Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Fuel 341 (2023) 127676

Contents lists available at ScienceDirect

Fuel
journal homepage: www.elsevier.com/locate/fuel

Full Length Article

Combustion chemistry of ammonia/C1 fuels: A comprehensive kinetic


modeling study
Xiaoyuan Zhang *, Kiran K. Yalamanchi , S. Mani Sarathy
Clean Combustion Research Center, King Abdullah University of Science and Technology, Thuwal 23955-6900, Saudi Arabia

A R T I C L E I N F O A B S T R A C T

Keywords: The application of ammonia (NH3) as a fuel could contribute to mitigating global warming and achieving carbon
Ammonia neutrality by mid-century. To enhance the reactivity of NH3 combustion, cofiring NH3 with hydrogen or C1 fuels
Syngas like syngas, methanol and methane is proposed. In previous work, we studied the combustion chemistry of NH3/
Methanol
H2 mixtures using a comprehensive kinetic model. In this work, we expanded our comprehensive kinetic model
Methane
to cover NH3/syngas, NH3/methanol and NH3/methane mixtures. Rate constants of critical cross reactions be­
Kinetic model
tween C- and N-containing species were evaluated based on previous experimental and theoretical studies for
selection and incorporation in the present model. Experimental data for both NOx/C1 and NH3/C1 fuels were
selected from literature to test the present model, including speciation data measured in shock tubes, flow re­
actors, jet-stirred reactors, burner-stabilized flames, and the global parameters like ignition delay times and
laminar flame speeds. The kinetic model validation conditions cover temperatures of 473–2000 K, pressures of
0.04–100 atm, and equivalence ratios of 0.04–116. In general, the present model adequately captures the
selected experimental data. The present model can not only be used to predict the combustion of NH3/C1 fuel
mixtures, but is also capable to predict the mutual interaction of NOx/C1 fuels. It is found that the rate constants
between C-containing species and H2-related radicals are generally faster than (or close to) those between C- and
N-containing species. At high temperatures, H2-related radicals (i.e. H, O, OH) have higher concentrations than
reactive N-related radicals (NH, NH2, NO2). Therefore, under these conditions, C-containing species are more
likely to react with H2-related radicals rather than N-related ones, making the C/N cross reactions less prominent.
However, under low- and intermediate-temperature conditions, the concentrations of H and O radicals become
lower, while those of NH2 and NO2 become higher, making the cross reactions between C- and N-containing
species possible to compete with C/H cross reactions. The present model can be used as the base chemistry model
for NH3/larger hydrocarbon fuels.

1. Introduction reactivity than NH3 and blending them can enrich the radical pools of
NH3 combustion [1]. In addition, NO can react with small hydrocarbon
Ammonia (NH3) is an attractive H2 carrier and carbon-free fuel that radicals (e.g., CH2, CH3, HCCO) to eventually form N2 via reburning
has received significant attention in recent years. It can either be con­ chemistry [2], resulting in lower NOx emissions. A reliable kinetic model
verted to H2 for energy supply, or used directly as a fuel, both of which for NH3 and C1 fuels is therefore necessary to understand fuel blending
could contribute to mitigating global warming and achieving carbon chemistry and to predict the operation conditions where the lowest NOx
neutrality by mid-century. However, its utilization as direct fuel has two emissions are reached.
main concerns: low reactivity and high NOx emissions. Regarding to Kinetic models for NH3/C1 fuels consist of pure NH3 and C1 fuel
these two issues, cofiring NH3 with C-containing fuels is a possible so­ subsets and the interaction mechanisms between C- and N-containing
lution. Particularly, small C-containing fuels like syngas, methanol species. Although C1 models are considered ubiquitous, well-
(CH3OH) and methane (CH4) are cleaner fuels than coal and heavy oils. established, and reliable, recent progress in non-thermal effects [3]
Blending these C1 fuels with NH3 can reduce emissions. From the per­ like prompt dissociation, are not fully adopted in widely used core C1
spectives of combustion chemistry, these C1 fuels have much higher reaction mechanisms [4]. Compared with the C1 subset, the pure NH3

* Corresponding author.
E-mail address: xiaoyuan.zhang@kaust.edu.sa (X. Zhang).

https://doi.org/10.1016/j.fuel.2023.127676
Received 7 December 2022; Received in revised form 31 January 2023; Accepted 1 February 2023
Available online 10 February 2023
0016-2361/© 2023 Elsevier Ltd. All rights reserved.
X. Zhang et al. Fuel 341 (2023) 127676

