You are on page 1of 32

ed

Effects of non-equilibrium plasma and spark discharge on low-

temperature combustion in lean propane/air mixtures

iew
Yangyang Ban1, Shenghui Zhong1,2, Jiajian Zhu3, Fan Zhang1*

ev
1State
Key Laboratory of Engines, Tianjin University, Tianjin 300350, China
2Beihang Hangzhou Innovation Institute Yuhang, Xixi Octagon City, Yuhang District, Hangzhou
310023, China
3Science and Technology on Scramjet Laboratory, College of Aerospace Science and

r
Engineering, National University of Defense Technology, Hunan 410073, China

Abstract
er
A zero-dimensional (0D) non-equilibrium plasma assisted ignition/combustion
pe
(PAI/PAC) model is established by coupling plasma kinetics and combustion kinetics

to investigate the effect of non-equilibrium nanosecond repetitively pulsed discharge

(NRPD) plasma and spark discharge on low-temperature combustion (LTC) stage of


ot

lean propane/air mixtures. The results show that the NRPD case has the fastest
tn

ignition, in comparison to spark ignition (SI) with the same discharge energy. Kinetic

pathways analysis for non-equilibrium PAI/PAC shows that active species induced by

plasma can significantly accelerate low-temperature combustion pathways while SI


rin

only provides thermal enhancement. With NRPD before LTC, electrons, O(1D), as

well as O and OH, are formed during discharge, promoting propane consumption
ep

dramatically. Besides, electron collision with propane can promote alkyl and alkenes

and further contribute to OH formation through low-temperature pathways. After the


Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
pulse, propyl oxidation into propyl peroxyl, electron attachment, and relaxation of

vibrationally excited states are the dominant pathways of heat release, leading to the

iew
continuous temperature rising after NRPD and the reduction of ignition delay time

(IDT). Finally, the negative temperature coefficient (NTC) performances under auto-

ignition, NRPD, and SI are investigated. It is found that SI shifts the NTC to the low-

ev
temperature zone because with SI mode, the input discharge energy can increase the

initial temperature (T0) immediately yielding an earlier NTC, while NRPD eliminates

NTC phenomenon. It is explained that on one hand, O and O(1D) radicals induced by

r
NRPD can promote NC3H7 formation, accelerating NC3H7O2 production, enhancing

er
low-temperature reaction pathways and leading to a weakened negative correlation

between the oxidation rate of NC3H7 and temperature. On the other hand, the higher
pe
concentrations of NC3H7 and HO2 radicals during discharge can result in an advanced

intermediate/high temperature reaction process.


ot

Keywords: Non-equilibrium plasma; Low temperature combustion; Plasma assisted


tn

combustion, Propane; Plasma/chemical kinetics


rin

*Corresponding author: fanzhang_lund@tju.edu.cn


ep

1. Introduction
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
Low-temperature combustion strategies in ICEs have attracted more attention due

to low emissions and high thermal efficiencies [1,2]. However, a lean burn mixture

iew
will lead to a longer IDT or even misfire which results in a limited engine operation

range [3]. With a high-energy spark, the ground electrode restrains the high-

temperature region and flame kernel size [4]. A significant part of discharge energy is

ev
lost by heat transfer to the surroundings [5,6], leading to a strict requirement for

minimum ignition energy (MIE). Non-equilibrium discharge has been used to stabilize

or stimulate ignition in a lean fuel-air mixture [7,8], which suggested the significant

r
advantages compared with a spark in ICEs [9,10]. For example, with NRPD, IDT was

er
three to four orders of magnitude shorter and approximately 13-17% narrower in

combustion duration than SI [9]. Also, the faster flame propagation [10] has been
pe
verified during NRPD in comparison to SI. However, how do non-equilibrium plasma

and spark discharge influence the ignition for a hydrocarbon fuel with an NTC regime
ot

in high pressure is still lacking.

During non-equilibrium discharge, a large proportion of discharge energy is used to


tn

accelerate electrons leading to higher electron temperature instead of converting into

heat. There are four major paths of PAI/PAC summarized in [7] including the thermal
rin

effect via electron elastic collision and gas heating, the kinetic effect brought by

various excited species and radicals produced during inelastic collisions, the diffusion
ep

transport enhancement by molecular dissociation into fragments and the convective

transport enhancement via ion wind. Numerous researches are devoted to revealing

the kinetic effect of non-equilibrium plasma on small hydrogen molecules [8,11-14].


Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
Uddi et al. [12] carried out a 0D simulation of plasma-assisted CH4/air during a single

pulse and found that atomic oxygen is the key species for PAC. Ju et al. [8] confirmed

iew
the importance of O(1D) in producing the initial radicals for PAC, and some plasma-

generated species i.e., O2(a1Δg) which can lead to flame propagation enhancement

[15]. The effect of non-equilibrium plasma on high-temperature ignition of small

ev
hydrocarbon fuels has been well-studied.

For LTC assisted by plasma, recently a few papers have been done, focusing on the

chemical kinetics of low-temperature pyrolysis and oxidation of methane/O2/He [15],

r
propane/O2/N2 [16], pentane/O2/He [17] and n-dodecane/O2/N2 [18]. Mao et al. [15]

er
numerically modeled the ignition enhancement of CH4/O2/He mixtures using a

detailed plasma mechanism. They found that NSD and DC hybrid plasmas can
pe
enhance ignition significantly, especially at low temperatures, because the combustion

chain branching process is slow and fewer radicals for ignition are produced. The
ot

induction process becomes more dependent on plasma activation. Meanwhile, they

suggested a further analysis of the rate constant at low temperatures which still shows
tn

a large discrepancy for low-temperature fuel oxidation involving large fuel molecules.

Regarding to LTC within the NTC regime, Filimonova [19] modeled the discharge
rin

impact on multistage ignition of C3H8/air mixture neglecting detailed plasma kinetics.

The results showed that discharge reduced the induction period of LTC stage by one
ep

order of magnitude due to the formation of hydroperoxide and O atoms promoted by

plasma. More recently, Filimonova et al. [20] further investigated the influence of

non-equilibrium discharge on LTC stage in the HCCI engine. The results suggested
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
that the optimal discharge initiation time is at the beginning of the low-temperature

heat release (LTHR) or intermediate-temperature heat release. It is concluded that a

iew
non-equilibrium discharge can control the combustion timing and accelerate ignition

if the plasma discharge is before LTHR. However, they ignore metastable states, ions,

and also the heat released from relaxation in non-equilibrium plasma. In a nutshell,

ev
few studies have been conducted to investigate propane/air ignition by non-

equilibrium plasma as well as comparison with conventional spark discharge.

