ArticlePDF Available

An experimental and kinetic modeling study of ammonia/n-heptane blends

Authors:

Abstract and Figures

Ammonia is carbon free and hence is a promising renewable fuel to achieve a reduction in CO2 emissions. However, due to its relatively low reactivity, ammonia is often blended with other high reactivity fuels in practical combustors. This study aims to understand the chemical kinetics of ammonia blended with n-heptane, which is a primary reference fuel and an important component in diesel and gasoline surrogate models. A high-pressure shock tube is used to measure the ignition delay times of ammonia/n-heptane blends with different blending ratios, for stoichiometric mixtures at 10 atm pressure in the temperature range 1000–1400 K. The experimental results show that fuel reactivity decreases with increasing ammonia concentration. The oxygen concentration also shows a large effect on the reactivity of ammonia/n-heptane blends. A new detailed kinetic model is developed to simulate these new ignition delay times in addition to experimental data available in the literature. Overall, the current kinetic model can predict well the auto-ignition behavior and laminar burning velocities of ammonia/n-heptane blends over a wide range of experimental conditions. Flux and sensitivity analyzes show that the interaction reaction pathways between ammonia and n-heptane via H-atom abstraction from n-heptane by ṄH2 radicals are important in predicting the fuel reactivity of ammonia/n-heptane blends.
Content may be subject to copyright.
1
An experimental and kinetic modeling study of ammonia/n-heptane
1
blends
2
3
Shijun Dong1, Bowen Wang1, Zuozhou Jiang1, Yuhang Li1, Wenxue Gao1, Zhaowen Wang1, 4
Xiaobei Cheng1*, Henry J. Curran2
5
6
1School of Energy and Power Engineering, Huazhong University of Science and Technology, Wuhan, China
7
2Combustion Chemistry Centre, School of Chemistry, Ryan Institute, MaREI, NUI Galway, Ireland. 8
Abstract
9
Ammonia is carbon free and hence is a promising renewable fuel to achieve a reduction in CO2
10
emissions. However, due to its low fuel reactivity, ammonia is often blended with other high
11
reactivity fuels in practical combustors. This study aims to understand the chemical kinetics of
12
ammonia blended with n-heptane, which is a primary reference fuel and an important component
13
in diesel and gasoline surrogate models. A high-pressure shock tube is used to measure the ignition
14
delay times of ammonia/n-heptane blends with different blending ratios, for stoichiometric
15
mixtures and 10 atm pressure in the temperature range of 1000 1400 K. The experimental results
16
show that fuel reactivity decreases with increasing concentration of ammonia. The oxygen
17
concentration also shows a large effect on the fuel reactivity of ammonia/n-heptane blends. A new
18
detailed kinetic model is developed to simulate these new ignition delay times in addition to
19
experimental data available in the literature. Overall, the current kinetic model can predict well the
20
auto-ignition behavior and laminar burning velocities of ammonia/n-heptane blends over a wide
21
range of experimental conditions. Flux and sensitivity analyses show that the interaction reaction
22
pathways between ammonia and n-heptane via H-atom abstraction from n-heptane by ṄH2 radicals
23
are important in predicting the fuel reactivity of ammonia/n-heptane blends.
24
Keywords: Ammonia, n-heptane, kinetics modeling, shock tube, ignition delay time
25
26
2
1. Introduction
1
Ammonia is carbon free and hence does not produce carbon dioxide (CO2) during combustion.
2
Meanwhile, ammonia can be produced via renewable energy such as from wind and solar energy
3
[1]. As a hydrogen carrier, the physical and chemical properties of ammonia facilitate easier
4
storage and transportation compared to hydrogen. Therefore, ammonia is a promising renewable
5
fuel used to achieve CO2 emissions reduction and carbon neutral [2]. However, ammonia shows
6
much lower fuel reactivity compared to hydrocarbon fuels, i.e. it has a much higher octane number
7
and much lower flame speeds compared to conventional fuels. Comparisons of fuel properties for
8
ammonia and hydrogen carbon fuels are provided in Table 1. Therefore, to improve its combustion
9
performance, ammonia is often blended with high reactivity fuels such as hydrogen, natural gas
10
and diesel when burned in practical facilities. This study aims to understand the chemical kinetics
11
of ammonia blended with n-heptane, which is a primary reference fuel and an important
12
component in diesel and gasoline surrogate models.
13
Table 1: Comparisons of fuel properties for ammonia and hydrogen carbon fuels.
14
ammonia
methane
hydrogen
n
-
heptane
iso
-
Lower heating value
(
MJ/kg
)
18.8[2] 50.05[2] 120.0[2] 44.6[4] 44.654[4]
Laminar burning velocity
(m/s)
-
close to stoich.
0.07[2] 0.38[2] 3.51[2] 0.37[3] 0.33[3]
Combustion limit
(volume fraction %)
15.0 28.0[2]
5.0 15.0[2]
4.7 75.0[2]
1.05 6.7[6]
0.95 6.5[6]
Octane number
130
[2]
120
[1]
-
0
100
Adiabatic flame
temperature
(with air)
(K)
1850[1] 2223[1] 2483[1] 2294[5] 2300[4]
*LBVs measured with fuel in ‘air’ mixture, at 1 atm, initial temperature of 298 K close to stoichiometric
15
condition.
16
Over the past decade, a number of experimental and modeling studies have focused on the
17
combustion characteristics of ammonia and its blends with different fuels, and most of these are
18
carbon free or low-carbon fuels, including hydrogen, methane, methanol and dimethyl ether
19
(DME). Tian et al. [7] measured the intermediates and products formed in premixed
20
NH3/CH4/O2/Ar flames at low pressure, using tunable synchrotron vacuum ultraviolet (VUV)
21
photoionization and molecular-beam mass spectrometry. A detailed kinetic model was proposed
22
to simulate these data. Mathieu et al. [8] studied the high-temperature oxidation of ammonia in a
23
shock tube at pressures of 1.4 30 atm and at highly diluted conditions. Shu et al. [9] studied the
24
auto-ignition behavior of ammonia in ‘air’ at intermediate temperatures using a shock tube.
25
Subsequently, He and Shu et al. further investigated the auto-ignition behavior of
26
ammonia/hydrogen blends [10] and ammonia/methane blends [11] at lower temperatures using a
27
3
rapid compression machine (RCM). Dai et al. studied the auto-ignition behavior of ammonia,
1
ammonia/hydrogen [12], ammonia/methane [13] and ammonia/DME [14] at low to intermediate
2
temperatures using an RCM, and proposed a kinetic model based on that from Glarborg et al. [15]
3
which showed good agreement with the data. Chen et al. [16] experimentally studied the auto-
4
ignition behavior of ammonia/hydrogen blends in a shock tube at 1.2 and 10 atm at highly diluted
5
conditions. The kinetic model proposed by Glarborg et al. [15] could simulate these data well.
6
The flame speed of ammonia with different fuel blends have been a focus of previous studies
7
over the years, as the significantly lower flame speed of ammonia is the major issue which
8
restricting its application in practical facilities. Okafor et al. [17] measured the laminar burning
9
velocities (LBVs) of ammonia/CH4/air mixtures. Han and Wang et al. [18], [19] studied the LBVs
10
of ammonia and ammonia blended with different fuels, including hydrogen, carbon monoxide,
11
methane and alcohol fuels. Mei et al. [20], [21] measured the LBVs of ammonia and
12
ammonia/syngas mixtures at different oxygen conditions and at elevated pressures. In these studies,
13
a number of kinetic models have been proposed and can simulate well these LBVs.
14
Currently, the ammonia/diesel dual fuel combustion strategy has been proved to be a promising
15
strategy to reduce engine CO2 emissions [22], [23]. Feng et al. [24] measured ignition delay times
16
(IDTs) of ammonia/diesel blends using an RCM at low-temperatures in the range of 670 – 910 K
17
at both 10 and 20 atm, with energy fraction of ammonia up to 50%. As n-heptane is one of the
18
primary reference fuels and is an important component of diesel and gasoline surrogate models, it
19
is essential to understand the chemical kinetics of ammonia/n-heptane blends. Yu et al. [25] studied
20
the auto-ignition behavior of ammonia/n-heptane blends in an RCM at 10 and 15 bar in the
21
temperature range of 600 1000 K. However, the kinetic model proposed in their study does not
22
capture well the low-temperature chemistry of ammonia/n-heptane blends, which is mainly
23
because some important interaction reaction pathways between ammonia and n-heptane are
24
missing as they state in their paper. Moreover, Lavadera et al. [26] experimentally studied the
25
effects of ammonia addition on LBVs of n-heptane and iso-octane in ‘air’ mixtures, and a semi-
26
detailed model which only includes the high-temperature oxidation of n-heptane and NOx was
27
used to simulate these LBV data [27], [28]. To the author’s knowledge, the model predictions of
28
ammonia/n-heptane IDTs available in the literature need to be improved at lower temperatures
29
(600–1000 K). Moreover, currently there is no shock tube IDT data measured for ammonia/n-
30
heptane blends at temperatures above 1000 K. These IDT data measured from the low- to high-
31
temperature range will be needed to explore the interaction reaction pathways between ammonia
32
and n-heptane.
33
4
In this study, a high-pressure shock tube (HPST) is used to measure the IDTs of different
1
ammonia/n-heptane blends in the high temperature range (1000–1400 K) and at 10 atm. A new
2
kinetic model is developed and validated against these new IDT data in addition to experimental
3
data available in the literature.