submechanism contains larger uncertainties. In our recent modeling HO2 + HO2 = OH + OH + O2 (R2)
study on NH3/H2, the performance of different comprehensive models
displayed large discrepancies [5]. Earlier combustion chemistry studies To better predict CH3OH oxidation under high-pressure and low-
on the interaction mechanism between C- and N-containing species were temperature conditions, the rate constant of R3 was updated based on
mainly motivated by the reburning process. In the modeling studies of the estimation from Burke et al. [34]. To better predict the flame
Miller and Glarborg [6,7], cross reactions between N- and C-containing propagation and ignition of CH4, the rate constants of R4 and R5 were
species (up to C2) were considered. These studies are important building updated from Troe et al. [35] and Srinivasan et al. [36], respectively.
blocks for the further development on C/N interaction chemistry CH3OH + HO2 = CH2OH + H2O2 (R3)
[2,8–14]. Among these studies, the most recent and comprehensive
modeling study is from Glarborg et al. [2], which was developed based CH3 + H(+M) = CH4(+M) (R4)
on their previous studies over the past 30 years and widely adopted by
CH4 + O2 = CH3 + HO2 (R5)
recent literature modeling work on NH3/hydrocarbon fuels [15–17].
Nevertheless, recent theoretical calculations [1,18–20] on the cross re­ In addition, the reactions in Chebyshev format were replaced by Troe
actions between C- and N-containing species provide chances to further format, and rate constants with a negative pre-exponential factor were
improve the C/N interaction mechanism. On the other hand, most pre­ refitted and replaced by positive ones. The re-fitting uncertainties are
vious chemical kinetic models for NH3/syngas [21,22], NH3/CH3OH within 10 % under combustion related conditions. This modification
[1,23] and NH3/CH4 [15,24–26] focused on specific combustion con­ ensures facile application of the present model into various simulation
ditions and limited experimental targets for model validation. software. Since the rate constants of R1-R5 were updated, selected
Comprehensive kinetic models for NH3/C1 fuels, which are validated validation targets for pure fuels were used to re-examine the model
against experimental data covering chemically reacting flows at various performance, as found in subsequent sections of this paper.
temperatures, pressures, and equivalence ratios in various reactors, are For the NH3 subset, the rate constants of isomerization and decom­
scarce. position reactions of HONO radical (i.e. R6-R9) were updated based on
In this work, a comprehensive kinetic model for NH3/syngas/ recent theoretical calculations by Chen et al [37]. They identified a well-
CH3OH/CH4 is proposed based on our ongoing modeling studies for pure skipping product channel, i.e. R8, which was not found in previous
NH3 [5] and small hydrocarbon fuels [27–32]. Rate constants of critical theoretical work by Rasmussen et al. [10], and which was missing in our
cross reactions between C- and N-containing species were evaluated previous NH3 model [5]. The rate constants of R8 are comparable with
based on previous experimental and theoretical studies. The proposed those of R7 at 1 atm, and 3–5 times lower than R7 at 100 atm.
model is validated against previous experimental data for both NH3/C1
fuels and NOx/C1 fuels, covering wide temperature, pressure and NO2 + H = NO + OH (R6)
equivalence ratio conditions. The role of C/N cross reactions under OH + NO = HONO (R7)
different conditions is discussed based on the present model.
OH + NO = HNO2 (R8)
2. Model development
HNO2 = HONO (R9)
2.1. Pure ammonia and pure C1 fuels Bimolecular reactions of HONO isomers with H atom (i.e. R10-R15)
were updated based on theoretical calculations in [19]. According to
The starting kinetic models for NH3 and C1 fuels (i.e. syngas, CH3OH their calculation results, at atmospheric pressure, R10, R11 and R12 are
and CH4) were taken from our previous modeling studies [5,27–32]. The competitive for the consumption of HONO at high-temperatures, while
C1 fuel models were developed hierarchically and adopted the same R13 is dominant amongst R13-R15 for the consumption of HNO2 over
hydrogen mechanism as the NH3 model; therefore, they were merged the whole temperature range. These modifications on HONO/HNO2
together. Our recent NH3 model [5] has been tested against compre­ have negligible influences on the predictions of NH3 oxidation. As seen
hensive experimental data for pure NH3, NH3/H2, H2/NO, H2/N2O, from Figs. S1-S6 in the Supplementary material, the present model well
NH3/NO, NH3/NO2 mixtures; it can reproduce the validation targets predicts speciation data in H2/NO oxidation.
over a wide range of pyrolysis and oxidation conditions. The original C1
fuel models have incorporated recent progress in non-thermal effects [3] HONO + H = NO2 + H2 (R10)
like prompt dissociation and termolecular reactions. The CH3OH subset
HONO + H = HNO + OH (R11)
was taken from our previous modeling work on formaldehyde (CH2O)
and CH3OH [29], which was comprehensively tested against both HONO + H = NO + H2O (R12)
speciation data and global combustion parameters under wide com­
bustion conditions. The CH4 subset includes both C1 and C2 chemistry. HNO2 + H = NO2 + H2 (R13)
The high-temperature oxidation mechanism of CH4 was taken from our HNO2 + H = HNO + OH (R14)
previous modeling studies on CO/CH4 [30] and ethylene (C2H4) [27], in
which experimental targets obtained under intermediate- and high- HNO2 + H = NO + H2O (R15)
temperature conditions were selected to test the model. The low-
temperature oxidation mechanism of CH4 was adopted from our
modeling studies on acetaldehyde (CH3CHO) [28] and propanal
(C2H5CHO) [32], in which the low-temperature chemistry of CH3 and 2.2. Interactions between N- and C1-containing species
C2H5 was visited.
Recent theoretical calculation study from Klippenstein et al. [33] Important radicals/intermediates in ammonia oxidation, such as
revisited the reaction of HO2 + HO2, and found that besides the triplet NH2, NO and NO2, can react with C1 fuels and their related radicals.
product channel R1, the singlet product channel to HO3 + OH plays an Reburning chemistry between C1 species and NO is the basis for un­
important role in leading to the formation of O2 and two OH radicals, i.e. derstanding NH3/C1 chemistry, since NO is one of the most important
R2. Klippenstein et al.’s calculation results for R1 and R2 were adopted intermediate/product in ammonia oxidation. NO has a relatively high
in the present model. concentration and has a radical character, indicating that the reactions
with NO could be competitive. NO is easily oxidized to NO2 via R16,
HO2 + HO2 = H2O2 + O2 (R1)
converting HO2 to OH in the process.

2
X. Zhang et al. Fuel 341 (2023) 127676

NO + HO2 = NO2 + OH (R16) CH2O + NO2 = HCO + cis/trans-HONO (R24)