In this paper, the discharge and combustion processes are simulated simultaneously

r
by coupling plasma kinetics solver ZDPlasKin and combustion solver Chemkin. The

er
non-equilibrium PAI/PAC mechanism of propane/air is established and validated.

With the detailed mechanism, the effects of non-equilibrium plasma and spark
pe
discharge on LTC of a lean propane/air mixture are investigated when the discharge

timings are before LTC stage. Reaction pathways and heat release rates of the plasma
ot

kinetics and LTC oxidation reactions are carried out to show the influences of

discharging timing on critical radicals and ignition enhancement. Finally, the effect of
tn

NRPD and SI on NTC is investigated and analyzed.


rin

2. Kinetic mechanism and numerical method

2.1 Kinetic mechanism


ep

At first, IDTs in the three oxidation mechanisms of C3H8/air are compared with

experiments under high pressure and different equivalence ratios, as shown in Figure
Pr

1. The right-most figure in Fig. 1 is conducted by San Diego mechanism including 58

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
species and 270 reactions which is highly recommended in [21]. The left-most one is

by Petersen mechanism involving 118 species and 663 reactions [22] and the middle

iew
is by Gokulakrishnan mechanism including 966 reactions and 136 species [23]. It is

found that three mechanisms can predict IDT well at high temperatures and show NTC

behavior. While Gokulakrishnan mechanism with detailed NOx chemistry at low

ev
temperatures provides a rather good comparison with experimental data and the

prediction is better than the others. Since it has been verified that NOx can enhance

combustion at low temperatures, and reduce IDT [23] and the enhancement effect is

r
even greater than that of atomic oxygen production [11].

100

10-1
Petersen
(2007) (2014)
er
Gokulakrishnan SanDiego
(2016)
pe
S118R663 S136R966 S58R270

10-2
IDT (s)

10-3

10-4  = 0.5, Exp  = 0.5, Exp  = 0.5, Exp


ot

 = 1.0, Exp  = 1.0, Exp  = 1.0, Exp


 = 2.0, Exp  = 2.0, Exp  = 2.0, Exp
10-5  = 0.5, Sim  = 0.5, Sim  = 0.5, Sim
 = 1.0, Sim  = 1.0, Sim  = 1.0, Sim
 = 2.0, Sim  = 2.0, Sim  = 2.0, Sim
tn

10-6
0.6 0.9 1.2 1.5 0.6 0.9 1.2 1.5 0.6 0.9 1.2 1.5
1000/T (1/K) 1000/T (1/K) 1000/T (1/K)

Fig. 1. IDT prediction among three mechanisms. Curves are calculations, symbols are
experimental data [24].
rin

Then, the mechanism of PAI/PAC C3H8/air is developed based on the air plasma

kinetic mechanism [25], propane/air plasma kinetics [16] and oxidation reactions from
ep

Gokulakrishnan [23]. For plasma kinetics, there are 1599 reactions involving various

processes such as excitation, ionization, the relaxation of excited states, and charge
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
exchange reactions. 187 species are included such as electrons, ions, and

vibrationally/electronically excited molecules. The collisions between electrons and

iew
intermediate species/products are ignored due to the low concentrations [13]. Rate

constants of electron impact reactions depending on electron energy distribution

function (EEDF) are calculated by two-term approximation Boltzmann equation using

ev
BOLSIG+ [26] based on the corresponding collision cross sections. The collision

cross sections of O2 and N2 are taken from Phelps [27] and SIGLO [28] databases and

the dissociative excitation collision cross sections of propane are computed according

r
to Ref. [29]. It is noteworthy that the dissociation energies of the C-H bond are larger

er
than that of the C-C bond in propane molecule [30]. Thus, the electron collision

process can not only produce various radicals such as H atoms and H2 but also smaller
pe
hydrocarbon molecules such as C2H4 and CH4. In total, the detailed mechanism

consists of 187 species and 2565 reactions (see Supplementary materials). To the best
ot

of the authors’ knowledge, experimental data related to non-equilibrium PAI/PAC

C3H8/air mechanism validation under high pressures is not available, however, the air
tn

plasma kinetics [25] and propane/air plasma kinetics used in this paper are validated

in [16]. The full mechanism can be found in the supplementary file.


rin

2.2 Numerical method


ep

For plasma kinetics modeling, the following equations (1)-(2) are solved. In energy

conservation Eq. (2), the temperature rising caused by Joule heating is considered by

the penultimate term.


Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
d  Ni  jmax
  Qij (t ) (2)
dt j 1

N gas dTgas jmax

iew
    R j  Pelast   N e   Qsrc (3)
  1 dt j 1

where, [Ni] is the number density for specie i, Qij the rate of production for specie i in

reaction j, Rj the reaction rate for reaction j, Ngas the gas number density, Tgas the gas

ev
temperature, γ the specific heat ratio, Pelast the elastic collision power, [Ne] the electron

number density, Qsrc the source term. δε is the heat release from enthalpy change,

r
which is not considered in plasma kinetics but in combustion kinetics.

er
In combustion reaction kinetics, the evolution of mass fractions for specie k and

temperature yields:
pe
dYk
 Wk k  S  Yk , k  1, 2, , N s (4)
dt

d c pTgas Ns
   hk k  0 (5)
ot

dt k 1

where Wk, Yk, ωk and hk are the molar mass, mass fraction, chemical reaction rate and
tn

specific enthalpy for specie k. ρ is the mixture density, cp the specific heat at constant

pressure. The whole Ns+1 ordinary differential equations are computed by Chemkin-
rin

III [31].

Similar to [13], multiple discharge pulses (100 pulses) are involved, where both
ep

plasma kinetics and combustion kinetics are coupled simultaneously during each pulse

duration. Within one time step, the concentrations of various components and

temperatures are calculated by ZDPlasKin [32] and then solved by the constant
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
pressure chemical solver Chemkin-III for combustion oxidation. After that, all the

species and temperatures are returned to the plasma kinetics solver. The interaction

iew
between both solvers repeats until the end of discharge. After a series of time

dependence tests, the time step size of 10-9 s is employed for temporal evolution. This

model verification can be referred to the previous works [33,34]. Fig. 2 shows the

ev
comparison of OH number density between experiment and simulation. The

experiment was conducted in C3H8/air mixture during single-pulse nanosecond

discharge, where the initial temperature is 300 K, the pressure of 1 atm, equivalence

r
ratios of 0.1 and 1.0, discharge duration of 10 ns and a repetition rate of 10 Hz. During

er
simulation, the voltage of nanosecond pulse discharge is reduced to a square wave.