4
2. Experimental specifications
5
2.1. Ignition delay time measurements
6
A HPST made of stainless steel was used to conduct the experiments. The facility has been
7
described in detail in previous work [29], [30] and only a brief introduction is given here. The
8
shock tube has an internal diameter of 100 mm and a total length of 12.2 m, divided into three
9
parts: an 8.0 m long driven section, a 4.0 m long driver section and a 0.2 m long middle section.
10
These sections are separated by two polyester terephthalate (PET) diaphragms in the experiments.
11
During the experiments, the driver gas pressure in the middle section is first filled to half that used
12
in the driver section. The diaphragm is burst by venting the middle section. The gas temperature
13
and pressure behind the reflected shock wave were calculated using GasEq [31] based on the
14
measured incident shock velocity and initial conditions of temperature, pressure and mixture
15
composition.
16
Table 2: Experimental conditions studied for ammonia/n-heptane in the HPST.
17
Mixture
Mole fraction
Ratios of energy
φ p (bar)
T (K)
NH3 n-heptane
N2 O2
NH3 n-heptane
1 0 0.0095 0.8865
0.1040
0 1.0
1.0 10 1000 – 1400
2 0.0264
0.0074 0.8647
0.1015
0.2 0.8
3 0.0518
0.0055 0.8437
0.0990
0.4 0.6
4 0.0980
0.0103 0.7044
0.1873
0.4 0.6
The IDTs of ammonia/n-heptane were measured for fuel/O2/N2 mixtures at 10 bar in the
18
temperature range 1000 1400 K. The experimental conditions studied are summarized in Table
19
2, and these conditions are consistent with the RCM experiments performed by Yu et al. [25],
20
except that nitrogen is used as the diluent gas in the current experiments whereas Yu et al. [25]
21
used argon as the diluent. As suggested in previous studies [12], [32], ammonia has a strong
22
tendency to be absorbed by the steel surface. Therefore, the inner surfaces of the mixing tank and
23
the shock tube were passivated before the mixture preparation and each shot in the experiments.
24
For the passivation process, 100 mbar pure ammonia was introduced into the mixing tank and the
25
shock tube, maintained for about 10 min and then evacuated.
26
5
The definitions of IDTs measured in the HPST are shown in Fig. 1, which are based on the
1
measured pressure trace and OH* emission at 308 nm. The pressure trace used to determine the
2
IDTs is measured by the piezoelectric pressure transducer (PCB 113B24) installed in the sidewall,
3
which is embedded 20 mm away from the endwall. OH* emission is detected using a
4
photomultiplier, with a band-pass filter of 307 ± 10 nm, mounted at the endwall. The IDT is defined
5
as the time interval between the arrival of the reflected shock wave and the onset of the ignition,
6
and the onset of ignition corresponds to the intersection of the maximum slope of the OH*
7
emission profile back to the zero baseline. As the same methods were used for mixture preparation
8
and gas temperature and temperature calculation, a 20% uncertainty is assigned to all of the IDT
9
measurements in the HPST according to previous studies [33]–[35].
10
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
2.64% NH
3
/ 0.74% n-heptane / 10.15% O
2
(Mixture 2),
=
1.0
T
c
= 1150.1 K, p
c
= 10 bar,

= 1.194 ms
Signal
time / ms
Pressure
Normalized Emission
11
Figure 1. Definition of ignition delay time measured in the shock tube.
12
The purity of ammonia is > 99% and the purity of n-heptane is > 99%. High purity oxygen (>
13
99%), nitrogen (> 99%) were used in the experiments. The new IDT data obtained in this study
14
are provided in Table S1 of the Supplementary material.
15
3. Model development
16
The current kinetic model is developed based on NUIGMech1.2, in which the NOx chemistry is
17
mostly from Glarborg et al. [15] and the interaction reaction pathways between NOx and C1 C3
18
alkanes have been updated and well validated [36], [37]. The n-heptane sub-mechanism in
19
NUIGMech1.2 is originally taken from [38] and was constructed based on alkane rate rules [39].
20
The n-heptane sub-chemistry has been well validated previously [40], [41], and hence no further
21
updates were made to the n-heptane sub-chemistry in this study. The pyrolysis chemistry of
22
ammonia was updated based on the study of Alturaif et al. [42]. As suggested in their paper, the
23
updates of the ammonia pyrolysis chemistry significantly improve ammonia LBV predictions from
24
the Glarborg et al. [15] mechanism. Another major update of ammonia chemistry is the ṄH2 + 2
25
reaction system. Stagni et al. [43] theoretically calculated the rate constant for the reaction of NH3
26
6
+ O2 ṄH2 + 2. Cañas et al. [44] and Klippenstein et al. [45] theoretically studied the ṄH2 +
1
2 reaction system, and provided rate constants as well as branching ratios for all three reaction
2
pathways. The calculation results of Klippenstein et al. [45] suggest that ṄH2 + 2 mainly
3
proceeds via chain termination forming NH3 + O2. This is significantly different with most of the
4
rate constants in previous studies, which suggest that this channel mainly forms H2and ȮH
5
radicals. The rate constants of ṄH2 + 2 NH3 + O2 calculated in both the studies of Stagni et
6
al. [43] and Klippenstein et al. [45] are within a factor of two of one another at low temperatures
7
and are very similar, within 20% difference at high temperatures. In this study, the rate constant
8
calculated by Stagni et al. [43] are used. This update affects the prediction of low-temperature
9
chemistry of ammonia/n-heptane blends. Meanwhile, the rate constants of NH3+ȮH are increased
10
by a factor of two compared to the original very high temperature measurements by Salimian et al.
11
[46]. The most important updates in the current model are the H-atom abstraction reactions from
12
n-heptane by ṄH2. The importance of fuel + ṄH2 reactions have been discussed in a few previous
13
studies, for instance, CH4 + ṄH2 by Song et al. [47], DME + ṄH2 by Dai et al. [14] and Issayev et
14
al. [48]. The rate constants are estimated by analogy with CH4 + ṄH2 reactions measured by Song
15
et al. [47], with the activation energy decreased for 2 kcal mol–1 and 5 kcal mol–1 respectively for
16
primary and second carbon sites. These reactions are the major interaction reaction pathways
17
between the ammonia and n-heptane systems. As shown in Fig. 2, these interaction reactions
18
significantly increase the predicted fuel reactivity in the low- to intermediate-temperature regime.
19
The chemistry of how these interaction reactions affect fuel reactivity is discussed in detail in
20
Section 4.3. Other updates include the reactions between linear alkenes and ṄH2 radicals, with rate
21
constants estimated by analogy with ethylene + ṄH2 calculated by Mai et al. [49]. Some reactions
22
in the C-N interaction subsets are also updated based on new calculations. A list of all the important
23
updated reactions are provided in Table S2 of the Supplementary material.
24
0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6
0.1
1
10
100
RCM 10 bar
1st-stage IDT 10 bar
HPST 10 bar
Ignition delay time / ms
1000 K / T
1400 1200 1000 800
40%NH
3
/60%NC
7
H
16
/9.9%O
2
,

=
1.0
Temperature
(
K
)
25
Figure 2. Effects of the nC7H16 + ṄH2 reaction class on model predictions of ammonia/n-heptane
26
IDTs. Solid symbols: new HPST IDT data obtained in this study. Half-filled symbols and open
27
7
symbols represent the first-stage and second-stage RCM IDT data from Yu et al. [25]. Solid lines:
1
constant-volume simulations. Dashed lines: RCM simulations including facility effect. Short
2
dotted lines: constant-volume simulations without the nC7H16 + ṄH2 reaction class.
3
All of the simulations in this study were performed using Chemkin software [50]. The HPST
4
data are simulated assuming constant volume conditions. For the simulations of the RCM data
5
from Yu et al. [25] effective volume history profiles derived from the non-reactive pressure traces
6
were used to account for the volume change and the facility effects in the RCM (i.e. heat losses,
7
etc.). Moreover, flux and sensitivity analyses were performed using the Chemkin software to
8
understand the fuel chemistry.
9
4. Results and discussions
10
4.1. Model performance against experimental IDT data
11
Figure 3 shows the current model performance against the shock tube IDTs measured in this
12
study as well as the RCM IDTs taken by Yu et al. [25]. The blending effects of ammonia/n-heptane
13
over the low-temperature range were discussed in detail by Yu et al. In general, the ammonia
14
addition significantly slows down the fuel reactivity of ammonia/n-heptane blends in the low-
15
temperature regime, Fig. 3(a). However, at high temperatures, the experimental results show that
16
the IDTs of the ammonia/n-heptane blends slightly increase with increasing ammonia
17
concentration, and the current model can capture this trend well. This will be discussed in detail
18
in Section 4.3. With increasing oxygen concentration, the fuel reactivity of the ammonia/n-heptane
19
blends increase over the studied temperature range, Fig. 3(b). The fuel reactivity of the ammonia/n-
20
heptane blends also increases with increasing pressure and equivalence ratio in the low-
21
temperature regime.
