni and K. Frederickson et al. / Combustion and Flame 206 (2019) 239–248

[

[14] Y. Xiong , M. Cha , S. Chung , AC electric field induced vortex in laminar coflowdiffusion flames, Proc. Combust. Inst. 35 (2015) 3513–3520 .

[15] D. Park , S. Chung , M. Cha , Bidirectional ionic wind in non-premixed counter-flow flames with dc electric fields, Combust. Flame 168 (2016) 138–146 .

[16] Y. Xiong , D. Park , B. Lee , S. Chung , M. Cha , DC field response of one-di-mensional flames using an ionized layer model, Combust. Flame 163 (2016)

317–325 . [17] D.G. Park , S. Chung , M. Cha , Visualization of ionic wind in laminar jet flames,

Combust. Flame 184 (2017) 246–248 .

[18] Y. Ren , S. Li , W. Cui , Y. Zhang , L. Ma , Low-frequency AC electric field inducedthermoacoustic oscillation of a premixed stagnation flame, Combust. Flame

176 (2017) 479–488 . [19] Y. Ren , W. Cui , S. Li , Electrohydrodynamic instability of premixed flames under

manipulations of dc electric fields, Phys. Rev. E 97 (2018) 013103 . [20] K. Criner , A. Cessou , J. Louiche , P. Vervisch , Stabilization of turbulent lifted jet

flames assisted by pulsed high voltage discharge, Combust. Flame 144 (2006)

422–425 . [21] G. Pilla , D. Galley , D.A. Lacoste , F. Lacas , D. Veynante , C.O. Laux , Stabilization

of a turbulent premixed flame using a nanosecond repetitively pulsed plasma,IEEE Trans. Plasma Sci. 34 (2006) 2471–2477 .

[22] K. Criner , A. Cessou , P. Vervisch , A comparative study of the stabilization ofpropane lifted jet-flames by pulsed, AC and DC high-voltage discharges, Third

European Combustion Meeting, ECM, Chania, Crete. April (2007) 11–13 .

[23] A.Y. Starikovskii , N.B. Anikin , I.N. Kosarev , E.I. Mintoussov , M.M. Nud-nova , A.E. Rakitin , D.V. Roupassov , S.M. Starikovskaia , V.P. Zhukov , Nanosec-

ond-pulsed discharges for plasma-assisted combustion and aerodynamics, J.Propuls. Power 24 (2008) 1182–1197 .

[24] W. Kim , H. Do , M.G. Mungal , M.A. Cappelli , Plasma-discharge stabilization ofjet diffusion flames, IEEE Trans. Plasma Sci. 34 (2006) 2545–2551 .

[25] W. Kim , M.G. Mungal , M.A. Cappelli , The role of in situ reforming in plasma

enhanced ultra-lean premixed methane/air flames, Combust. Flame 157 (2010)374–383 .

[26] M.S. Bak , S.K. Im , M.G. Mungal , M.A. Cappelli , Studies on the stability limit ex-tension of premixed and jet diffusion flames of methane, ethane, and propane

using nanosecond repetitive pulsed discharge plasmas, Combust. Flame 160(2013) 2396–2403 .

[27] T. Ombrello , S. Won , Y. Ju , S. Williams , Flame propagation enhancement by

plasma excitation of oxygen. Part I: effects of O 3 , Combust. Flame 157 (2010)1906–1915 .

[28] T. Ombrello , S. Won , Y. Ju , S. Williams , Flame propagation enhancement byplasma excitation of oxygen. Part II: effects of O 2 (a

1 �g ), Combust. Flame 157

(2010) 1916–1928 . [29] W. Sun , M. Uddi , S.H. Won , T. Ombrello , C. Carter , Y. Ju , Kinetic effects of

non-equilibrium plasma-assisted methane oxidation on diffusion flame extinc-

tion limits, Combust. Flame 159 (2012) 221–229 . [30] D.A . Lacoste , D.A . Xua , P. Moeck , C.O. Laux , Dynamic response of a weakly

turbulent lean-premixed flame to nanosecond repetitively pulsed discharges,Proc. Combust. Inst. 34 (2013) 3259–3266 .

[31] D.A. Lacoste , J.P. Moeck , D. Durox , C.O. Laux , T. Schuller , Effect of nanosecondrepetitively pulsed discharges on the dynamics of a swirl-stabilized lean pre-

mixed flame, J. Eng. Gas Turbines Power 135 (2013) 101501 . [32] W. Kim , J. Snyder , J. Cohen , Plasma assisted combustor dynamics control, Proc.