Fig. 2 indicates a good agreement between simulation and experiment, especially the
pe
delay time of OH after discharge.

1.00
=0.1, EXP
Normalized OH Number density

=0.1, SIM
ot

=1.0, EXP
0.75
=1.0, SIM
tn

0.50

0.25
rin

0.00
10-7 10-6 10-5
Time (s)
ep

Fig. 2. Normalized OH number density during single-pulse nanosecond discharge in C3H8/air


plasma.
Pr

3. Initial condition and computational cases

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
Five cases are carried out to study the effect of spark discharge and NRPD with

various discharge timings on ignition and fuel consumption. The discharge time is

iew
specified at 0 ms (before LTHR), where the LTHR stage is characterized by the first

appearance of the maximum OH [19]. The initial thermo-dynamic conditions for all

cases are a pressure of 30 atm and an equivalence ratio of 0.5. Case 4 and Case 5 are

ev
designed to study the influence of NRPD on NTC. Besides, the repetitive rate of

NRPD is 1 MHz, and the pulse duration is 25 ns [12]. Within one pulse the reduced

electric field (E/N) is fixed at 100 Td [19,20], while in the pulse interval, no energy is

r
inputted (E/N = 0 Td). To keep the same discharge energy at different equivalence

er
ratio cases, the electron density has to be adjusted. In the current cases, the electron

density is specified as 5×1011 cm-3 for φ = 0.5 and kept constant during the simulations.
pe
Nevertheless, the quenching of ions is calculated normally.

Table 1 Initial parameters of simulation cases


Case discharge type discharge time T0 / K φ
ot

1 Auto ignition - 700 0.5


2 NRPD before LTHR 700 0.5
tn

3 Spark before LTHR 700 0.5


4 Auto ignition - 800 0.5
5 NRPD before LTHR 800 0.5
rin

4. Results and discussion


ep

4.1 Electron energy loss fraction

The Electron energy loss fraction represents electron energy transfer direction at a
Pr

certain E/N, which is conducive to understanding the discharge energy deposition

10

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
channels [26]. As shown in Fig. 3, at low E/N (< 30 Td), most of the discharge energy

is used to excite vibrational states of C3H8, O2 and N2. With E/N increasing, electron

iew
energy is spent in the dissociation of C3H8, and excitation of O2 and N2 increases. As

E/N continues to increase, the proportion of electron energy depositing into the

ionization channel boosts and dominates in the discharge. The gray area represents the

ev
current simulation condition where the dissociation of C3H8, vibrationally and

electronically excited N2 and O2 should be included. In addition, dissociation and

ionization of O2 and N2 are added into the kinetic mechanism, while the C3H8

r
vibrational states and C3H8+ are not considered for lack of corresponding reactions.

100
vib
er
pe
N2(exc) O2(dis)
Energy loss fraction

ion
10-1
C3H8(dis)
rot
ot

ela

10-2 O2(exc)
tn

100 101 102 103


E/N (Td)

Fig. 3. The distribution of electron energy loss fractions in E/N space.


rin

4.2 Comparison between NRPD and spark discharge


ep

It is well recognized that spark discharge, characterized by equilibrium plasma,

ignites the charge by importing higher temperature, without chemical selectivity. A

comparison of ignition between equilibrium and non-equilibrium discharge with the


Pr

11

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
same discharge power and duration is carried out, as shown in Fig. 4. The NRPD

power is calculated as follows:

iew
p  j  E  ene vdr E (1)

where, j is the discharge current, E the electric field, e the elementary charge, ne the

electron number density and vdr the electron drift velocity, obtained from BOLSIG+

ev
[26]. According to Eq. (1), the input energy density by NRPD in Case 2 is 0.715 J·cm-
3. The spark discharge is reduced to an external heat source adding to the energy

r
equation.

In Fig. 4a, it can be found that IDT with SI (Case 3) is shortened by 42%, while
er
NRPD (Case 2) is 77% in comparison with autoignition (Case 1). Under the same
pe
input energy, NRPD is more effective than SI in reducing IDT, which is consistent

with the result in Ref. [9]. It is noted that although using the same input energy, the

temperature of Case 3 is higher than that of Case 2 at 0.1 ms (the end of discharge),
ot

indicating that partial energy during NRPD is transferred to the internal energy of

molecules. The temperature in Case 3 keeps constant immediately as soon as the


tn

discharge is off. However, after discharge, the temperature in Case 2 continues to

increase and exceeds that of Case 3 at about 0.2 ms. From Fig. 4b, it can be seen that
rin

HRR oscillates with pulsed discharge and considerable heat has been released during

NRPD. While the heat release during spark discharge can be ignorable. The results
ep

also show that for both discharges, the LTC stage is shortened and then, the ignition

event occurs in advance compared with Case 1 without discharge.


Pr

12

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
2000 10-3 1.5
1.5
a b

HRR 1011 (Jm-3s-1)


1800 T, Case 1
1.0
T, Case 2

HRR 1011 (Jm-3s-1)


1600 T, Case 3
1.0
OH, Case 1 10-6 0.5

Mole fraction

iew
1400 OH, Case 2
T (K)

OH, Case 3 0.0

760
discharge
1200

3
4
0
1
4

6
02
02
06
06
10

10
0.
0.
0.
0.
0.