22
0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6
0.1
1
10
100
Share by heating value
100% NC
7
H
16
80% NC
7
H
16
20% NH
3
60% NC
7
H
16
40% NH
3
Ignition delay time / ms
1000 K / T
(a)
16001400 1200 1000 800
=
1.0, p
c
=10 bar, ~ 10% O
2
Temperature
(
K
)
0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6
0.1
1
10
100
9.9% O
2
RCM
1st-stage IDT
HPST
18.37% O
2
RCM
1st-stage IDT
HPST
Ignition delay time / ms
1000 K / T
1400 1200 1000 800
40%NH
3
/60%NC
7
H
16
,

=
1.0, 10 bar
Temperature
(
K
)
(b)
23
8
1.0 1.1 1.2 1.3 1.4 1.5 1.6
10
100
40%NH
3
/60%NC
7
H
16
, ~85% dilution

= 1.0, 10 bar

= 2.0, 10 bar

= 1.0, 15 bar

= 2.0, 15 bar
Ignition delay time / ms
1000 K / T
(c)
1000 900 800 700
Temperature
(
K
)
1
Figure 3. Effects of (a) blending ratios, (b) dilution degree and (c)equivalence ratio and pressure
2
on ammonia/n-heptane IDTs. Solid symbols: new HPST IDT data obtained in this study. The half-
3
filled symbols and open symbols represent the first-stage and second-stage RCM IDT data from
4
Yu et al. [25]. Solid lines: constant-volume simulations. Dashed lines: RCM simulations including
5
facility effect. Dotted lines: RCM simulations of first-stage IDTs.
6
In the low- to intermediate temperature range, the current model slightly over-predicts the IDTs
7
at the relative high temperatures of the RCM data. This is due to the n-heptane sub-chemistry as
8
the under-predicted fuel reactivity is consistent for both pure n-heptane and ammonia/n-heptane
9
blends. Overall, the current model can capture well both the first-stage and second-stage IDTs at
10
different experimental conditions for different mixtures. The comparisons between the simulation
11
results using the current model and the previous model [25] are shown in Fig. S1 of the
12
Supplementary material. The current model predictions are improved significantly compared to
13
those of the original model in the low- to intermediate-temperature range.
14
Figure 4 shows the comparisons between the simulation results of the current model and the
15
experimental RCM IDTs of ammonia. As can be seen, the current model slightly over-predicts the
16
RCM IDTs at φ = 1.0 and 2.0, while the trend is well captured by the current model. Therefore, the
17
current model can capture well the IDTs of both pure n-heptane and ammonia, as well as their
18
blends over a wide temperature range.
19
0.82 0.84 0.86 0.88 0.90 0.92 0.94
10
100
Dai et al. 2020
= 0.5
= 1.0
= 2.0
Ignition delay time / ms
1000 K / T
1200 1175 1150 1125 1100 1075
100%NH
3
, 60 bar, 75% dilution
Temperature
(
K
)
20
9
Figure 4. Effects of equivalence ratio on ammonia IDTs. Open symbols: RCM data from Dai et
1
al. [12]. Dashed lines: RCM simulations including facility effect.
2
4.2. Model performance against experimental LBV data
3
The current model is also validated against experimental LBVs of ammonia and ammonia/n-
4
heptane blends available in the literature [26], Fig. 5. The LBVs of ammonia in ‘air’ at 298 K and
5
at 1 atm have been measured by different groups using different facilities over the years [18], [20],
6
[51], [52]. There was large discrepancy (~2 cm·s-1) between these experimental data of the fuel
7
rich conditions, Fig. 5(a). The peak value of LBVs measured at this condition is around 8 cm·s-1,
8
and the lower flame speed of ammonia compared to hydrocarbon fuels lead to large uncertainties
9
in the measurements. The simulation results of the current model match well with the experimental
10
results from Ronney et al. [51] and Mei et al. [20]. The n-heptane sub-chemistry of the current
11
model was validated previously against the LBVs measured by Sileghem et al. [53], open symbols
12
in Fig. 5(b). In a recent study, Lavadera et al. [26] corrected the measured data of Sileghem et al.
13
[53], half-filled symbols in Fig. 5(b). Lavadera et al. [26] also measured the LBVs of n-heptane
14
and ammonia/n-heptane blends, solid symbols in Fig. 5(b). As can be seen, the current model can
15
predict well the LBVs of n-heptane originally measured by Sileghem et al. [53], while the
16
simulation results are slightly slower compared to the LBVs measured by Lavadera et al. [26] and
17
the corrected results for the fuel lean conditions. For the ammonia/n-heptane blends, the LBVs of
18
blends decrease with increasing ammonia percentage. The equivalence ratios (φ) of the peak flame
19
speed for different fuel blends are around φ = 1.1. The current model slightly under-predicts the
20
LBVs at the fuel lean conditions, which is consistent with that of the pure n-heptane. The
21
simulation results without the nC7H16 + ṄH2 reaction class are also given in Fig. 5(b). It is observed
22
that this reaction class only affects the fuel-lean predicted LBVs when the mole fraction of
23
ammonia is 50%. However, the energy fraction from ammonia for this condition is only 5.9%,
24
which indicates that more LBV data with higher mole fractions of ammonia are needed for model
25
validation. Overall, the current model predicts the LBVs of ammonia and different ammonia/n-
26
heptane blends reasonably well at the test conditions.
27
10
0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7
0
2
4
6
8
10
Ronney et al. 1988
Mei et al. 2019
Jabbour et al. 2004
Han et al. 2019
Laminar Burning Velocity (cm/s)
Equivalence ratio
Ammonia in 'air', T
u
=298 K, 1 atm
(a)
0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4
0
10
20
30
40
50
60
(b)
Laminar Burning Velocity (cm/s)
Equivalence ratio
Mole fractions of fuels
100% NC
7
H
16
, Sileghem et al. 2013
100% NC
7
H
16
, Lavadera et al. 2021
75% NC
7
H
16
/ 25% NH
3
, Lavadera et al. 2021
50% NC
7
H
16
/ 50% NH
3
, Lavadera et al. 2021
Ammonia/n-heptane in 'air', T
u
=338 K, 1 atm
1
Figure 5. Experimental and simulated results of LBVs for (a) ammonia, experimental data from
2
Ref. [18],[20],[51],[52] and (b) n-heptane and ammonia/n-heptane blends. Solid and half-filled
3
symbols from Ref. [26] and open symbols from Ref. [53]. Symbols represent experimental data
4
and solid lines represent simulations based on current model. The dashed lines represent
5
simulations without the reaction class of nC7H16 + ṄH2.
6
4.3. Flux and Sensitivity analyses using the current model
7
Figure 6 shows the species traces during the auto-ignition process of the 40% ammonia/60% n-
8
heptane (by energy) blends (Mixture 4) at an initial pressure p = 10 atm and at an initial
9
temperature T = 650 K. Over 70% of n-heptane is consumed during the first-stage heat release,
10
while only around 4% of ammonia is consumed at this stage. Therefore, the first stage heat release
11
is dominated by the oxidation of n-heptane. The mole fractions of HȮ2 and ȮH radicals increase
12
during the first-stage heat release, and H2O2 accumulates at this stage. The temperature increases
13
slowly after the first-stage heat release, and consequently the decomposition of H2O2 is accelerated
14
at a temperature of approximately 900 K. At this stage almost all of the n-heptane is consumed,
15
and ammonia starts to be consumed quickly due to the formation of ȮH radicals.
16
0.00
0.01
0.02
0.03
0.04
0.05
0.06
Mole fraction
Mole fraction
NC7H16*10
NH3
Temperature
Mole fraction
700
800
900
1000
1100
1200
Temperature / K
0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 0.09
0.0
2.0x10
-7
4.0x10
-7
Time
(
s
)
( )
HO
2
0.01
OH
( )
( )
0.0
5.0x10
-4
1.0x10
-3
H
2
O
2
17
Figure 6. Temperature and species mole fraction traces for 40% ammonia/60% n-heptane (by
18
energy) at φ = 1.0 and ~85% dilution (Mixture 4), with initial pressure p = 10 atm and initial
19
temperature T = 650 K. Solid symbols: 20% fuel consumption timing.
20
11
To further investigate the reaction pathways, flux and sensitivity analyses were performed for
1
Mixture 4 at 10 atm and at different temperatures using the Chemkin [50] software, and the results
2
are depicted in Figs. 7 and 8. The three temperatures are selected to represent different temperature
3
regimes controlled by different reaction pathways, with 650 K and 1200 K representing low and
4
high temperature regime and 750 K representing the negative temperature coefficient (NTC)
5
region. As the low-temperature reaction pathways of n-heptane are well known, only simplified
6
flux analyses results are given here. At 650 K, more than 90% of n-heptane is consumed via H-
7
atom abstraction by ȮH radicals, and around 10% of n-heptane is consumed via H-atom
8
abstraction by ṄH2 leading to the formation of ammonia and heptyl radicals. The heptyl radicals
9
so formed mainly add to O2 leading to the formation of RȮ2 radicals. These radicals subsequently
10
undergo isomerization and 2 elimination reactions, similar to the low-temperature chemistry of
11
pure n-heptane. For 2 species formed in the low-temperature chemistry of n-heptane, the
12
reactions with ṄO forming 2 and alkyl oxide radicals are also included. However, these reaction
13
pathways are not important at this condition as the competing channels of 2 isomerization to
14
Q
OOH have lower energy barriers.
15
The ṄH2 radicals are mainly formed via H-atom abstraction from ammonia by ȮH radicals. As
16
only ~4% of ammonia is consumed at this stage, ȮH and HȮ2 radicals are mainly formed during
17
the oxidation of n-heptane. As discussed above, H-atom abstraction reactions from n-heptane by
18
ṄH2 significantly increase the predicted fuel reactivity of the ammonia/n-heptane blends in the
19
low- to intermediate-temperature regime. These reactions are the major interaction reaction
20
pathways between ammonia and n-heptane in the low- to intermediate-temperature range. The
21
ṄH2 radicals can also react with 2 radicals forming ammonia and O2, which is chain terminating
22
and hence slows down the consumption rate of ammonia. Meanwhile, about 10% of ṄH2 radicals
23
undergo recombination forming N2H4. Over 30% of ṄH2 radicals react with 2 and 2 radicals
24
forming H2radicals, which further react with 2 and forms HONO. The HONO then react
25
with ṄH2 forming ammonia and NȮ2.