Combust. Inst. 35 (2015) 3479–3486 .

[33] D.A. Lacoste , Y. Xiong , J.P. Moeck , S.H. Chung , W.L. Roberts , M.S. Cha , Transferfunctions of laminar premixed flames subjected to forcing by acoustic waves,

AC electric fields, and non-thermal plasma discharges, Proc. Combust. Inst. 36(2017) 4183–4192 .

[34] M. Belhi , P. Domingo , P. Vervisch , Direct numerical simulation of the effect ofan electric field on flame stability, Combust. Flame 157 (2010) 2286–2297 .

[35] M. Belhi , P. Domingo , P. Vervisch , Modeling of the effect of DC and AC electricfields on the stability of a lifted diffusion methane/air flame, Combust. Theory

Model. 17 (2013) 749–787 .

[36] J. Hu , B. Rivin , E. Sher , The effect of an electric field on the shape of co-flowingand candle-type methane-air flames, Exp. Therm. Fluid Sci. 21 (20 0 0) 124–133 .

[37] K. Yamashita , S. Karnani , D. Dunn-Rankin , Numerical prediction of ion currentfrom a small methane jet flame, Combust. Flame 156 (2009) 1227–1233 .

[38] M. Belhi , B.J. Lee , F. Bisetti , H.G. Im , A computational study of the effects ofDC electric fields on non-premixed counterflow methane-air flames, J. Phys. D

Appl. Phys. 50 (2017) 494005 .

[39] S.O. Macheret , M.N. Shneider , R.B. Miles , Magnetohydrodynamic and electro-hydrodynamic control of hypersonic flows of weakly ionized plasmas, AIAA J.

42 (2004) 1378–1387 . [40] A. Roettgen , I. Shkurenkov , M. Simeni Simeni , I.V. Adamovich , W.R. Lempert ,

Time-resolved electron temperature and electron density measurements in ananosecond pulse filament discharge in H 2 -He and O 2 -He mixtures, Plasma

Sources Sci. Technol. 25 (2016) 055008 .

[41] S.B. Leonov , V. Petrishchev , I.V. Adamovich , Dynamics of energy coupling andthermalization in barrier discharges over dielectric and weakly conducting sur-

faces on μsec to msec time scales, J. Phys. D Appl. Phys. 47 (2014) 465201 . [42] A. Dogariu , B.M. Goldberg , S. O’Byrne , R.B. Miles , Species-independent fem-

tosecond localized electric field measurement, Phys. Rev. Appl. 7 (2017)024024 .

[43] B.M. Goldberg , T.L. Chng , A. Dogariu , R.B. Miles , Electric field measurements

in a near atmospheric pressure nanosecond pulse discharge with picosecondelectric field induced second harmonic generation, Appl. Phys. Lett. 112 (2018)

064102 . 44] M. Simeni Simeni , Y. Tang , K. Frederickson , I.V. Adamovich , Electric field distri-

bution in a surface plasma flow actuator powered by ns discharge pulse trains,Plasma Sources Sci. Technol. 27 (2018) 104001 .

[45] M. Simeni Simeni , Y. Tang , Y.-C. Hung , Z. Eckert , K. Frederickson ,

I.V. Adamovich , Electric field in ns pulse and AC electric discharges in ahydrogen diffusion flame, Combust. Flame 197 (2018) 254–264 .

[46] M. Frenklach, H. Wang, C.-L. Yu, M. Goldenberg, C.T. Bowman, R.K. Hanson,D.F. Davidson, E.J. Chang, G.P. Smith, D.M. Golden, W.C. Gardiner, V. Lissianski,

1995, http://combustion.berkeley.edu/gri _ mech/new21/version12/text12.html . [47] T.L. Chng , I.S. Orel , S.M. Starikovskaia , I.V. Adamovich , Electric Field Induced

Second Harmonic (E-FISH) generation for characterization of fast ionization

wave discharges at moderate and low pressures, Plasma Sources Sci. Technol.28 (2019) 045004 .

[48] K. Takashima , I.V. Adamovich , U. Czarnetzki , D. Luggenhölscher , Developmentof fast ionization wave discharges at high pulse repetition rates, Plasma Chem.

Plasma Process. 32 (2012) 471–493 . [49] K. Takashima , Z. Yin , I.V. Adamovich , Measurements and kinetic modeling of

energy coupling in volume and surface nanosecond pulse discharges, PlasmaSources Sci. Technol. 22 (2013) 015013 .


Recommended