0.
0.5

720 740
Time (s)
10-9

T (K)
1000
total HRR, Case 2
total HRR, Case 3

700
800 total HRR, Case 1
0.0 0.1 0.2 0.3 0.0
Time (ms)
600 10-12

ev
0.00 0.05 0.10 0.15 0.20 10-6 10-5 10-4 10-3 10-2 10-1
Time (s) Time (s)

Fig. 4. Temporal evolution of a) temperature, OH mole fraction, and b) HRR under auto-
ignition, SI and NRPD conditions for cases 1-3, respectively.

r
The reason for the continuous temperature rising in Case 2 can be explained in Fig.
er
5 and Fig. 6. In Fig. 5, the initial repetitive spikes denote discharge pulses. It is found
pe
that there are more excited states formed during NRPD such as O2(a1Δg), vibrationally

excited O2 and N2. For NRPD in Case 2, alkyls such as IC3H7 and NC3H7 and smaller

hydrocarbon molecules such as C2H4 are produced in advance compared with SI in


ot

Case 3 due to electron and electronically excited N2 collisions with propane [16],

which will be explained in Fig. 8. Also, an amount of O, O(1D) and OH are formed
tn

during discharge which accelerate propane oxidation [16,35]. On the contrary, radicals

are rarely generated during spark discharge (before 1 ms).


rin
ep
Pr

13

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
10-2
O103 O2(v1)

10-4 C 2 H4
OH10 N2(v1)

iew
Mole fraction
C 2 H4
10-6 OH10
NC3H7
-8
10
IC3H7 NC3H7
O2(a1g)
10-10

ev
IC3H7
1
O( D)
10-12
10-6 10-5 10-4 10-3 10-2
Time (s)

r
Fig. 5. The temporal evolution of mole fractions in terms of key species in Case 2 (solid line)

er
and Case 3 (dotted line).

It is suggested that for propane combustion there are three different heat release
pe
stages, namely, a high-temperature (T > 1200 K), a low-temperature (T < 900 K), and

an intermediate-temperature (T = 1000 – 1200 K) heat release stage [2,35]. We focus

on LTHR stage in cases 2 and 3 where discharge is initiated before LTC stage.
ot

Generally, if T is below LTC region, the ignition of hydrocarbon fuel is initiated by


tn

the abstraction of hydrogen through (I) and (II). Then the formation of peroxide via

(III) and hydroperoxide is followed by hybridization (IV). Subsequently,

hydroperoxide combines with O2 to produce O2QOOH through (V). Finally, hydroxyl


rin

radicals and oxygen-containing hydrocarbons are produced via (VI) and (VII). During

NTC region, peroxide RO2 and hydroperoxide QOOH usually decompose into R, HO2
ep

and alkene via RO2 => R + O2 and QOOH => HO2 + alkene. Thus, the chain branching

reactions (VI) and (VII) are replaced and the fuel consumption rate is slowed down,
Pr

14

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
leading to the NTC effect. The detailed NTC analysis for C3H8/air can be referred to

Appendix.

iew
RH+O2=R+HO2 (I)

RH+X=R+XH (X=O, H, OH) (II)

R+O2=RO2 (III)

ev
RO2=QOOH (IV)

QOOH+O2=O2QOOH (V)

O2QOOH=OQ’OOH+OH (VI)

r
OQ’OOH=Q’’O+Q’’’O+OH (VII)

er
Heat release rates (HRRs) of main exothermic reactions after NRPD are plotted in

Fig. 6. It is found that the accumulated alkyls IC3H7 and NC3H7 during discharge are
pe
oxidized to form propyl peroxy radicals IC3H7O2 and NC3H7O2 through R478, R476

and dominate heat release during LTC, which is illustrated in reactions (I)-(VII). In
ot

addition, reactions R2078 and R1966 also play a significant role, indicating the

importance of electron attachment reaction and O2(v1) vibrational-relaxation (VT)


tn

process in LTC stage. Following that, oxidation pathways involving N and H atoms

through reactions R616 and R11 also release a majority of the heat meanwhile.
rin
ep
Pr

15

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
109

iew
HRR (Jm-3s-1) 108 R478: O2+IC3H7=>IC3H7O2
R2078: O2+E+O2=>O2-+O2
R476: O2+NC3H7=>NC3H7O2
R616: N+O2=NO+O

ev
R1966: O2(v1)+N2=>O2+N2
R11: H+O2(+M)=HO2(+M)
107
0.10 0.15 0.20
Time (ms)

r
Fig. 6. The temporal evolution of HRRs of main reactions in Case 2 after discharge.

er
Fig. 7 shows the key species in terms of OH production at LTC stage during NRPD

(Case 2) and auto-ignition (Case 1) processes. Compared to Case 1, there are much
pe
more low-temperature products related to the OH low-temperature generation

pathway under NRPD, such as NC3H7O2 (RO2), C3H6OOH (QOOH), O2C3H6OOH

(O2QOOH) and OC3H5OOH (OQ’OOH) which is known as KETO. Especially during


ot

the isomerization of O2QOOH and KETO, OH-carriers break down to form OH


tn

radicals and release heat. The more low-temperature products produced during NRPD

promote OH production and enhance low-temperature reaction pathways.


rin
ep
Pr

16

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
Case 2: OH NC3H7O2 C3H6OOH O2C3H6OOH OC3H5OOH
Case 1: OH NC3H7O2 C3H6OOH O2C3H6OOH OC3H5OOH
10-3 2000
T, Case 2
T, Case 1
10-4

iew
10-5
Mole fraction 1500
10-6

T (K)
10-7
1000
10-8

ev
10-9

10-10 500
10-6 10-5 10-4 10-3 10-2 10-1
Time (s)

r
Fig. 7. The temporal evolution of main radical mole fractions at LTC stage and temperature in
Case 2 and Case 1.

er
Fig. 8 shows the Path Flux Analysis (PFA) of propane consumption in Case 2. Here,
pe
PFA is defined as the proportion of the contribution from a certain reaction to the total

production/consumption of one species integrating over the entire discharge process.

Propane dissociation directly from electron collision via R2475, R2476 occupies
ot

38.5% of propane destruction. Also, plasma directly generated high energy excited

atoms O(1D) reacting with fuel (R1000 and R1001) accompanied with the rapid
tn

generation of IC3H7 and NC3H7 is the second important propane consumption

pathway, which can explain the early generation of propyl radicals in Fig. 5.
rin

Meanwhile, neutral specie OH is still the primary factor of propane destruction,

accounting for 43.7%. Conversely, H atoms play a minor role in propane


ep

consumption [35], as well as N2(A3Σu+) which is not shown for brevity.