26
12
1
Figure 7. Flux analyses for 40% ammonia / 60% n-heptane (by energy) at φ = 1.0 and ~85%
2
dilution (Mixture 4), p = 10 atm and 20% n-heptane consumed. Numbers represent the percentage
3
of fuel flux that goes into a particular species. Black numbers represent flux at 650 K, blue numbers
4
represent flux at 750 K, and red numbers represent flux at 1200 K. Ṙ is the sum of ȮH, HȮ2 and
5
ĊH3 radicals and Ḣ and Ö atoms.
6
At 750 K, the flux of H-atom abstraction from n-heptane by ṄH2 increases. To simulate the
7
ammonia/n-heptane blends, adding the reaction pathways of H-atom abstraction by ṄH2 radicals
8
promotes the consumption of n-heptane and reduces the flux to chain termination, and
9
consequently increases the predicted fuel reactivity of the ammonia/n-heptane blends in the low-
10
to intermediate-temperature range. The reactivity of n-heptane in the NTC region depends on the
11
competition between chain branching (first and second Q
OOH addition to O2) and chain
12
propagation (HȮ2 elimination from 2 species) reaction pathways. The HȮ2 radicals so formed
13
promote the chain terminating pathway of ṄH2 with 2 forming ammonia and oxygen, and hence
14
decrease the reactivity. This explains why ammonia addition has large effects on the fuel reactivity
15
of ammonia/n-heptane blends in the NTC region.
16
At 1200 K, over 58% of ṄH2 radicals react with n-heptane producing heptyl radicals and
17
ammonia, and hence fewer ṄH2 radicals undergo chain termination reaction pathways. This
18
explains why ammonia fraction has relatively smaller effects on fuel reactivity at high
19
temperatures than that at lower temperatures.
20
Figure 8 shows brute force sensitivity analyses for IDTs of ammonia/n-heptane blends at
21
different temperatures. The sensitivity coefficient, Sτ, is defined as: = ln(
) /ln (
),
22
which is computed for every reaction by perturbing the A-factor by factors of 2.0 and 0.5. Here
23
is the original pre-exponential factor of the target reaction; and are the factor is multiplied
24
by 2.0 and 0.5, respectively; and are the corresponding IDTs obtained with and ,
25
13
respectively. Positive values indicate reactions that inhibit reactivity and negative values
1
indicate reactions that promote reactivity. The reaction pathways highlighted in the sensitivity
2
analysis results are consistent with those in the flux analysis results. At 650 K, the reaction
3
pathways of H-atom abstraction from n-heptane forming Ċ7H15-2 and Ċ7H15-4 radicals promote
4
fuel reactivity, as these radicals mainly undergo chain branching via and Q
OOH radical
5
additions to O2. Both the reaction channels of ṄH2 radicals with 2 and 2 forming H2
6
radicals promotes reactivity as they compete with the two main chain terminating reaction
7
pathways. The most sensitive reaction inhibiting fuel reactivity is H-atom abstraction from
8
ammonia by ȮH radicals, and the ṄH2 radicals so formed react with HȮ2 producing NH3 and O2,
9
which also inhibits reactivity. At 750 K, both the channels of ṄH2 radicals reacting with NȮ2 and
10
2 producing H2radicals promote reactivity, which is similar to that at 650 K. Q
OOH radical
11
addition to O2 also promotes reactivity as these lead to chain branching. The most sensitive
12
reactions which inhibit reactivity are the same as those at 650 K. At 1200 K, the sensitivity analysis
13
results show that H-atom abstraction from ethylene by ṄH2 significantly promotes fuel reactivity.
14
This is because a high concentration of ethylene is produced during the oxidation of n-heptane at
15
high temperatures, and the vinyl radicals formed after H-atom abstraction from ethylene can
16
undergo chain branching reaction pathways by reacting with O2. It is interesting to see that the
17
reaction of vinyl radicals reacting with oxygen producing formaldehyde, Ḣ and CO, R315, slightly
18
decreases fuel reactivity. This is because this channel competes with another channel forming Ö
19
atoms, R310.
20
Ċ2H3 + O2 CH2O + + CO R315
21
Ċ2H3 + O2 ĊH2CHO + Ö R310
22
The reaction of ṄH2 and HȮ2 radicals producing NH3 and O2 inhibit fuel reactivity in the low-
23
to high-temperature range.
24
14
-0.6 -0.4 -0.2 0.0 0.2 0.4 0.6
NH
3
+OH<=>NH
2
+H
2
O
NH
3
+O
2
<=>NH
2
+HO
2
HO
2
+HO
2
<=>H
2
O
2
+O
2
HCO+O
2
<=>CO+HO
2
NH
2
+NO
2
<=>N
2
O+H
2
O
OH+NO<=>HONO
NC
7
H
16
+OH<=>C
7
H
15
-2+H
2
O
NH
2
+HO
2
<=>OH+H
2
NO
NH
2
+NO
2
<=>H
2
NO+NO
Sensitivity Coefficient
NC
7
H
16
+OH<=>C
7
H
15
-4+H
2
O
= 1.0, 10 atm, 650 K
(a)
-0.6 -0.4 -0.2 0.0 0.2 0.4 0.6
NH
3
+O
2
<=>NH
2
+HO
2
NH
3
+OH<=>NH
2
+H
2
O
CH
3
NO
2
(+M)=CH
3
+NO
2
(+M)
NH
2
+NO
2
<=>N
2
O+H
2
O
C
7
H
15
-4O
2
<=>C
7
H
14
-3+HO
2
C
7
H
14
OOH4-2+O
2
<=>C
7
H
14
OOH4-2O
2
C
7
H
14
OOH2-4+O
2
<=>C
7
H
14
OOH2-4O
2
OH+NO<=>HONO
NH
2
+NO
2
<=>H
2
NO+NO
Sensitivity Coefficient
NH
2
+HO
2
<=>OH+H
2
NO
= 1.0, 10 atm, 750 K
(b)
1
-0.6 -0.4 -0.2 0.0 0.2 0.4 0.6
NH
3
+O
2
<=>NH
2
+HO
2
NH
3
+OH<=>NH
2
+H
2
O
C
2
H
3
+O
2
=>CH
2
O+H+CO
CH
3
+HO
2
<=>CH
4
+O
2
NC
7
H
16
+H<=>C
7
H
15
-3+H
2
C
2
H
5
+O
2
<=>C
2
H
4
+HO
2
NC
7
H
16
<=>PC
4
H
9
+NC
3
H
7
NH
2
+HO
2
<=>OH+H
2
NO
O
2
+H<=>O+OH
Sensitivity Coefficient
C
2
H
4
+NH
2
<=>C
2
H
3
+NH
3
= 1.0, 10 atm, 1200 K
(c)
2
Figure 8. Sensitivity analyses for IDTs of 40% ammonia / 60% n-heptane (by energy) at φ = 1.0
3
and ~85% dilution (Mixture 4), with p = 10 atm and T = 650, 750, 1200 K. The positive values
4
indicate reactions that inhibit reactivity and negative values promote reactivity.
5
The LBV sensitivity analyses for the 50% ammonia/50% n-heptane blends at φ = 0.8 and φ =
6
1.1 are given as Fig. S2 of the Supplementary material. The most sensitive reactions are from the
7
core hydrocarbon chemistry. The H-atom abstraction reactions from n-heptane by ṄH2 radicals
8
and the overall ammonia chemistry are less important in simulating LBVs of the fuel-rich mixture
9
for ammonia/n-heptane blends at the condition studied.
10
5. Conclusions
11
(1) In this study, a high-pressure shock tube was used to measure the IDTs of different ammonia/n-
12
heptane blends in the high temperature regime. By comparing the new ST IDTs with the RCM
13
IDTs available in the literature, the experimental results show that ammonia fraction has a relative
14
smaller effect on the IDTs at high temperatures compared to the low- to intermediate-temperature
15
range.
16
(2) A new kinetic model for ammonia/n-heptane blends was developed and validated against the
17
new measured ST IDTs in addition to existing experimental data available in the literature. Overall,
18
the current model can predict well both the auto-ignition behavior and flame speeds of pure
19
ammonia and different ammonia/n-heptane blends. The current model performance shows
20
significant improvement compared to the previous ammonia/n-heptane model in the literature.
21
(3) The flux and sensitivity analysis results highlight the importance of interaction reactions of H-
22
atom abstraction from n-heptane by ṄH2, which significantly increase the predicted fuel reactivity
23
15
in the low- to intermediate-temperature range. Moreover, this interaction reaction class also
1
improves the predicted LBVs of the fuel-lean mixture of the ammonia/n-heptane blends, but only
2
when the mole fraction of ammonia is higher than 50%.
3
Acknowledgements
4
5
6
16
References
1
[1] Robert F. Service. Ammonia made from sun, air, and water could turn Australia into a
2
renewable energy. Science, 2018, 361(6398): 120-123.
3
[2] A. Valera-Medina, H. Xiao, M. Owen-Jones, et al. Ammonia for power. Progress in Energy
4
and Combustion Science, 2018, 69: 63-102.
5
[3] P. Dirrenberger, P.A. Glaude, R. Bounaceur, H. Le Gall, A.P. Da Cruz, A.A. Konnov, F.