NO2 is an unusual intermediate in NH3 oxidation, which exhibits a


high affinity, since every atom in the molecule has a radical character. CH2O + NO2 = HCO + HNO2 (R25)
The reactions related to NO2 play an important role in NH3/hydrocarbon CH2O + NH2 = HCO + NH3 (R26)
oxidation under low and intermediate temperatures [1,15,26]. Besides
nitrogen oxides, NHi radicals can also react with C1 species. These re­ Besides R24 and R25, Wu et al. [18] also investigated the kinetics of
actions could play an important role in determining global combustion CH3OH + NO2 theoretically. Their computed global rate constant lies
parameters at low and intermediate temperatures [1] and in predicting between previous measurements [45,46]. Similar to previous theoretical
speciation profiles [15]. In this work, the original interaction mecha­ calculations by Xiao et al. [47], they found the branching channels to
nism between N- and C-containing species, including kinetic, thermo­ CH2OH formation are more important than CH3O formation. However,
dynamic and transport data, was adopted from Glarborg et al. [2]. branching ratios amongst the product channels display large discrep­
Important cross reactions between C- and N-containing species were ancies between Wu et al. [18] and Xiao et al. [47]. According to Wu et al.
revisited in this work, with full consideration of more recent studies. [18], the channels to CH2OH + cis-HONO and R28 are dominant at 300
to 2000 K. In contrast, the results from Xiao et al. [47] show that the
2.2.1. Interaction mechanism between N-containing species and CO subset product channels to CH2OH + cis-HONO and CH2OH + trans-HONO are
Since CO and CO2 are relatively stable molecules, only a few radicals dominant. In this model, we adopted the more recent theoretical cal­
are reactive enough to attack them. Recently, Kroupnov et al. [38] culations by Wu et al. [18] to describe the kinetics of CH3OH + NO2,
theoretically studied R17 and reported both forward and reverse rate since they adopted a more advanced theoretical calculation method,
constants. They also compared their results with literature data. It was including proper treatments of quantum tunneling effects and partition
found that their computed rate constants match well with those functions.
measured at intermediate and high temperatures [39,40], while they are
CH3OH + NO2 = CH2OH + cis/trans-HONO (R27)
10 times higher than earlier measurements at 500 K [41]. Since we
focused on combustion related conditions, where the rate parameter CH3OH + NO2 = CH2OH + HNO2 (R28)
accuracy at intermediate and high temperatures are more important
than 500 K, the computed result from [38] was adopted in the present CH3OH + NO2 = CH3O + cis/trans-HONO (R29)
model. CH3OH + NO2 = CH3O + HNO2 (R30)
CO + NO2 = CO2 + NO (R17) The reaction between CH3OH and NH2 was theoretically investi­
Poskrebyshev et al. [42] theoretically investigated the reaction be­ gated in a recent modeling work on NH3/CH3OH [1]. According to this
tween NH2 and CO, as well as the decomposition of NH2CO to HNCO and study, R31 and R32 are the main product channels. R33 has a much
H. The computed rate constant for R18 at 300 K and 1 atm is 1.3×106 higher energy barrier and is not important. Interestingly, according to Li
cm3/mol/s, which is several orders of magnitude lower than common et al. [1], the rate constant of R32 is roughly-two times faster than that of
addition reactions. Therefore, it was not incorporated in the present R31 at 1000 K, which shows an opposite trend to the H-abstraction re­
model. The addition-elimination reaction of NH2 + CO can lead to the actions of CH3OH by other radicals, such as H, O, OH and HO2 [29].
formation of HNCO and H (i.e. R19), which transfers NH2 to more Based on their PES calculations, the energy barrier of R32 is only 1 kJ/
reactive H atom and varies the radical pool. Therefore, it was incorpo­ mol higher than that of R31. However, R32 has a looser transition state
rated in this model with a rate constant taken from Baulch et al. [43], in than R31, making k32 faster than k31. More experimental data are
which the reverse rate constant of R19 was evaluated. CO2 has two required to confirm these theoretical results [1]. Due to the limited
double bonds and is more stable than CO. Its reactions with N-containing studies on the kinetics of CH3OH + NH2, the theoretical calculations
radicals like amines were reviewed in previous work [25]. Similar to CO, from Li et al. [1] were adopted for R31-R33.
the reactions between N-containing radicals and CO2 are too slow to be CH3OH + NH2 = CH2OH + NH3 (R31)
of importance under combustion related conditions.
CH3OH + NH2 = CH3O + NH3 (R32)
NH2 + CO = NH2CO (R18)
CH3OH + NH2 = NH2OH + CH3 (R33)
NH2 + CO = HNCO + H (R19)
CH2OH is a dominant intermediate radical in CH3OH oxidation [29].
The interaction mechanism between HCO and NOx mainly includes It can combine with NOx species to proceed R34-R36, and react with NHi
R20-R23. Details on the kinetic parameter evaluation can be found in via R37-R40. Details on the kinetic parameter review and selection are
the Supplementary material. available in the Supplementary material. Note that although the inter­
HCO + NO = HNO + CO (R20) action reactions between N-containing species and CH2OH were incor­
porated in the present model, they are not competitive with R41 due to
HCO + NO2 = NO + CO2 + H (R21) much higher concentrations of O2 compared to N-containing species and
comparable rate constants between R41 and R34-R40.
HCO + NO2 = HONO + CO (R22)
CH2OH + NO = HNCO + H2O (R34)
HCO + NO2 = HO + NO + CO (R23)
CH2OH + NO2 = CH2O + HONO (R35)

CH2OH + NO2 = CH2O + HNO2 (R36)


2.2.2. Interaction mechanism between N-containing species and CH3OH
subset CH2OH + NH2 = CH2O + NH3 (R37)
Since CH2O is the most important intermediate in CH3OH oxidation, CH2OH + NH2 = NH + CH3OH (R38)
the reactions between CH2O and N-containing species, i.e. R24-R26,
were revisited. Their rate constants were taken from Wu et al. [18] CH2OH + NH2 = HCOH + NH3 (R39)
and Li et al. [44]. Details on the rate constant evaluation can be found in
CH2OH + NH2 = CH2NH2 + OH (R40)
the Supplementary material.