Pr

17

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
R317: C3H8+OH=H2O+NC3H7

R2475: E+C3H8=>E+C2H4+CH4

R326: C3H8+OH=H2O+IC3H7

iew
R2476: E+C3H8=>E+C3H6+H2

R327: C3H8+O=OH+IC3H7

R1001: O(1D)+C3H8=>NC3H7+OH

R1000: O(1D)+C3H8=>IC3H7+OH

R318: C3H8+O=OH+NC3H7

ev
R2477: E+C3H8=>E+C3H4+2H2

R325: C3H8+H=H2+IC3H7

R2483: E+C3H8=>E+NC3H7+H

R2482: E+C3H8=>E+IC3H7+H

r
R996: N2(A3u+)+C3H8=>N2+C3H6+H2

-0.3 -0.2 -0.1 0.0 0.1 0.2 0.3

er
Normalized ROP of C3H8

Fig. 8. PFA of propane integrated into the discharge duration for Case 2. Negative values denote
specie destruction.
pe
According to Fig. 8, about 51.4% of propane is consumed by reacting with OH and

O in total during discharge. Subsequently, Fig. 9 shows the PFA of (a) O and (b) OH
ot

in the whole discharge duration (0.1 ms), respectively. Normally, it is difficult to

produce an O atom at low temperatures since oxygen-oxygen bond energy is relatively


tn

high. However, Fig. 9a shows that almost all O atoms generation is related to plasma

reactions which is the kinetic effect of plasma modifying the combustion process. N2
rin

electronically excited states and E contribute 23.8% and 8.3% of O via dissociating

O2, respectively. About 37.6% of O is produced due to N oxidation, primarily because

of the reaction of N with O2. In addition, N2 vibrationally excited states via reaction
ep

N2(v) + O => NO + N [11] dominates N production, which is the source of N radicals

in R616 (c.f. Fig. 6). Here, N2(v) is the group of N2(v1)-N2(v8). It should be noted that
Pr

18

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
the rate constant for N2(v) relaxation reactions, i.e., N2(v1) + O => NO + N is

estimated as k = 3.01×1013 cm3/(mol·s), much larger than that of N2(v) VT relaxation

iew
with hydrocarbons of 6.02×109 cm3/(mol·s) in [16]. Besides, electronically excited

molecules of N2 and O2 and O(1D) provide 49.7% of O in total. Electron collision with

O2 and N2 can also promote concentrations of O and N.

ev
In Fig. 9b, the top three pathways of OH at LTC stage are propane oxidation,

O2C3H6OOH isomerization decomposition and HO2 decomposition. In detail, 33.4%

of OH is directly formed by the impact of O(1D) and O with propane and 28.2% by

r
the isomerization of O2C3H6OOH. Besides, at low temperature and high pressure, H

er
is usually consumed to produce HO2 by reaction R11: H + O2(+M) = HO2(+M), which

is a chain-termination process [7]. While with the existence of NO, a chain branching
pe
reaction NO + HO2 = NO2 + OH takes place resulting in the HO2 decomposing into

OH and benefits the subsequent combustion [11]. Furthermore, NO2 promotes H atom
ot

oxidation to produce OH and N2(v) can also promote H by decomposing HO2 as

explained in [14].
tn
rin
ep

Fig. 9. PFA of a) O and b) OH before the end of discharge in Case 2.


Pr

19

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
4.3 NTC under NRPD and SI

At a certain temperature after LTC stage of C3H8/air auto-ignition process, RO2,

iew
QOOH and O2QOOH tend to decompose to produce R, alkene and HO2, causing chain

termination and lowering the radical production rate. As a result, the fuel oxidation

rate and heat release rate slow down, leading to IDT increasing with a temperature

ev
rising, which is called an NTC effect. The knocking and oscillatory combustion are

closely related to NTC [2]. It has been proven that the non-equilibrium discharge can

r
weaken the NTC effect [7], while detailed physical understanding is still rare. This

section will explore how the NTC during the combustion process of C3H8/air is

affected by NRPD. er
pe
From Fig. 10, it can be seen that the NTC region is in the temperature range of 765-

865 K for φ = 1.0 and 772-833 K for φ = 0.5 at auto-ignition condition. With the

energy of 0.715 J·cm-3 inputted by spark, the NTC region shifts towards the lower
ot

temperature range of 733-790 K for φ = 1.0 and 725-769 K for φ = 0.5, respectively.

The reason can be ascribed to the thermal effect brought by spark. With the enhanced
tn

initial thermal condition, the NTC phenomenon can be found earlier as the time

involves such that the NTC region is “pushed” into a lower temperature range in Fig.
rin

10a. However, the NTC phenomenon disappears in both cases if NRPD is employed

in Fig. 10b, where the induction time decreases monotonically with the growth of the
ep

initial temperature which is consistent with the results in Ref. [19]. Furthermore, the

ignition delay time of NRPD mode among the three ignition modes is the shortest.
Pr

20

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
0.10 0.10
NTC  = 0.5, auto ignition
765 K 815 K  = 0.5, auto ignition
 = 1.0, auto ignition
0.08  = 1.0, auto ignition
 = 0.5, SI
 = 0.5, NRPD
 = 1.0, SI
 = 1.0, NRPD

IDT (s)
0.06
a

iew
IDT (s)

NTC 0.05 b
725 K 769 K
0.04
NTC
772 K 833 K

0.02

733 K NTC 790 K


0.00 0.00
700 750 800 850 900 950 1000

ev
700 750 800 850 900 950 1000
T (K) T (K)

Fig. 10. The evolution of IDT with different initial temperatures for φ = 0.5 and φ = 1.0 under
conditions of a) auto ignition and SI, b) auto ignition and NRPD.

r
For understanding the NTC disappearance with NRPD, concentrations of key

er
species of low-temperature pathways are compared between NRPD (Case 5) and auto-

ignition (Case 4) for T0 = 800 K in Fig. 11, where Case 4 is inside of NTC region as
pe
shown in Fig. 10. It can be found that with NRPD, an amount of NC3H7 is formed

from the beginning due to abstraction of H from the fuel oxidation by O, OH and
ot

O(1D) as shown in Fig. 9, and grows dramatically to its peak. Subsequently, NC3H7O2

is generated from NC3H7 oxygenation via reaction R476 (O2 + NC3H7 => NC3H7O2)
tn

with a high reaction rate under NRPD in Fig. 11a. While for the auto-ignition Case 4,

due to the relatively low initial temperature, the reaction rate of R476 is rather low
rin

and NC3H7 is hardly to generate until 0.04 s.