6
Battin-Leclerc, Laminar burning velocity of gasolines with addition of ethanol, Fuel. 115 (2014)
7
162–169.
8
[4] Y. Xu, T.I. Farouk, M.C. Hicks, C.T. Avedisian, Initial diameter effects on combustion of
9
unsupported equi-volume n-heptane/iso-octane mixture droplets and the transition to cool
10
flame behavior: Experimental observations and detailed numerical modeling, Combust. Flame.
11
220 (2020) 82–91.
12
[5] T. Hara, K. Tanoue, Laminar flame speeds of ethanol, n-heptane, iso-octane air mixtures,
13
FISITA 2006 World Automot. Congr. Yokohama, Japan. (2006) Technical Report F2006SC40.
14
[6] R. Li, Z. Liu, Y. Han, M. Tan, Y. Xu, J. Tian, J. Chai, J. Liu, Extended adiabatic flame
15
temperature method for lower flammability limits prediction of fuel-air-diluent mixture by
16
nonstoichiometric equation and nitrogen equivalent coefficients, Energy and Fuels. 31 (2017)
17
351–361.
18
[7] Z. Tian, Y. Li, L. Zhang, P. Glarborg, F. Qi, An experimental and kinetic modeling study of
19
premixed NH3/CH4/O2/Ar flames at low pressure, Combust. Flame. 156 (2009) 1413–1426.
20
[8] O. Mathieu, E.L. Petersen, Experimental and modeling study on the high-temperature
21
oxidation of Ammonia and related NOx chemistry, Combust. Flame. 162 (2015) 554–570.
22
[9] B. Shu, S.K. Vallabhuni, X. He, G. Issayev, K. Moshammer, A. Farooq, R.X. Fernandes, A
23
shock tube and modeling study on the autoignition properties of ammonia at intermediate
24
temperatures, Proc. Combust. Inst. 37 (2019) 205–211.
25
[10] X. He, B. Shu, D. Nascimento, K. Moshammer, M. Costa, R.X. Fernandes, Auto-ignition
26
kinetics of ammonia and ammonia/hydrogen mixtures at intermediate temperatures and high
27
pressures, Combust. Flame. 206 (2019) 189–200.
28
[11] B. Shu, X. He, C.F. Ramos, R.X. Fernandes, M. Costa, Experimental and modeling study
29
on the auto-ignition properties of ammonia/methane mixtures at elevated pressures, Proc.
30
Combust. Inst. 38 (2021) 261–268.
31
[12] L. Dai, S. Gersen, P. Glarborg, H. Levinsky, A. Mokhov, Experimental and numerical
32
analysis of the autoignition behavior of NH3 and NH3/H2 mixtures at high pressure, Combust.
33
Flame. 215 (2020) 134–144.
34
17
[13] L. Dai, S. Gersen, P. Glarborg, A. Mokhov, H. Levinsky, Autoignition studies of NH3/CH4
1
mixtures at high pressure, Combust. Flame. 218 (2020) 19–26.
2
[14] L. Dai, H. Hashemi, P. Glarborg, S. Gersen, P. Marshall, A. Mokhov, H. Levinsky, Ignition
3
delay times of NH3/DME blends at high pressure and low DME fraction: RCM experiments
4
and simulations, Combust. Flame. 227 (2021) 120–134.
5
[15] P. Glarborg, J.A. Miller, B. Ruscic, S.J. Klippenstein, Modeling nitrogen chemistry in
6
combustion, Prog. Energy Combust. Sci. 67 (2018) 31–68.
7
[16] J. Chen, X. Jiang, X. Qin, Z. Huang, Effect of hydrogen blending on the high temperature
8
auto-ignition of ammonia at elevated pressure, Fuel. 287 (2021) 119563.
9
[17] E.C. Okafor, Y. Naito, S. Colson, A. Ichikawa, T. Kudo, A. Hayakawa, H. Kobayashi,
10
Experimental and numerical study of the laminar burning velocity of CH4–NH3–air premixed
11
flames, Combust. Flame. 187 (2018) 185–198.
12
[18] X. Han, Z. Wang, M. Costa, Z. Sun, Y. He, K. Cen, Experimental and kinetic modeling
13
study of laminar burning velocities of NH3/air, NH3/H2/air, NH3/CO/air and NH3/CH4/air
14
premixed flames, Combust. Flame. 206 (2019) 214–226.
15
[19] Z. Wang, X. Han, Y. He, R. Zhu, Y. Zhu, Z. Zhou, K. Cen, Experimental and kinetic study
16
on the laminar burning velocities of NH3 mixing with CH3OH and C2H5OH in premixed
17
flames, Combust. Flame. 229 (2021) 111392.
18
[20] B. Mei, X. Zhang, S. Ma, M. Cui, H. Guo, Z. Cao, Y. Li, Experimental and kinetic
19
modeling investigation on the laminar flame propagation of ammonia under oxygen
20
enrichment and elevated pressure conditions, Combust. Flame. 210 (2019) 236–246.
21
[21] B. Mei, S. Ma, Y. Zhang, X. Zhang, W. Li, Y. Li, Exploration on laminar flame
22
propagation of ammonia and syngas mixtures up to 10 atm, Combust. Flame. 220 (2020) 368–
23
377.
24
[22] A.J. Reiter, S.C. Kong, Combustion and emissions characteristics of compression-ignition
25
engine using dual ammonia-diesel fuel, Fuel. 90 (2011) 87–97.
26
[23] A. Yousefi, H. Guo, S. Dev, B. Liko, S. Lafrance, Effects of ammonia energy fraction and
27
diesel injection timing on combustion and emissions of an ammonia/diesel dual-fuel engine,
28
Fuel. 314 (2022) 122723.
29
[24] Y. Feng, J. Zhu, Y. Mao, M. Raza, Y. Qian, L. Yu, X. Lu, Low-temperature auto-ignition
30
characteristics of NH3/diesel binary fuel: Ignition delay time measurement and kinetic analysis,
31
Fuel. 281 (2020) 118761.
32
18
[25] L. Yu, W. Zhou, Y. Feng, W. Wang, J. Zhu, Y. Qian, X. Lu, The effect of ammonia addition
1
on the low-temperature autoignition of n-heptane: An experimental and modeling study,
2
Combust. Flame. 217 (2020) 4–11.
3
[26] M.L. Lavadera, X. Han, A.A. Konnov, Comparative effect of ammonia addition on the
4
laminar burning velocities of methane, n-heptane, and iso-octane, Energy and Fuels. 35 (2021)
5
7156–7168.
6
[27] M. Pelucchi, M. Bissoli, C. Cavallotti, A. Cuoci, T. Faravelli, A. Frassoldati, E. Ranzi, A.
7
Stagni, Improved kinetic model of the low-temperature oxidation of n-heptane, Energy and
8
Fuels. 28 (2014) 7178–7193.
9
[28] Y. Song, L. Marrodán, N. Vin, O. Herbinet, E. Assaf, C. Fittschen, A. Stagni, T. Faravelli,
10
M.U. Alzueta, F. Battin-Leclerc, The sensitizing effects of NO 2 and NO on methane low
11
temperature oxidation in a jet stirred reactor, Proc. Combust. Inst. 37 (2019) 667–675.
12
[29] C. Yang, W. Wang, Y. Li, X. Cheng, Quantitative study on chemical effects of
13
actual/simulated recirculated exhaust gases on ignition delay times of n-heptane /ethanol fuel
14
blends at elevated temperature, Fuel. 263 (2020) 116327.
15
[30] Y. Liu, C. Zou, J. Cheng, H. Jia, C. Zheng, Experimental and Numerical Study of the Effect
16
of CO2 on the Ignition Delay Times of Methane under Different Pressures and Temperatures,
17
Energy and Fuels. 32 (2018) 10999–11009.
18
[31] C. Morley, GasEq, Version 0.76. http://www.gaseq.co.uk, 2004.
19
[32] O. Mathieu, M.M. Kopp, E.L. Petersen, Shock-tube study of the ignition of multi-
20
component syngas mixtures with and without ammonia impurities, Proc. Combust. Inst. 34
21
(2013) 3211–3218.
22
[33] E.L. Petersen, M.J.A.A. Rickard, M.W. Crofton, E.D. Abbey, M.J. Traum, D.M. Kalitan,
23
A facility for gas- and condensed-phase measurements behind shock waves, Meas. Sci.
24
Technol. 16 (2005) 1716–1729.
25
[34] Y. Li, C.W. Zhou, H.J. Curran, An extensive experimental and modeling study of 1-butene
26
oxidation, Combust. Flame. 181 (2017) 198–213.
27
[35] S. Dong, K. Zhang, E.M. Ninnemann, A. Najjar, G. Kukkadapu, J. Baker, F. Arafin, Z.
28
Wang, W.J. Pitz, S.S. Vasu, S.M. Sarathy, P.K. Senecal, H.J. Curran, A comprehensive
29
experimental and kinetic modeling study of 1- and 2-pentene, Combust. Flame. 223 (2021)
30
166–180.
31
[36] A.B. Sahu, A.A.E.S. Mohamed, S. Panigrahy, C. Saggese, V. Patel, G. Bourque, W.J. Pitz,
32
H.J. Curran, An experimental and kinetic modeling study of NOx sensitization on methane
33
autoignition and oxidation, Combust. Flame. 238 (2022) 111746.
34
19
[37] A.A.E.S. Mohamed, S. Panigrahy, A.B. Sahu, G. Bourque, H. Curran, The effect of the
1
addition of nitrogen oxides on the oxidation of ethane: An experimental and modelling study,
2
Combust. Flame. 241 (2022) 112058.