3
X. Zhang et al. Fuel 341 (2023) 127676

Fig. 1. Demissy et al. [50] evaluated the rate constant of R45 experi­
mentally by flash photolysis using laser resonance absorption detection
of NH2. The temperature range covers 300–520 K. Hack et al. [51]
studied R45 in an isothermal discharge flow system over the tempera­
ture range of 743–1020 K, extending previous work [50] to intermediate
temperatures. At high temperatures, the measured rate constants by
Song et al. [52] are 3 times higher than those measured by Hennig et al.
[53] over 1600–2100 K. Both experiments were conducted in shock
tubes. In the experimental study of Hennig et al. [53], NH2 was formed
by the thermal decomposition of hydrazine and its concentration was
measured by laser absorption. In contrast, Song et al. [52] used
methylamine to generate NH2, and the NH2 decay was measured using a
frequency modulation (FM) absorption technique, which can achieve
much lower detection limit compared with direct laser absorption. Be­
sides these measurements, Yu et al. [54] theoretically studied R45 and
obtained both the forward and reverse rate constants over 300–2100 K.
Their computed forward rate constant is in reasonable agreement with
previous measurements from [51,53] over intermediate and high tem­
peratures, while overpredicting the rate constant measured at 300 K
from [50] by a factor of 10. Mebel et al. [55,56] also computed the rate
Fig. 1. Comparison of the rate constant of CH4 + NH2 = CH3 + NH3 (R45). constant of R45. In contrast, their calculation results agree well with the
Symbols [50–53] and lines [52,55–58] represent measured and theoretical measured results at low temperatures [50], while overpredicting the
calculation values. measured rate constants at intermediate [51] and high temperatures
[52,53]. More recent theoretical calculations on R45 were performed by
Siddique et al. [57] at the CBS-QB3 level. Their computed results agree
CH2OH + O2 = CH2O + HO2 (R41) well with the measured results of [50] at low temperatures and those of
Song et al. [52] at high temperatures, while overpredicting the mea­
For the methoxy radical (CH3O), its bimolecular reactions are less surements of Hack et al. [51] by a factor of 4 at intermediate tempera­
competitive than unimolecular decomposition reaction, i.e. R42. tures. Song et al. [52] not only performed an experimental study, but
Therefore, the interaction mechanism between N-containing species and also performed theoretical calculations. They computed the rate con­
CH3O are not important. Nevertheless, we still incorporated the inter­ stants of R45 over 300–2100 K, which well predict the measured rate
action reactions between N-containing species and CH3O in the present constants [50–52] over 300–2084 K. In the present model, we adopted
model for completeness. The reaction pathways and rate constants were the rate constant from the theoretical calculations of Song et al. [52],
adopted from previous modeling studies [2,48]. albeit large uncertainties remain for the rate constant of R45 at high
temperatures.
CH3O (+M) = CH2O + H (+M) (R42)
For the reactions between CH3 and NHi, multiple reaction channels
are possible, including H-atom abstraction, recombination, and
addition-elimination reactions, i.e. R46 and R47. Dean and Bozzelli [48]
2.2.3. Interaction mechanism between N-containing species and CH4 subset performed theoretical calculations to identify possible reaction chan­
H-atom abstraction reactions from CH4 by NO2 can lead to three nels, and obtained rate constants based on QRRK theory. Their results
product channels, i.e. R43-R44. Review on the rate constants of these were adopted in the present model.
reactions before 2008 can be found in [11]. The sub-mechanism of CH3NH2 was taken from a recent kinetic
modeling work from Glarborg et al. [20], where the rate constants of
CH4 + NO2 = CH3 + cis/trans-HONO (R43) critical reactions were computed theoretically.
CH4 + NO2 = CH3 + HNO2 (R44) CH3 + NH = CH2NH + H (R46)
According to a recent study at a high level of theory [49], the rate CH3 + NH2 = CH3NH2 (R47)
constant of CH3 + cis-HONO is the largest, while R44 is the smallest. The
Glarborg 2018 model [2] adopted the computed rate constants for R43- The reaction between CH3 and NO2 can proceed via a dispropor­
R44 from [49] without distinguishing cis-HONO and trans-HONO. tionation reaction (i.e. R48) or combination reaction (i.e. R49). A study
Duplicate reactions were defined in the model for cis-HONO and trans- by Glarborg et al. [59] evaluated k48 at 1180 K by adjusting this
HONO channels. To simplify the mechanism, recently, Fuller et al. [19] parameter in the mechanism to match measured NO2 profiles. In com­
assumed that cis-HONO and trans-HONO were lumped into a single bination with low temperature data, they suggested the k48 decreases
species. Transition state theory (TST) calculations for the direct H- slightly with temperature. Wollenhaupt et al. [60] measured the rate
abstraction performed in [49] were redone with a single HONO isomer constant of R48 using LIF detection of CH3O over 233–356 K.
that systematically treats the cis-/trans- conversion as a hindered inter­ Temperature-independent rate constants over this range were reported,
nal rotation in both the reactant and transition state. Their computed which agree (within 16 %) with those evaluated by Glarborg et al. [59].
overall rate constant for CH4 + NO2 is within a factor of two compared At high temperatures, the measured k48 by Srinivasan et al. [61] is 50 %
with that obtained from [49]. In the present model, the rate constants for higher than that from Glarborg et al. [59]. In the present model, the
R43 and R44 were taken from [49]. value from Glarborg et al. [59] was adopted. For R49, recent modeling
The H-atom abstraction reaction from CH4 by NH2 (R45) is another work from Sahu et al. [62] evaluated its rate constant at various tem­
chain-propagation reaction in NH3/CH4 oxidation. peratures and pressures. They recommended to adopt the theoretical
results from [63], which better match measurements [63,64] at various
CH4 + NH2 = CH3 + NH3 (R45) pressures compared to values from [59] and [65]. Therefore, Sahu
The rate constants for R45 from different sources are compared in et al.’s recommendations, i.e. theoretical results from [63], were adop­
ted in this model. Studies on the reactions related to nitromethane