Comparing the major species during the ignition process in Fig. 11b, a great number

of intermediate species C3H6OOH, O2C3H6OOH and OC3H5OOH have generated


ep

during plasma discharge period and even two orders of magnitude larger than those of

the corresponding auto-ignition case. It is well recognized that O2C3H6OOH and


Pr

21

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
OC3H5OOH show a notable impact on LTC [2,35] which results in a chain branching

reaction for OH production. As such, the low-temperature pathway of OH production

iew
is enhanced. An apparent two-stage ignition and a shorter ignition delay time under

NRPD can be seen even though the initial temperature is inside the NTC region.
Case 5: NC3H7O2 C3H6OOH O2C3H6OOH OC3H5OOH OH
Case 4: NC3H7O2 C3H6OOH O2C3H6OOH OC3H5OOH OH
0.025
Case 5: R476 NC3H7 10-4 2000
Case 4: R476 NC3H7 10-6

ev
Reaciton rate (molcm-3s-1)

0.020

10-7

Mole fraction
10-6 1500

Mole fraction
0.015
Reaciton rate (molcm-s-1)
0.002

T (K)
0.010 10-8
0.001

r
10-8 1000
0.005 10-9
0.000

0.04 0.06 T, NRPD


Time (s) T, auto ignition
0.000
10-6 10-5 10-4
Time (s)
10-3

Fig. 11. The temporal evolution of a) NC3H7 and R476 reaction rate and b) temperature and key
10-2
er
10-10 10-10
10-6 10-5 10-4
Time (s)
10-3 10-2
500
pe
low-temperature species for Case 4 and Case 5.

The reaction rates of the main reactions for NC3H7 and oxidation rates of
ot

NC3H7/IC3H7 during NRPD in Case 5 are shown in Fig. 12a, b respectively. New

pathways like R1001 and R2483 for NC3H7 production are induced by non-
tn

equilibrium plasma, besides R317, R318 and R477. According to Figs. 7 and 8, O

produced by excited states and electrons accelerates the reactions R318 and R327
rin

(C3H8 + O = OH + NC3H7/IC3H7). Therefore, NC3H7 and IC3H7 concentrations are

promoted even in NTC region, which further increases the oxidation reaction rates of
ep

NC3H7/IC3H7 in terms of R476 and R478 (O2 + IC3H7 => IC3H7O2). However, this

cannot be achieved in auto-ignition Case 4 since both the quantities and the reaction

rates of R476 and R478 are low within the NTC region. Finally, the promoted
Pr

22

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
NC3H7O2 and IC3H7O2 concentrations accelerate the production of hydroxy

compounds as well as OH through low-temperature chain branching reactions, leading

iew
to a shorter IDT.
R317: C3H8+OH=H2O+NC3H7 R1001: O(1D)+C3H8=>NC3H7+OH
R318: C3H8+O=OH+NC3H7 R2483: E+C3H8=>E+H+NC3H7
R477: NC3H7O2=>O2+NC3H7 R343: O2+NC3H7=C3H6+HO2 R351: O2+IC3H7=C3H6+HO2
0.025 0.006
a 10-4
b
Reaciton rate (molcm-3s-1)

Reaciton rate (molcm-3s-1)


0.020

ev
HO2 Mole fraction
0.004
0.015 10-6

0.010
0.002 R343, Case 5
R343, Case 4 10-8

r
0.005 R351, Case 5
R351, Case 4
HO2, Case 5
HO2, Case 4
0.000 0.000 10-10
10-6 10-5
Time (s)
10-4 10-3

er
Fig. 12 a) The primary ROP in terms of NC3H7 and b) reaction rate of R343, R351 and mole
10-6 10-5 10-4
Time (s)
10-3 10-2 10-1
pe
fraction of HO2 for Case 5.

Besides the enhanced low-temperature pathways in Case 5, it is found another way

affecting high-temperature ignition. As expected, with temperature increasing, the


ot

consumption pathway of NC3H7/IC3H7 transfers from oxygen addition to H-


tn

abstraction i.e., from R476 and R478 to R343 and R351 in Fig. 12b. Such that, the

low-temperature pathway weakens while intermediate/high temperature reaction

processes begin to dominate. Fig. 12b shows an advanced and enhanced H-abstraction
rin

reaction rate from NC3H7/IC3H7. Accordingly, the increased HO2 radicals contribute

to H2O2 and OH formation [2,35]. It is well recognized that this process is the major
ep

chain-branching reaction pathway at intermediate/high temperature. The detailed

pathways are not shown for brevity. Therefore, the promoted NC3H7/IC3H7 by NRPD
Pr

23

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
not only increases the production of NC3H7O2/IC3H7O2, but accelerates the production

of HO2 and further the intermediate/high temperature reaction processes.

iew
5. Conclusions

Zero-dimensional plasma kinetics and combustion kinetics are coupled to investigate

the effect of non-equilibrium plasma and spark discharge on LTC stage and NTC

ev
performance of lean propane/air mixture. The detailed PAI/PAC propane/air

mechanism composed of electron collision reactions is established. Results show that

r
NRPD can modify propane consumption pathways and shorten IDT.

For NRPD, electronically excited species such as O(1D), O2(v) and N2(v) as well as
er
propyl are produced in advance compared to SI. In addition, it is found that electrons
pe
and O(1D) play a significant role in propane destruction, and abundant alkyls such as

IC3H7 and NC3H7 are formed during the discharge period when the discharge timing

is before LTC. Meanwhile N, O(1D) and electronically excited states of N2 and O2


ot

produce O atoms, accelerating propane oxidation. The decomposition of O2C3H6OOH

from previously generated alkyls through LTC and propane oxidation by O(1D) and
tn

O contribute to a majority of OH which releases heat and shortens LTC stage.

Besides, NRPD can eliminate the NTC effect at fuel lean conditions (φ = 0.5), while
rin

spark discharge has a minor effect on the NTC region which shifts the NTC region

into the lower temperature zone. The negative coefficient relation between the reaction
ep

rate of NC3H7/IC3H7 + O2 => NC3H7O2/IC3H7O2 and temperature leads to the NTC

phenomenon. With T0 rising, the reaction rate of NC3H7 + O2 => C3H6 + HO2

dominates among NC3H7 consumption pathways and the low-temperature pathways


Pr

24

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
of OH are replaced. When the T0 is inside the NTC region, plasma-induced radicals O

and O(1D) can promote C3H8 oxidation to produce higher concentrations of OH and

iew
NC3H7 radicals. The higher NC3H7 concentration is not only beneficial for NC3H7

oxidation to NC3H7O2 offsetting the low production rate of NC3H7O2 brought by the

negative coefficient relation between NC3H7 oxygenation reaction and temperature,

ev
but also accelerates the intermediate/high temperature reaction rates to produce C3H6

and HO2. As a result, with NRPD the LTC stage in NTC region and intermediate/high

temperature reaction processes are advanced, leading to the disappearance of NTC

r
and a shorter ignition induction time.