3
[38] K. Zhang , C. Banyon , J. Bugler , H.J. Curran , A. Rodriguez , O. Herbinet , F. Bat- tin-
4
Leclerc , C. B’Chir , K.A. Heufer , An updated experimental and kinetic modeling study of n-
5
heptane oxidation, Combust. Flame 172 (2016) 116–135.
6
[39] K. Zhang, C. Banyon, U. Burke, G. Kukkadapu, S.W. Wagnon, M. Mehl, H.J. Curran, C.K.
7
Westbrook, W.J. Pitz, An experimental and kinetic modeling study of the oxidation of hexane
8
isomers: Developing consistent reaction rate rules for alkanes, Combust. Flame. 206 (2019)
9
123–137.
10
[40] A.A. El-Sabor Mohamed, S. Panigrahy, A.B. Sahu, G. Bourque, H.J. Curran, An
11
experimental and kinetic modeling study of the auto-ignition of natural gas blends containing
12
C1-C7 alkanes, Proc. Combust. Inst. 38 (2021) 365–373.
13
[41] Y. Wu, S. Panigrahy, A.B. Sahu, C. Bariki, J. Beeckmann, J. Liang, A.A.E. Mohamed, S.
14
Dong, C. Tang, H. Pitsch, Z. Huang, H.J. Curran, Understanding the antagonistic effect of
15
methanol as a component in surrogate fuel models: A case study of methanol / n-heptane
16
mixtures, Combust. Flame. 226 (2021) 229–242.
17
[42] S.A. Alturaifi, O. Mathieu, E.L. Petersen, An experimental and modeling study of
18
ammonia pyrolysis, Combust. Flame. (2021) 111694.
19
[43] A. Stagni, C. Cavallotti, S. Arunthanayothin, Y. Song, O. Herbinet, F. Battin-Leclerc, T.
20
Faravelli, An experimental, theoretical and kinetic-modeling study of the gas-phase oxidation
21
of ammonia, React. Chem. Eng. 5 (2020) 696–711.
22
[44] J.E. Chavarrio Cañas, M. Monge-Palacios, X. Zhang, S.M. Sarathy, Probing the gas-phase
23
oxidation of ammonia: Addressing uncertainties with theoretical calculations, Combust. Flame.
24
235 (2022) 111708.
25
[45] S.J. Klippenstein, P. Glarborg, Theoretical kinetics predictions for NH2 + HO2, Combust.
26
Flame. 236 (2022) 111787.
27
[46] S. Salimian, R.K. Hanson, C.H. Kruger, High temperature study of the reactions of O and
28
OH with NH3, Int. J. Chem. Kinet. 16 (1984) 725–739.
29
[47] S. Song, D.M. Golden, R.K. Hanson, C.T. Bowman, J.P. Senosiain, C.B. Musgrave, G.
30
Friedrichs, A shock tube study of the reaction NH2 + CH4 NH3 + CH3 and comparison
31
with transition state theory, Int. J. Chem. Kinet. 35 (2003) 304309.
32
20
[48] G. Issayev, B.R. Giri, A.M. Elbaz, K.P. Shrestha, F. Mauss, W.L. Roberts, A. Farooq,
1
Ignition delay time and laminar flame speed measurements of ammonia blended with dimethyl
2
ether: A promising low carbon fuel blend, Renew. Energy. 181 (2022) 1353–1370.
3
[49] T.V.T. Mai, M. v. Duong, L.K. Huynh, Theoretical kinetics of the C2H4 + NH2 reaction,
4
Combust. Flame. 215 (2020) 193–202.
5
[50] CHEMKIN-PRO 15101, Reaction Design, San Diego, 2010.
6
[51] P.D. Ronney, Effect of Chemistry and Transport Properties on Near-Limit Flames at
7
Microgravity, Combust. Sci. Technol. 59 (1988) 123–141.
8
[52] T. Jabbour, D.F. Clodic, Burning velocity and refrigerant flammability classification,
9
ashrae transactions, Atlanta 110 (2004) 522–533.
10
[53] L. Sileghem, V.A. Alekseev, J. Vancoillie, K.M. Van Geem, E.J.K. Nilsson, S. Verhelst,
11
A.A. Konnov, Laminar burning velocity of gasoline and the gasoline surrogate components
12
iso-octane, n-heptane and toluene, Fuel. 112 (2013) 355–365.
13
14
... Nonetheless, the practical utilization of pure NH 3 as a fuel still faces several challenges, such as high autoignition temperature, low burning velocity, and high NO emissions [4,5]. To address these problems, an effective method is the dual-fuel combustion strategy, which involves blending low-reactivity NH 3 with a high-reactivity fuel [6]. ...
... Therefore, the NH 3 sub-mechanism is directly taken from Liu et al. [12]. The reactions of fuel + NH 2 play a key role during the ignition process of NH 3 blended binary fuels [6,10,18,[22][23][24]. In this work, the rate constants of the MHX + NH 2 reactions are estimated by reducing the activation energy by 2 and 5 kcal mol −1 for the primary and secondary carbon based on the CH 4 + NH 2 reactions [55], respectively. ...
... In this work, the rate constants of the MHX + NH 2 reactions are estimated by reducing the activation energy by 2 and 5 kcal mol −1 for the primary and secondary carbon based on the CH 4 + NH 2 reactions [55], respectively. The rate constants of the interaction reactions between MHX and NO 2 , leading to the production of HONO, assume those at the same sites of -heptane (nC 7 H 16 ) [6]. Since the generation of HNO 2 was not considered by Dong et al. [6], the rate constants for reactions di-isopropyl ketone (C 7 H 14 O) + NO 2 [29] are taken for the values for MHX + NO 2 reactions. ...
Article
Full-text available
Ammonia (NH 3) is a promising carbon-free and alternative fuel, but its applications are hindered by high auto-ignition temperature and low burning velocity. A common approach to overcome such drawbacks is to blend NH 3 with a high-reactivity fuel. In this study, a heated shock tube is employed to measure ignition delay time of NH 3 blended with methyl hexanoate (MHX). The experiments are conducted at 6 atm, equivalence ratios of 0.5-2.0, temperatures of 1168-2115 K, and MHX blending ratios of 0, 20%, 50%, 70%, and 100%. Ignition delay time of the binary mixtures decreases monotonically with the addition of MHX. Compared with pure NH 3 , the reactivity of the binary mixtures increases significantly with the addition of only 20% MHX, leading to a 10 times faster ignition delay time at around 1500 K and 6 atm. The reactivity of the fuel-lean and stoichiometric ratio mixtures is similar, and higher than the fuel-rich mixtures. The promotion effect of ignition delay time decreases with increasing blending ratio and pressure, and decreasing temperature. The influence of equivalence ratio on the promotion effect of ignition delay time is less significant than that of blending ratio, temperature and pressure. A detailed NH 3 /MHX kinetic model is developed by updating the interaction reactions between MHX and NH 2 /NO 2 /NO radicals, and the NH 3 and MHX sub-mechanism. The present kinetic model can reproduce satisfactorily the ignition delay time of pure MHX and NH 3 , and the NH 3 /MHX mixtures in the whole experimental conditions explored here. The kinetic analyses reveal that the interaction reactions between MHX and NH 2 radical have a significant impact on the ignition of the binary mixtures. Moreover, the important intermediate N 2 H 2 is more prone to forming N 2 H 3 rather than NNH in the presence of MHX, different from the production of NNH in pure NH 3 combustion. The H-atom abstraction reaction, NH 3 , H + NH 3 = NH 2 + H 2 , can proceed in the reverse direction with the addition of MHX, resulting in the production of more active H radicals that facilitate ignition. Novelty and significance statement: The practical utilization of pure ammonia (NH 3) as a fuel still faces several challenges and an effective method is the dual-fuel combustion strategy which involves blending low-reactivity NH 3 with a high-reactivity fuel. This work measures the new ignition delay time of methyl hexanoate (MHX) and NH 3 /MHX mixtures. A newly detailed NH 3 /MHX kinetic model is also developed by updating the interaction reactions between MHX and NH 2 /NO 2 /NO radicals, and the NH 3 and MHX sub-mechanism. The kinetic analyses reveal that the interaction reactions between MHX and NH 2 radical have a significant impact on the ignition of the binary mixtures and the important intermediate N 2 H 2 is more likely to form N 2 H 3 rather than NNH in the presence of MHX. To our best knowledge, this is the first study on the effect of methyl ester MHX addition on the ignition behavior of NH 3. * Corresponding authors.
... This mechanism can predict the inhibition effect of ammonia on the ignition of n-heptane, but cannot predict the IDTs of diesel/ammonia mixtures over a wide range of temperatures, especially overpredict the first-stage ignition and the IDTs in the NTC region. Dong et al. 23 investigated the IDTs of ammonia/diesel mixtures in a shock tube. The experimental results showed that the IDTs of ammonia/n-heptane mixtures were prolonged with increasing ammonia mole fraction or decreasing oxygen concentration. ...
... Moreover, the interaction between NH 3 and C1-C2 is included in the Glarborg mechanism, and it does not need to be added to the merge mechanism anymore. Since the reaction n-C 7 H 16 + NH 2 <=> C 7 H 15 + NH 3 has been proven to be important for the ignition processes of ammonia/n-heptane mixtures, 23 the related reactions are added in the new mechanism. ...