4
X. Zhang et al. Fuel 341 (2023) 127676

both NOx/C1 and NH3/C1 fuel mixtures were selected for validation. The
validation targets are summarized in Table 1, including speciation data
measured in shock tubes, flow reactors, jet-stirred reactors (JSRs),
burner-stabilized premixed flames and the global parameters like
laminar flame speeds and ignition delay times. The temperature, pres­
sure and equivalence ratio conditions cover 473–2000 K, 0.04–100 atm
and 0.04–116, respectively. Two criteria are followed for choosing
experimental validation data: 1) the data cover as wide combustion
conditions as possible; and 2) the experiments adopt state-of-art
methods with quantified experimental uncertainties. Simulations were
performed with Chemkin-Pro software [77]. We followed the same
simulation methods used in our previous study on NH3/H2 [5], and
details can be found in [5]. Before performing simulations, the collision
limit compliance was checked based on the online tool developed in our
previous work [78] over 500–2500 K. The reactions that exceed collision
limits are mainly C2 related reactions at 500 K or above 2000 K, which is
beyond the normal combustion relevant conditions. These reactions are
not important in the present NH3/C1 systems, and will be evaluated in
our future modeling study on NH3/C2 mixtures. The model files,
Fig. 2. Comparison of the rate constant of CH3O2 + NO = CH3O + NO2 (R52). including kinetic, thermodynamic and transport data, are available in
Symbols and lines represent measured [68–72] and theoretical calculation/ the Supplementary material. In this section, comparisons between the
evaluation values [73–76], respectively.
experimental targets and the predicted results by the present model are
presented. In addition, the simulated results by Glarborg 2018 model [2]
chemistry are limited and important kinetic parameters available in the are also included for comparison, since this is the most recent and
literature are decades old. Further work is required in this area to comprehensive NH3 co-firing model, including NH3-syngas, NH3-
improve the understanding of nitromethane pyrolysis and combustion. CH3OH and NH3-CH4 subsets, and this model has served as the base
model for many other recent models, especially for the C/N interaction
CH3 + NO2 = CH3O + NO (R48)
mechanism.
CH3 + NO2 = CH3NO2 (R49)
3.1. Global parameters
For the kinetics of CH3 + NO, Miller et al. [66] computed the global
rate constant of CH3 + NO over 1000–2500 K, which is close to previous
3.1.1. Laminar flame speeds
shock tube measurements [67]. They concluded tentatively that the
Laminar flame speeds are helpful to examine the high-temperature
radical channel R51 should be dominant, which is in disagreement with
model, especially reactions involving H atoms. Experimental measure­
the measurements by Henning [67]. Similar to the Glarborg 2018 model
ments for laminar flame speeds of NH3/C1 fuels are available in the
[2], we assumed an equal contribution for R50 and R51 in the present
literature [21,23,82,83,101], covering wide unburned temperature,
model, and the global rate constant from Miller et al. [66] was adopted.
pressure and equivalence ratio conditions. Under some conditions, more
More studies on the branching ratio are required in the future.
than one set of experimental data are available [82,101,102], which is
CH3 + NO = HCN + H2O (R50) helpful to check the reliability of the experimental data. Fig. 3 presents
the comparison between the measured and predicted laminar flame
CH3 + NO = H2CN + OH (R51) speeds of stoichiometric NH3/syngas/air, NH3/CH3OH/air and NH3/
CH4/air at unburned gas temperature of 298 K. The experimental data
A number of experimental studies on the kinetics of CH3O2 + NO are
show that with increasing content of C1 fuels, the laminar flame speeds
available at low temperatures (200–400 K) and low pressures [68–72].
are increased. The enhancement effects of syngas are the strongest,
According to theoretical calculations [73], R52 is the dominant reaction
followed by CH3OH and CH4. This trend is well captured by the present
channel at 300–1500 K and 0.01–100 atm.
and the Glarborg 2018 model [2]. However, the present model can
CH3O2 + NO = CH3O + NO2 (R52) better predict the laminar flame speeds of NH3/C1 fuels under conditions
with low C1 fuel concentration. Under these conditions, the pure NH3
Atkinson et al. [74,75] reviewed previous studies and recommended mechanism becomes more important. The present model adopted the
k52 over ~200–400 K, as seen from Fig. 2. In the two review studies, the subset of NH3 from our previous model [5], which has a better perfor­
rate constants of R52 at 300 K are the same, while the temperature- mance than the Glarborg 2018 model [2] in predicting laminar flame
dependencies are different. The measurements from Villalta et al. [69] speeds for pure NH3 and NH3/H2 mixtures [5].
generally lie between those evaluated by Atkinson et al. [74,75]. Fig. 4 presents the measured and predicted laminar flame speeds
Measured results for k52 at temperatures higher than 430 K are not across wide equivalence ratio conditions and unburned gas temperature
available. Nguyen et al. [73] reported computed rate constants of R52 at of 298 K. For the NH3/syngas/air and NH3/CH4/air cases, as seen from
300–1500 K and 0.01–100 atm. Interestingly, k52 is nearly temperature- Fig. 4a and c, the present model can reasonably predict the experimental
independent at 300–430 K, which is inconsistent with previous negative- data under various unburned pressures, while the Glarborg 2018 model
temperature dependent trends observed [69,70,74,75]. At higher tem­ [2] can only reasonably predict NH3/CH4/air mixtures. For the pre­
peratures, the computed k52 has an obvious positive-temperature dictions of NH3/CH3OH/air mixtures, the present model better predicts
dependent trend. More experimental studies under higher tempera­ the pure CH3OH/air case than the Glarborg 2018 model [2]. Both the
tures are desired to confirm their calculations. In this work, we adopted present and Glarborg 2018 models [2] can reasonably predict cases with
the measurements from Villalta et al. [69] for R52. CH3OH contents of 60–80 %. With lower CH3OH contents, the present
and Glarborg 2018 models underpredict and overpredict the laminar
3. Model validation flame speeds, respectively. Additional validations against laminar flame
speeds of NH3/CO/air, NH3/syngas/air, NH3/CH3OH/air, NO/CH4,
To test the reliability of the present model, experimental data for NO/CH4/O2/N2, NO/CH4/O2/Ar, NH3/CH4/air can be found in Figs S7-

5
X. Zhang et al. Fuel 341 (2023) 127676

Table 1
A list of validation data for the present model.
Fuel Method Mixture T (K) P (atm) ϕ Refs

H2-NO JSR H2/NO/O2/N2 700–1150 1–10 0.1–1.0 [79]


Flow reactor H2/NO/O2/N2 802 10 0.25 [80]

CO-NO Flow reactor CO/H2O/O2/NO/N2 850–1400 1 1.24–1.68 [81]

CO-NH3 Flame speed CO/NH3/air 298 1 0.7–1.7 [82]


Flame speed CO/NH3/air 298 1–5 0.7–1.5 [83]
Flow reactor CO/NH3/H2O/O2/NO/N2 775–1300 1 N/A [84]

CO/H2-NO JSR CO/H2/O2/NO/N2 800–1400 1 0.1–2.0 [85]


Flow reactor CO/H2/O2/NO/NO2/N2 600–900 20–100 0.06 [10]

CO/H2-NH3 Flame speed CO/H2/NH3/air 298 1 0.7–1.6 [22]


Flame speed CO/H2/NH3/air 298 1–5 0.7–1.6 [83]
Flame speed CO/H2/NH3/air 298 1–10 0.7–1.5 [21]

CH3OH-NOx JSR CH3OH/O2/H2O/NO/N2 700–1100 10 0.3–1.0 [86]