Acknowledgements
er
pe
We thank Dr. Xingqian Mao for his valuable comments and proofreading.
ot

Funding

This work was supported by the National Natural Science Foundation of China (Grant
tn

No. 51876139) and special funds for the construction of innovative provinces in

Hunan (Grant No. 2019RS2028).


rin

Supplementary materials
ep

Supplementary material is submitted along with the manuscript.


Pr

Appendix

25

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
According to Ref. [7,35], with temperature increasing, the pathways of OH production

and fuel consumption are changed. The appendix explains the reason for the NTC

iew
phenomenon and the transition process between LTC and intermediate/high

temperature stage for C3H8/air in detail.

As shown in Fig. A1, the reduction of R476 and R478 reaction rate constants with

ev
temperature rising will restrain the production of NC3H7O2/IC3H7O2. Within a wide

temperature range, the reaction rates of R343 and R351 are lower than that of R476

and R478, leading to a restricted H-abstraction rate at low temperatures. It can be

r
concluded that the negative coefficient relation between reaction rate constants of

R476 and R478 and temperature leads to lower RO2 and OH production rates and

finally the NTC phenomenon.


er
pe
Reaction rate constant 1013 (cm3mol-1s-1)

1.6
R343: O2+NC3H7=C3H6+HO2
R351: O2+IC3H7=C3H6+HO2
1.2 R476: O2+NC3H7=>NC3H7O2
R478: O2+IC3H7=>IC3H7O2
ot

0.8
tn

0.4

0.0
500 1000 1500 2000
rin

T (K)

Fig. A1. Reaction rate constants of R343, R351, R476 and R478.

The ROP of NC3H7 for the initial temperature of 905 K is shown in Fig. A2 to
ep

explain the disappearing NTC phenomenon under the auto-ignition condition. In Fig.

A2, the reaction rate of R343 begins to be larger than that of R476 at about 0.023 s
Pr

26

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
marked by the dotted line, and the consumption path of NC3H7 is changed. Reaction

R343 promotes HO2 production, further accelerating the production of H2O2

iew
respectively by C3H8 + HO2 = H2O2 + C3H7 and 2HO2 = H2O2 + O2. Finally, OH is

promoted by H2O2 decomposition via H2O2(+M) = 2OH(+M), the low-temperature

generation pathway of which is replaced, and the propane consumption is accelerated

ev
[7,35].

0.000
ROP of NC3H7 (molcm-3s-1)

r
-0.002
er
R343: O2+NC3H7=C3H6+HO2
R476: O2+NC3H7=>NC3H7O2
pe
0.000 0.005 0.010 0.015 0.020 0.025 0.030
Time (s)

Fig. A2. ROP of NC3H7 under the initial temperature of 905 K for φ = 0.5.
ot

References
tn

[1] Krishnamoorthi M, Malayalamurthi R, He Z, Kandasamya S. A review on low

temperature combustion engines: Performance, combustion and emission


rin

characteristics. Renew Sust Energ Rev 2019; 116: 109404.

https://doi.org/10.1016/j.rser.2019.109404.
ep

[2] Ju Y, Reuter CB, Yehia OR, Farouk TI, Won SH. Dynamics of cool flames. Prog

Energ Combust 2019; 75: 100787. https://doi.org/10.1016/j.pecs.2019.100787.


Pr

[3] Saxena S, Bedoya ID. Fundamental phenomena affecting low temperature

27

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
combustion and HCCI engines, high load limits and strategies for extending these

limits. Prog Energ Combust 2013; 39: 457-488.

iew
https://doi.org/10.1016/j.pecs.2013.05.002.

[4] Nguyen MT, Shy SS, Chen YR, Lin BL, Huang SY, Liu CC. Conventional spark

versus nanosecond repetitively pulsed discharge for a turbulence facilitated ignition

ev
phenomenon. Proc Combust Inst 2021; 38: 2801-2808.

https://doi.org/10.1016/j.proci.2020.06.020.

[5] Shiraishi T, Kakuho A, Urushihara T, Cathey C, Tang T, Gundersen M. A Study

r
of Volumetric Ignition Using High-Speed Plasma for Improving Lean Combustion

er
Performance in Internal Combustion Engines. SAE Int J Engines 2008; 1: 399-408.

https://doi.org/10.4271/2008-01-0466.
pe
[6] Lovascio S, Hayashi J, Stepanyan S, Stancu GD, Laux CO. Cumulative effect of

successive nanosecond repetitively pulsed discharges on the ignition of lean


ot

mixtures. Proc Combust Inst 2019; 37: 5553-5560.

https://doi.org/10.1016/j.proci.2018.06.029.
tn

[7] Ju Y, Sun W. Plasma assisted combustion: Dynamics and chemistry. Prog Energy

Combust Sci 2015; 48: 21-83. https://doi.org/10.1016/j.pecs.2014.12.002.


rin

[8] Ju Y, Lefkowitz JK, Reuter CB, Won SH, Yang X, Yang S, et al. Plasma assisted

low temperature combustion. Plasma Chem Plasma P 2016; 36: 85-105.


ep

https://doi.org/10.1007/s11090-015-9657-2.

[9] Shiraishi T, Urushihara T, Gundersen M. A trial of ignition innovation of gasoline

engine by nanosecond pulsed low temperature plasma ignition. J Phys D Appl Phys
Pr

28

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
2009; 42: 135208. http://dx.doi.org/10.1088/0022-3727/42/13/135208.

[10] Cathey C, Cain J, Hai W, Gundersen MA, Carter C, Ryan M. OH production by

iew
transient plasma and mechanism of flame ignition and propagation in quiescent

methane–air mixtures. Combust Flame 2008; 154: 715-727.

https://doi.org/10.1016/j.combustflame.2008.03.025.

ev
[11] Togai K. Kinetic modeling and sensitivity analysis of plasma-assisted

combustion, PhD thesis, The Pennsylvania State University, Commonwealth of

Pennsylvania, U.S., 2015.

r
[12] Uddi M. Non-Equilibrium Kinetic studies of repetitively pulsed nanosecond

er
discharge plasma assisted combustion, PhD thesis, The Ohio State University, Ohio

State, U.S., 2008.


pe
[13] Lefkowitz JK, Guo P, Rousso A, Ju Y. Species and temperature measurements

of methane oxidation in a nanosecond repetitively pulsed discharge. Phil Trans R


ot

Soc A 2015; 373: 20140333. https://doi.org/10.1098/rsta.2014.0333.