... Figure 8a shows that the simulated IDTs of NC 7 H 16 agree well with experimental data over the entire temperature range. 22,23 At n-heptane energy fractions of 80% and 60%, the predicted IDTs of NC 7 H 16 / NH 3 mixtures are in good agreement with experimental data at medium temperatures, while they are slightly higher than experimental data at low and high temperatures. Figure 8B shows that, at 15 bar, the predicted IDTs of NC 7 H 16 agree well with experimental data. ...
Article
Full-text available
Pilot‐ignited ammonia‐fueled engines have drawn more and more attention for low carbon emissions compared to traditional diesel engines. The ignition processes of NH3/NC7H16 mixtures under compression ignition engine‐like conditions are numerically investigated. By comparing the ignition delay times (IDTs) calculated by six ammonia mechanisms with experimental data, the Glarborg mechanism is selected. Then, the Glarborg mechanism and the Zhang detailed n‐heptane mechanism are merged into a new mechanism, which is adopted in the present study. Results show that the negative temperature coefficient behavior of the IDTs is only observed as the ammonia mass fraction is 70%. Only temperature has a significant effect on IDTs at all research conditions, and the effect of ammonia mass fraction is significant when the temperature is lower than 1000 K. However, the effects of equivalence ratio and pressure are small, especially at high temperatures, high equivalence ratios, and high pressures. Interestingly, the IDTs are categorized into three regions by temperature and ammonia mass fraction. The sensitivity analysis indicates that the sensitivity coefficients of most reactions associated with ammonia decrease with a decrease in ammonia mass fraction, whereas only R4210 is sensitive to ammonia mass fraction for n‐heptane‐related reactions. Rates of production and consumption (ROP) analyses indicate that the ammonia mass fraction mainly affects the ROPs of NC7H16, NH3, and NNH at low and medium temperatures, whereas the ammonia mass fraction affects the ROP of H2NO before the temperature of 2000 K. The ROPs of NC7H16, NH3, and NNH significantly increase with increasing temperature, whereas the ROP of H2NO slightly increases with increasing temperature. The increase of temperature in the early and middle stages is mainly contributed by the oxidation of n‐heptane, while the increase of temperature in the middle and late stages is mainly contributed by the oxidation of ammonia.
Article
Full-text available
Ammonia has been receiving increasing interest as a hydrogen carrier and carbon-free fuel to tackle the issue of greenhouse gas (GHG) emissions from transportation. In this study, an ammonia/diesel dual-fuel (ADDF) engine is experimentally and numerically investigated, with focus on its feasibility to reduce GHG emissions while achieving a diesel-like efficiency. A single-cylinder, heavy-duty diesel engine is used to investigate the effect of ammonia energy fraction and start of diesel injection (SODI) timing on the combustion performance and emissions of the ADDF engine. Results revealed that due to the low flame speed of ammonia, increasing the ammonia energy fraction decreased the thermal efficiency of the ADDF combustion mode compared to the diesel-only combustion mode. Increasing ammonia energy fraction from 0 to 40% reduced the nitrogen oxides (NOx) emissions by 58.8% at a given SODI due to the effect of the thermal DeNOx process. However, increasing the ammonia energy fraction at a given SODI increased the nitrous oxide (N2O) emissions, which offsets the benefit of lower intrinsic carbon dioxide (CO2) emissions of ADDF combustion and resulted in a higher GHG emission compared to diesel-only combustion. Advancing SODI helped reduce the N2O and overall GHG emissions while achieving a diesel-like thermal efficiency in ADDF combustion mode. The lowest GHG emissions of ADDF combustion achieved by advancing the SODI were 12% lower compared to those of diesel-only combustion. The thermal efficiency of ADDF combustion mode at the optimum point of GHG emissions (i.e., ITE = 37.85%) was slightly lower than that of the diesel-only combustion mode (i.e., ITE = 38.53%). This, however, comes with a benefit of 10% and 20% reduction in NOx and CO emissions, respectively. The unburned ammonia concentration is high (i.e., about 4445 ppm) in the exhaust flow. The reduction of ammonia emissions in the exhaust flow should be further investigated in the future.
Article
Full-text available
Methanol is a widely used engine fuel, blend component, and additive. However, no systematic auto-ignition data or laminar flame speed measurements are available for kinetic studies of the effect of methanol as a blending or additive component. In this work, both ignition delay times and laminar flame speeds of pure methanol, n-heptane and their blends at various blending ratios were measured at engine-relevant conditions. Results show that increasing methanol in a blend promotes reactivity at high temperatures and inhibits it at low temperatures, with the crossover temperature occurring at approximately 970–980 K with it being almost independent of pressure. The experimental data measured in this work, together with those in the literature are used to validate NUIGMech1.1, which predicts well the experimental ignition delay times and laminar flame speeds of the pure fuels and their blends over a wide range of conditions. Furthermore, kinetic analyses were conducted to reveal the effects of methanol addition on the oxidation pathways of n-heptane and the dominant reactions determining the fuel reactivities. It is found that competition for ȮH radicals between methanol and n-heptane plays an important role in the auto-ignition of the fuel blends at low temperatures. At high temperatures, methanol produces higher concentrations of HȮ2 radicals which produce two ȮH radicals either through the production of H2O2 and its subsequent decomposition or through direct reaction with Ḣ atoms. This promotes the high temperature reactivity of methanol/n-heptane mixtures for ignition delay times and laminar flame speeds, respectively.
Article
This study reports new ignition delay time (IDT) measurements of ethane (C2H6)/‘air’ mixtures with NOx (nitric oxide (NO), nitrogen dioxide (NO2), and nitrous oxide (N2O)) addition in the range 0 – 1000 ppm at stoichiometric fuel to air (φ) ratios, at compressed temperatures (TC) of 851 – 1390 K and at compressed pressures (pC) of 20 – 30 bar. In addition, new IDT measurements of three highly diluted C2H6/NO2 mixtures at φ = 0.5, TC = 805 – 1038 K, and pC = 20 – 30 bar are also studied. These new experimental data, together with data already available in the literature, are used to validate NUIGMech1.2 with an updated NOx sub-mechanism. Although the addition of 200 ppm of NO or NO2 to ethane shows a minimal promoting effect, the addition of 1000 ppm significantly promotes its reactivity. The similarity of the effect of the addition of both NO and NO2 addition is due to the fast conversion of NO into NO2 in the presence of molecular oxygen. However, the 1000 ppm NO doped ethane mixtures exhibit ∼20% shorter IDTs compared to the NO2 blended ones. The addition of 1000 ppm of N2O exhibits no effect on ethane oxidation at the conditions studied. The NUIGMech1.2 predictions can reproduce the sensitisation effect of NOx on ethane with good agreement over a wide range of pressure, temperature, equivalence ratio, and percentage dilution. Sensitivity and flux analyses of C2H6/NOx are performed to highlight the key reactions controlling ignition over the different temperature regimes studied. The analyses show that there is a competition between the reactions Ṙ + NO2 ↔ RȮ + NO and Ṙ + NO2 (+M) ↔ RNO2 (+M). This governs NOx sensitization on C2H6 ignition.
Article
Recent modeling studies of NH3 oxidation, which are motivated by the prospective role of ammonia as a zero-carbon fuel, have indicated significant discrepancies between existing literature mechanisms. In this study high level theoretical kinetics predictions have been obtained for the reaction of NH2 with HO2, which has previously been highlighted as an important reaction with high sensitivity and high uncertainty. The potential energy surface is explored with coupled cluster calculations including large basis sets and high-level corrections to yield high accuracy (∼0.2 kcal/mol) estimates of the stationary point energies. Variational transition state theory is used to predict the microcanonical rate constants, which are then incorporated in master equation treatments of the temperature and pressure dependent kinetics. For the radical-radical channels, the microcanonical rates are obtained from variable reaction coordinate transition state theory implementing directly evaluated multireference electronic energies. The analysis yields predictions for the total rate constant as well as the branching to the NH3 + O2, H2NO + OH, and HNO + H2O channels. Rate constants are also reported for the H2NO + OH reaction as they arise naturally from the analysis. The rate constant and branching fraction determined in this work for the NH2 + HO2 reaction deviate significantly from values used in most previous modeling studies. The fact that the main product channel is chain terminating, rather than propagating, has strong implications for modeling NH3 ignition and oxidation, in particular at intermediate temperatures and elevated pressure.
Article
Ammonia (NH3) has recently received much attention as a promising future fuel for mobility and power generation. The use of ammonia as a fueling vector can help curb global warming by cutting CO2 emissions because it is a carbon-free fuel and a hydrogen carrier with a high percentage of hydrogen atoms per unit volume. Liquid ammonia contains a higher volumetric density of hydrogen than liquid hydrogen. The low reactivity of ammonia, however, hinders its direct usage as a combustible fuel. One feasible way to boost the reactivity of ammonia is to target a dual-fuel system comprising of ammonia and a suitable combustion promoter. In this work, combustion properties of ammonia were investigated by blending it with various proportions of dimethyl ether (DME) using a rapid compression machine (RCM) and a constant volume spherical reactor (CVSR) over a wide range of experimental conditions. DME is a highly reactive fuel that may be produced in a sustainable carbon cycle with a net zero-carbon emission. Ignition delay times (IDTs) of NH3/DME blends were measured over a temperature (T) range of 649–950 K, pressures (P) of 20 and 40 bar, equivalence ratios (Φ) of 0.5 and 1 for a range of DME mole fractions (χDME) of 0.05–0.5 in the blends. In addition, the laminar burning velocities of NH3/DME blends were measured at P = 1, 3 and 5 bar, Φ = 0.8 to Φ = 1.3 and T = 300 K for χDME ranging 0.18 to 0.47. Our results suggest that DME is a good ignition promoter, resulting in a significant shortening of IDTs and an increase of flame speeds of NH3. A detailed chemical model has been developed and validated against the experimental data. Overall, our kinetic model offered reasonable predictive capabilities capturing the experimental trends over a wide range of conditions. In the worst-case scenario, our model underpredicted IDTs by a factor of ∼2.5 while overpredicting laminar flame speed by ∼20%.