JSR CH3OH/O2/H2O/NO2/N2 700–1100 10 0.3–1.0 [86]
JSR CH3OH/O2/NO/N2 550–1000 10 1.0 [87]
Flow reactor CH3OH/O2/H2O/NO/N2 700–1450 1 0.068–2.7 [88]

CH3OH-NH3 Ignition delay CH3OH/O2/Ar/N2 845–1100 40 0.5–2.0 [1]


Flame speed CH3OH/air 298–448 1 0.7–1.8 [23]

CH4/NO Flow reactor CH4/O2/NO/Ar 473–973 6 0.7–2.1 [89]


Flow reactor CH4/O2/NO/N2/CO2 1173–1773 1 0.125–1.4 [12]
Flow reactor CH4/O2/NO/N2/H2O 800–1250 1 0.12 [90]
JSR CH4/O2/NO/Ar 650–1200 1 0.5–2 [91]
JSR CH4/O2/NO/N2 800–1150 1 0.1 [92]
JSR CH4/O2/NO/N2 800–1150 10 0.5–1.0 [93]
Burner-stabilized flame CH4/NO/O2/N2 400–1800 0.05 0.95 [94]
Flame speed CH4/NO/air 298 1 0.4–1.8 [16]
Flame speed CH4/O2/NO/N2/Ar 298 1 0.6–1.5 [95]

CH4/NO2 Shock tube CH4/NO2/He/Ar 1213–1913 1 0.5 [96]


JSR CH4/O2/NO2/Ar 650–1200 1 0.5–2 [91]
Flow reactor CH4/O2/NO/NO2/N2 598–898 20–100 0.04–116 [11]
Flow reactor CH4/O2/NO2/N2/H2O 800–1250 1 0.12 [90]
Ignition delay CH4/O2/NO2/N2/Ar 901–1650 15–30 1.0 [62]
Ignition delay CH4/O2/NO2/Ar 1016–1981 5–16 1 [97]

CH4/NH3 Flow reactor CH4/NH3/NO/NO2/O2/Ar 730–1035 1 N/A [98]


Flow reactor CH4/NH3/O2/He 1200–2000 1 1 [26]
Flow reactor CH4/NH3/O2/N2 973–1773 1 0.13–1.55 [25]
JSR CH4/NH3/O2/He 600–1200 1 0.5–2.0 [26]
Ignition delay CH4/NH3/air 1369–1804 2–5 0.5–2 [99]
Ignition delay CH4/NH3/O2/N2/Ar 900–1100 20–40 0.5–2 [100]
Flame speed CH4/NH3/air 298 1–5 0.6–1.6 [101]
Flame speed CH4/NH3/air 298 1–5 0.7–1.3 [102]
Flame speed CH4/NH3/air 298 1 0.7–1.5 [82]
Burner-stabilized flame CH4/NH3/O2/N2 400–1800 0.05 0.95 [94]

Fig. 3. Laminar flame speeds of stoichiometric (a) NH3/syngas/air, (b) NH3/CH3OH/air and (c) NH3/CH4/air at unburned gas temperature of 298 K. Symbols are
measurements from [21,23,82,83,101]. Solid and dashed lines are the simulation results by the present and Glarborg 2018 models [2], respectively.

6
X. Zhang et al. Fuel 341 (2023) 127676

Fig. 4. Laminar flame speeds of (a) NH3/syngas/air, (b) NH3/CH3OH/air and (c) NH3/CH4/air at unburned temperature of 298 K. Symbols are measurements from
[22,23,82,83,101,102]. Solid and dashed lines are the simulation results by the present and Glarborg 2018 models [2], respectively.

Fig. 5. Ignition delay times of stoichiometric NO2/CH4/O2/Ar/N2 mixtures at (a) 5 atm, (b) 10 atm, (c) 16 atm, (d) 15 atm and (e) 30 atm. Symbols are mea­
surements from [62,97]. Solid and dashed lines are the simulation results by the present and Glarborg 2018 models [2], respectively.

Fig. 6. Ignition delay times of (a-c) NH3/CH3OH/O2/Ar/N2 mixtures at 40 atm, (d-f) NH3/CH4/O2/Ar/N2 mixtures at 20 and 40 atm. Symbols are measurements
from [1,100]. Solid and dashed lines are the simulation results by the present and Glarborg 2018 models [2], respectively.

7
X. Zhang et al. Fuel 341 (2023) 127676

Fig. 7. (a-b) Mole fractions of CO in the atmospheric flow reactor oxidation of CO/H2O/O2/N2 with and without NO; Mole fractions of (c) NH3 and (d) NO in the
atmospheric flow reactor oxidation of NH3/H2O/O2/NO/N2 with and without CO. Symbols are measurements from [81,84]. Solid and dashed lines are the simulation
results by the present and Glarborg 2018 models [2], respectively.