[14] Starikovskiy A, Aleksandrov N. Plasma-assisted ignition and combustion. Prog


tn

Energy Combust 2013; 39: 61-110. https://doi.org/10.1016/j.pecs.2012.05.003.

[15] Mao X, Rousso AC, Chen Q, Ju Y. Numerical modeling of ignition enhancement


rin

of CH4/O2/He mixtures using a hybrid repetitive nanosecond and DC discharge.

Proc Combust Inst 2019; 37: 5545-5552.


ep

https://doi.org/10.1016/j.proci.2018.05.106.

[16] Adamovich IV, Li T, Lemper WR. Kinetic mechanism of molecular energy

transfer and chemical reactions in low-temperature air-fuel plasmas. Phil Trans R


Pr

29

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
Soc A 2015; 373: 20140336. https://doi.org/10.1098/rsta.2014.0336.

[17] Rousso AC, Mao X, Chen Q, Ju Y. Kinetic studies and mechanism development

iew
of plasma assisted pentane combustion. Proc Combust Inst 2019; 37: 5595-5603.

https://doi.org/10.1016/j.proci.2018.05.100.

[18] Zhong H, Mao X, Rousso AC, Patrick CL, Yan C, Xu W, et al. Kinetic study of

ev
plasma-assisted n-dodecane/O2/N2 pyrolysis and oxidation in a nanosecond-pulsed

discharge. Proc Combust Inst 2021; 38: 6521-6531.

https://doi.org/10.1016/j.proci.2020.06.016.

r
[19] Filimonova EA. Discharge effect on the negative temperature coefficient

er
behaviour and multistage ignition in C3H8-air mixture. J Phys D Appl Phys 2015;

48: 015201. http://dx.doi.org/10.1088/0022-3727/48/1/015201.


pe
[20] Filimonova E, Bocharov A, Bityurin V. Influence of a non-equilibrium discharge

impact on the low temperature combustion stage in the HCCI engine. Fuel 2018;
ot

228: 309-322. https://doi.org/10.1016/j.fuel.2018.04.124.

[21] Bramlette RB, Depcik CD. Review of Propane-Air Chemical Kinetic


tn

Mechanisms for a Unique Jet Propulsion Application. J Energy Inst 2020; 93: 857-

877. https://doi.org/10.1016/j.joei.2019.07.010.
rin

[22] Petersen EL, Kalitan DM, Simmons S, Bourqu G, Curran HJ, Simmie JM.

Methane/propane oxidation at high pressures: Experimental and detailed chemical


ep

kinetic modeling. Proc Combust Inst 2007; 31: 447-454.

https://doi.org/10.1016/j.proci.2006.08.034.

[23] Gokulakrishnan P, Fuller CC, Klassen MS, Joklik RG, Kochar YN, Vaden SN,
Pr

30

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
et al. Experiments and modeling of propane combustion with vitiation. Combust

Flame 2014; 161: 2038-2053. https://doi.org/10.1016/j.combustflame.2014.01.024.

iew
[24] Gallagher SM, Curran HJ, Metcalfe WK, Healy D, Simmie JM, Bourque G. A

rapid compression machine study of the oxidation of propane in the negative

temperature coefficient regime. Combust Flame 2008; 153: 316-333.

ev
https://doi.org/10.1016/j.combustflame.2007.09.004.

[25] Defilippo AC, Chen JY. Modeling plasma-assisted methane-air ignition using

pre-calculated electron impact reaction rates. Combust Flame 2016; 172: 38-48.

r
https://doi.org/10.1016/j.combustflame.2016.07.005.

er
[26] Hagelaar GJM, Pitchford LC. Solving the Boltzmann equation to obtain electron

transport coefficients and rate coefficients for fluid models. Plasma Sources Sci T
pe
2005; 14: 722-733. http://dx.doi.org/10.1088/0963-0252/14/4/011.

[27] Phelps AV. Phelps database, available at <www.lxcat.net>.


ot

[28] Phelps AV, Pitchford LC. SIGLO database, available at <www.lxcat.net>.

[29] Janev RK, Reiter D. Collision processes of C2,3Hy and C2,3Hy+ hydrocarbons with
tn

electrons and protons. Phys Plasmas 2004; 11: 780-829.

https://doi.org/10.1063/1.1630794.
rin

[30] Moreau N, Pasquiers S, Blin-Simiand N, Magne L, Jorand F, Postel C, et al.

Propane dissociation in a non-thermal high-pressure nitrogen plasma. J Phys D Appl


ep

Phys 2010; 43: 619-625. http://dx.doi.org/10.1088/0022-3727/43/28/285201.

[31] Kee RJ, Rupley FM, Meeks E, Miller JA. Chemkin-III: a Fortran chemical

kinetics package for the analysis of gas-phase chemical and plasma kinetics, Report
Pr

31

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040
ed
No. SAND96-8216, Sandia National Laboratories, 1996.

[32] Pancheshnyi S, Eismann B, Hagelaar GJM, Pitchford LC. Computer code

iew
ZDPlasKin, available at <http://www.zdplaskin.laplace.univ-tlse.fr>.

[33] Feng R, Zhu J, Wang Z, Zhang F, Ban Y, Zhao G, et al. Suppression of

combustion mode transitions in a hydrogen-fueled scramjet combustor by a multi-

ev
channel gliding arc plasma. Combust Flame 2022; 237: 111843.

https://doi.org/10.1016/j.combustflame.2021.111843.

[34] Ban Y, Zhang F, Zhong S, Zhu J. The numerical simulation of nanosecond-pulsed

r
discharge-assisted ignition in lean-burn natural gas HCCI engines. Frontiers in

Mechanical Engineering

https://doi.org/10.3389/fmech.2022.930109.
er 2022; 8: 930109.
pe
[35] Zhukov VP, Sechenov VA, Starikovskii A. Autoignition of a Lean Propane-Air

Mixture at High Pressures. Kinet Catal 2005; 46: 319-327.


ot

https://doi.org/10.1007/s10975-005-0079-7.
tn
rin
ep
Pr

32

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4270040

You might also like