Article
An experimental and kinetic modeling study of the influence of NOx (i.e. NO2, NO and N2O) addition on the ignition behavior of methane/‘air’ mixtures is performed. Ignition delay time measurements are taken in a rapid compression machine (RCM) and in a shock tube (ST) at temperatures and pressures ranging from 900–1500 K and 1.5–3.0 MPa, respectively for equivalence ratios of 0.5–2.0 in ‘air’. The conditions chosen are relevant to spark ignition and homogeneous charge compression ignition engine operating conditions where exhaust gas recirculation can potentially add NOx to the premixed charge. The RCM measurements show that the addition of 200 ppm NO2 to the stoichiometric CH4/oxidizer mixture results in a factor of three increase in reactivity compared to the baseline case without NOx for temperatures in the range 600–1000 K. However, adding up to 1000 ppm N2O does not show any appreciable effect on the measurements. The promoting effect of NO2 was found to increase with temperature in the range 950–1150 K, while the sensitization effect decreases at higher pressures. The experimental results measured are simulated using NUIGMech1.2 comprising an updated NOx sub-chemistry in this work. A kinetic analysis indicates that the competition between the reactions ĊH3 + NO2 ↔ CH3Ȯ + NO and ĊH3 + NO2 (+M) ↔ CH3NO2 (+M), the former being a propagation reaction and the latter being a termination reaction governs NOx sensitization on CH4 ignition. Recent calculations by Matsugi and Shiina (A. Matsugi, H. Shiina, J. Phys. Chem. A. 121 (2017) 4218–4224) for the nitromethane formation reaction CH3 + NO2 (+M) ↔ CH3NO2 (+M), together with the recently calculated rate constants for HONO/HNO2 reactions significantly improve ignition delay time predictions in the temperature range 600–1000 K. Furthermore, the experiments with NO addition reveal a non-monotonous sensitization impact on CH4 ignition at lower temperatures with NO initially acting as an inhibitor at low NO concentrations and then as a promoter as NO concentrations increase in the mixture. This non-monotonous trend is attributed to the role of the chain-termination reaction ĊH3 + NO2 (+M) ↔ CH3NO2 (+M) and the impact of NO on the transition to the chain-branching steps CH2O + HȮ2 ↔ HĊO + H2O2, H2O2 (+M) ↔ ȮH + ȮH (+M), HĊO ↔ CO + Ḣ followed by CO + O2 ↔ CO2 + Ö and Ḣ + O2 ↔ Ö + ȮH. NUIGMech1.2 is systematically validated against the new ignition delay measurements taken here together with species measurements and high temperature ignition delay time data available in the literature for CH4/oxidizer mixtures diluted with NO2/N2O/NO and is observed to accurately capture the sensitization trends.
Article
The kinetics of the reactions H2NO + O2(³Σg−) → HNO(X˜1A′) + HO2 and NH2 + HO2 → NH3 + O2(³Σg−), which are, respectively, very sensitive chain-propagation and chain-termination reactions in ammonia kinetic models, have been revisited by means of high-level electronic structure and variational transition state theory calculations with the goal of improving former predictions and the performance of ammonia kinetic models. In addition, the rate constants of the reactions H2NO + O2(³Σg−) → HNO(a˜3A″) + HO2, NH2 + HO2 → H2NO + OH, and NH2 + HO2 → NH3 + O2(¹Δg), which take place on excited-state potential energy surfaces and/or yield the electronically excited species HNO(a˜3A″) and O2(¹Δg), have been also calculated for the first time in order to assess their importance in ammonia oxidation. We observed that spin contamination and multi-reference character are pronounced in many of the investigated reactions, and these features were handled by performing post-CCSD(T) electronic structure calculations with the W3X-L composite method as well as restricted open shell coupled cluster calculations. Branching ratios were also analyzed, and indicate that the contribution of the electronically excited species HNO(a˜3A″) and O2(¹Δg) are of little importance even at very high temperatures; however, we do not preclude an effect of those species at certain conditions that contribute to their yield. The calculated rate constants were implemented in two recent kinetic models to perform jet stirred reactor, rapid compression machine, and flow reactor simulations, concluding that the model predictions are very sensitive to the reactions H2NO + O2(³Σg−) → HNO(X˜1A′) + HO2 and NH2 + HO2 → NH3 + O2(³Σg−) .
Article
Ammonia (NH3) is a promising carbon-free fuel and a hydrogen carrier. In recent years, there has been a large number of experimental and numerical studies to understand the chemical kinetics of NH3 in moderate to extremely complex systems. This study focused on understanding the chemical kinetics of NH3 in a simple system (pyrolysis). Shock-tube experiments were performed to monitor the NH3 time history profiles during pyrolysis, with and without the presence of H2, using a new laser absorption diagnostic near 10.4 µm. The pyrolysis experiments were conducted for mixtures of ∼ 0.5% NH3/Ar and ∼ 0.42% NH3/2% H2/Ar behind reflected shock waves, near atmospheric pressure, and over a temperature range of 2100–3000 K. Using the data from the present study as a guide, along with NH3 pyrolysis data from the literature, a detailed chemical kinetics mechanism for NH3 pyrolysis is proposed herein. This mechanism was assembled using available reaction rate constants from the literature. The mechanism showed excellent agreement with the experimental results, as well as with the literature data. Additionally, an assessment of 15 detailed NH3 chemical kinetics mechanisms on their capabilities of predicting the new pyrolysis experiments was performed. The assessment showed that these literature mechanisms yield significantly different predictions, with only one model producing acceptable results for the majority of the NH3 pyrolysis experiments. The effect of pyrolysis reactions on the prediction of oxidation data was investigated by updating the pyrolysis sub-mechanism of selected literature models with the reactions of the present pyrolysis model. The prediction of the updated literature models significantly improved for ignition delay time and flame speed literature data, indicating the importance of pyrolysis reactions for high-temperature oxidation chemistry.
Article
Ammonia (NH3) is a promising carbon-free alternative fuel, and blending NH3 with other more active fuels can enhance the low burning rate of pure NH3 flames, which makes it more compatible with existing combustors. There are laminar burning velocity data of NH3+H2, NH3+CO/syngas, and NH3+CH4 flames in literature, however, blending NH3 with the oxygenated alternative fuel alcohols have not yet been checked. In the present work, the laminar burning velocities of NH3+CH3OH/air and NH3+C2H5OH/air flames were measured using the heat flux method at 1 atm with varied equivalence ratios and mixing ratios. Measurements were carried out at 298 K and elevated temperatures until 448 K unburnt temperatures, in order to check the data consistency via the temperature coefficient α in SLSL0=(TuTu0)α. A new updated kinetic mechanism, based on our previous mechanism named CEU-NH3, was proposed and validated based on the present and published experimental data for laminar burning velocity, ignition delay times and species profiles. This mechanism contains 91 species and 444 reactions, which is small in size and thus advantageous to be used in the modeling of complex combustion fields. By numerical analyses using the CEU-NH3 and the detailed Konnov mechanisms, the interactions between C- and N- containing species are found to be insignificant for the laminar burning velocities of the various NH3 blending mixtures.
Article
Autoignition delay times of ammonia/dimethyl ether (NH3/DME) mixtures were measured in a rapid compression machine with DME fractions of 0, 2 and 5 and 100% in the fuel. The measurements were performed at equivalence ratios φ=0.5, 1.0 and 2.0 and pressures in the range 10–70 bar; depending on the fuel composition, the temperatures after compression varied from 610 K to 1180 K. Admixture of DME is seen to have a dramatic effect on the ignition delay time, effectively shifting the curves of ignition delay vs. temperature to lower temperatures, up to ~250 K compared to pure ammonia. Two-stage ignition is observed at φ=1.0 and 2.0 with 2% and 5% DME in the fuel, despite the pressure being higher than that at which pure DME shows two-stage ignition. At φ=0.5, a reproducible pre-ignition pressure rise is observed for both DME fractions, which is not observed in the pure fuel components. A novel NH3/DME mechanism was developed, including modifications in the NH3 subset and addition of the NH2+CH3OCH3 reaction, with rate coefficients calculated from ab initio theory. Simulations faithfully reproduce the observed pre-ignition pressure rise. While the mechanism also exhibits two-stage ignition for NH3/DME mixtures, both qualitative and quantitative improvement is recommended. The overall ignition delay times for ammonia/DME mixtures are predicted well, generally being within 50% of the experimental values, although reduced performance is observed for pure ammonia at φ=2.0. Simulating the ignition process, we observe that the DME is oxidized much more rapidly than ammonia. Analysis of the mechanism indicates that this ‘early DME oxidation’ generates reactive species that initiate the oxidation of ammonia, which in turn begins heat release that raises the temperature and accelerates the oxidation process towards ignition. The reaction path analysis shows that the low-temperature chain-branching reactions of DME are important in the early oxidation of the fuel, while the sensitivity analysis indicates that several reactions in the oxidation of DME, including cross reactions between DME and NH3 species presented here, are critical to ignition, even at fractions of 2% DME in the fuel.