S21 in the Supplementary material. In general, the present model can times of NH3/CH3OH mixtures is mainly due to the CH3OH subset, since
reasonably predict these experimental targets and has a marginally it fails to predict the ignition delay times of pure CH3OH. Sensitivity
better performance than the Glarborg 2018 model [2]. analysis for pure CH3OH mixture at ϕ = 1.0, Pc = 40 atm and Tc = 850 K
was performed based on both the present and Glarborg model [2], as
3.1.2. Ignition delay times seen from Fig. S22 in the Supplementary Material. It shows that R3
Ignition delay times measured in shock tubes are helpful to examine (CH3OH + HO2 = CH2OH + H2O2) plays a dominant role in promoting
the intermediate- and high-temperature oxidation mechanism, while the ignition in both models. The rate constant of R3 updated in this
those obtained in rapid compression machines (RCMs) are valuable for model is around 5.5 times higher than that in the Glarborg model [2],
validations at low and intermediate temperatures. Fig. 5 presents the which can explain the large prediction discrepancies between the pre­
comparison between measured and simulated ignition delay times for sent and Glarborg model [2]. Fig. 6(d – f) presents the comparison be­
NO2/CH4 mixtures. The ignition delay data shown in Fig. 5(a – c) were tween the measured and predicted ignition delay times for NH3/CH4
obtained from a shock tube study at various pressures and doping ratios mixtures. The experimental data were taken from RCM measurements
[97]. Both the present and Glarborg 2018 models [2] can capture pure [100]. With increasing CH4 content, the ignition reactivity is enhanced.
CH4 oxidation behavior over 5–16 atm. With increasing content of NO2, Both the present and Glarborg 2018 models [2] can reasonably predict
the present model slightly underpredicts ignition delay times. Compared this experimental data, although the present model has better perfor­
with the Glarborg 2018 model [2], the present model shows better mance under some cases. Additional validations against the ignition
performance. The ignition delay time data shown in Fig. 5(d and e) were delay times measured in shock tubes can be found in Figs. S23-S24 in the
taken from RCM measurements [62]. Both models can reasonably pre­ Supplementary material. It shows that both models perform quite well.
dict pure CH4 cases at 15 and 30 atm, while the present model better
captures NO2/CH4 mixture cases, especially at higher pressures and
lower temperatures. 3.2. Speciation data
Fig. 6(a – c) presents the comparison of ignition delay times of NH3/
CH3OH mixtures between measurements and simulations. The experi­ 3.2.1. NH3/NO/syngas
mental data were obtained in an RCM with varying CH3OH contents and To further constrain the present model, speciation profiles of fuels,
equivalence ratios [1]. With increasing CH3OH content, the ignition intermediates and products were also selected to test the present model.
occurs at lower temperatures, indicating CH3OH blending can enhance Fig. 7(a and b) presents the measured and predicted mole fraction
ignition reactivity of NH3. The present model can reasonably predict the profiles of CO in the flow reactor oxidation of CO/NO/H2O/O2/N2
ignition delay times of both pure CH3OH and NH3/CH3OH mixtures. In mixtures [81]. Both the measured and predicted results show that
contrast, the Glarborg 2018 model [2] fails to predict these data. The blending NO results in higher temperature windows for CO conversion.
deficiency of the Glarborg model [2] in predicting the ignition delay Both the present and Glarborg 2018 models [2] can reasonably capture
the experimental data. Fig. 7(c and d) presents the validation results

8
X. Zhang et al. Fuel 341 (2023) 127676

in high-pressure JSR oxidation. The validation results against the


speciation data in the oxidation of CH3OH/NOx/O2/N2 mixtures are
presented in Fig. 8 and Figs. S33-S40 in the Supplementary material. In
general, the predicted results between the present and Glarborg 2018
models [2] are close, both of which can reasonably capture the experi­
mental data.

3.2.3. NH3/NOx/Ch4
Figs. 9 and 10 present the measured and predicted speciation data in
the JSR oxidation of CH4/NO/O2/Ar and CH4/NO2/O2/Ar mixture,
respectively. Compared with the Glarborg 2018 model [2], the present
model can better predict the temperature windows of fuel conversion
and the yields of products. Additional validation results against the
speciation data measured in flow reactors, JSRs, shock tubes and pre­
mixed flames can be found in Figs S41-S51 in the Supplementary ma­
terial. In general, the present model has similar performance with the
Glarborg 2018 model [2] against the speciation data obtained at low and
atmospheric pressures, while can better predict the yields of products in
the flow reactor oxidation at high pressures [11], as seen from Fig. S43
Fig. 8. Mole fractions of (a) CH3OH, (b) CH2O, (c) CO and (d) CO2 in the JSR in the Supplementary material.
oxidation of CH3OH/NO/O2/N2 mixture at 10 atm, ϕ = 1.0 and τ = 1 s. Figs. 11 and 12 present the speciation data measured in the JSR
Symbols are measurements from [87]. Solid and dashed lines are the simulation oxidation of NH3/CH4/O2/He mixtures at fuel-lean and fuel-rich con­
results by the present and Glarborg 2018 models [2], respectively.
ditions, respectively. Compared with the Glarborg 2018 model [2], the
present model can better predict the temperature windows of fuel con­
against the speciation data measured in flow reactor oxidation of CO/ version and the yields of products, especially under fuel-lean condition.
NH3/NO/H2O/O2/N2 mixtures [84]. Both the present and Glarborg The over-predicted reaction rates shown in Fig. 11 by the Glarborg
2018 models [2] can accurately capture the effects of CO blending on the model [2] are mainly due to the deficiency of NH3 subset. In our pre­
conversion of NH3 and NO, while the present model has better pre­ vious study [5], it was shown that the Glarborg model [2] greatly over-
dictions in capturing NO conversion temperature windows. Additional predicts the NH3 conversion under lean conditions, since the rate con­
validations against speciation data measured in high-pressure flow re­ stant of the chain-termination reaction, i.e. NH2 + HO2 = NH3 + O2, is
actors and JSRs are provided in Figs S25-S30 in the Supplementary largely under-estimated. Additional validations against the speciation
material. In general, the present model can adequately predict these data measured in flow reactors and burner-stabilized flames are pre­
experimental data. sented in Figs S52-S59 in the Supplementary material. In general, both
the present and Glarborg 2018 models [2] can reasonably capture these
3.2.2. NH3/NO/Ch3oh experimental data.
Model validations against speciation data in the oxidation of pure
CH3OH are presented in Figs S31 and S32 in the Supplementary mate­ 4. Discussion
rial. Compared with the Glarborg 2018 model [2], the present model can
better capture the temperature windows of CH3OH conversion in at­ When blending C1 fuels in NH3 combustion, C1 fuel molecules and
mospheric flow reactor oxidation, and better predict the yield of CH2O their radicals can either react with H2-related radicals like H, O, OH and

Fig. 9. Mole fractions of (a) CH4, (b) NO2, (c) NO, (d) CH2O, (e) CO and (f) CO2 in the JSR oxidation of CH4/NO/O2/Ar mixtures at 1 atm and τ = 1.5 s. Symbols are
measurements from [91]. Solid and dashed lines are the simulation results by the present and Glarborg 2018 models [2], respectively.

9
X. Zhang et al. Fuel 341 (2023) 127676

Fig. 10. Mole fractions of (a) CH4, (b) NO2, (c) NO, (d) CH2O, (e) CO and (f) CO2 in the JSR oxidation of CH4/NO2/O2/Ar mixtures at 1 atm and τ = 1.5 s. Symbols
are measurements from [91]. Solid and dashed lines are the simulation results by the present and Glarborg 2018 models [2], respectively.

Fig. 11. Mole fractions of (a) NH3, (b) NO, (c) C2H6, (d) CH4, (e) CO and (f) CO2 in the JSR oxidation of CH4/NH3/O2/He mixture at 1 atm, ϕ = 0.5 and τ = 1.5 s.
Symbols are measurements from [26]. Solid and dashed lines are the simulation results by the present and Glarborg 2018 models [2], respectively.

Fig. 12. Mole fractions of (a) NH3, (b) NO, (c) C2H6, (d) CH4, (e) CO and (f) CO2 in the JSR oxidation of CH4/NH3/O2/He mixture at 1 atm, ϕ = 2.0 and τ = 1.5 s.
Symbols are measurements from [26]. Solid and dashed lines are the simulation results by the present and Glarborg 2018 models [2], respectively.

10
X. Zhang et al. Fuel 341 (2023) 127676

Fig. 13. Sensitivity analysis results of the present model for stoichiometric 50 % NH3/50 % C1 fuels/air under (a-c) Tu = 298 K, Pu = 1 atm and (d-e) Tc = 850 K, Pc
= 40 atm. C/N interaction reactions are highlighted with red color. (For interpretation of the references to color in this figure legend, the reader is referred to the web
version of this article.)

HO2, or react with N-containing species like NH, NH2, NO and NO2. R53. This reaction has a much higher rate constant than the C/N
Therefore, the C/H and C/N interaction reactions could be competitive. interaction reactions (i.e. CO + NH2 = HNCO + H and CO + NO2 = CO2
To study the role of cross reactions between C- and N-containing species, + NO), as seen from Fig. 14a. Besides, OH radical has much higher
sensitivity analyses were performed for the NH3/syngas, NH3/CH3OH concentration than NH2 and NO2, indicating the C/H cross reaction R53
and NH3/CH4 mixtures under both flame and autoignition conditions, as plays a dominant role in NH3/syngas oxidation compared with C/N
seen from Fig. 13. The results show that the C/N interaction reactions cross reactions.
are not sensitive under flame conditions, while they can be sensitive
CO + OH = CO2 + H (R53)
reactions under autoignition conditions. To explain the sensitivity ana­
lyses results, rate constants of C/N interaction reactions are compared In NH3/CH3OH oxidation, the H-abstraction reactions of CH3OH by
with those between C-containing species and H2-related radicals, as seen H2-related and N-related radicals are important chain-propagation re­
from Fig. 14. Meanwhile, the critical radical pools of stoichiometric actions and are responsible for CH3OH conversion. Their rate constants
NH3/C1 fuel oxidation under typical flame and ignition conditions are are compared in Fig. 14b. It is found that the H-abstraction reactions of
compared in Fig. 15. The peak mole fractions of H, O, OH, HO2, NH, CH3OH by H2-related radicals (i.e. H, O, OH, black lines in Fig. 14b) are
NH2, NO, NO2 are predicted by the present model under flame propa­ more than one order of magnitude higher than those by N-related rad­
gation and ignition conditions. icals (i.e. NH2, NO, NO2, red lines in Fig. 14b), while the concentration
In NH3/syngas oxidation, the most important C/H cross reaction is levels between H2-related radicals and N-related radicals are in the same

11
X. Zhang et al. Fuel 341 (2023) 127676

Fig. 14. Rate constants comparison between selected C/H and C/N cross reactions. The rate constants are taken from the present model.

Fig. 15. Predicted radical pool distributions in the oxidation of stoichiometric 50 % NH3/50 % C1 fuels under typical (a) flame (unburned temperature of 298 K,
pressure of 1 atm) and (b) ignition (850 K, 40 atm) conditions.

order of magnitude (see Fig. 15). Therefore, for the CH3OH related re­ = CH3O + OH and R48 (CH3 + NO2 = CH3O + NO) are close, indicating
actions, C/N interaction reactions can hardly compete with C/H cross the C/N interaction reaction R48 could play an important role under
reactions in NH3/CH3OH oxidation. As seen from Fig. 13e, the sensi­ ignition conditions. This analysis can explain the reactions between CH3
tivity coefficients of H-abstraction reactions of CH3OH by NH2/NO2 are and NO2 become sensitive reactions under ignition conditions, as shown
two orders of magnitude smaller than that by HO2. The rate constants of in Fig. 13.
CH2OH related reactions are not compared, since the contribution of In summary, the rate constants of C/H cross reactions are generally
R41 (CH2OH + O2 = CH2O + HO2) is dominant compared with the re­ faster than (or close to) those of C/N ones in NH3/C1 systems. Under
actions between CH2OH and radicals in NH3/CH3OH oxidation [1]. flame conditions, H2-related radicals (i.e. H, O, OH) have generally
In NH3/CH4 oxidation, the rate constants of H-abstraction reactions higher concentrations than reactive N-related radicals (NH, NH2, NO2).
of CH4 by H2-related and N-related radicals are compared in Fig. 14c. Therefore, under these conditions, C-containing species are more likely
Similar to NH3/CH3OH oxidation, the concentrations between H2- to react with H2-related radicals rather than N-related ones, making the
related radicals and N-related radicals are in the same order of magni­ C/N cross reactions less competitive. This can explain the conclusions
tude (see Fig. 15), while the rate constants of CH4 + H2-related radicals drawn in previous studies under flame propagation conditions [23]—
are much higher than those of CH4 + N-related reactions (see Fig. 14c), the cross reactions between C- and N-containing species are insignifi­
making the C/N cross reactions less competitive. Unlike CH2OH radical, cant. In contrast, under low- and intermediate-temperature conditions, i.
CH3 radical has multiple important consumption pathways. As seen e. typical ignition conditions, the concentrations of H and O radicals
from Fig. 14d, the rate constant of R4 (CH3 + H(+M) = CH4(+M)) is in decrease, while those of NH2 and NO2 increase, thereby making C/N
the same order of magnitude with R47 (CH3 + NH2 = CH3NH2), and the interaction reactions possible to compete with C/H reactions.
concentrations between H and NH2 in NH3/CH4 oxidation are close
under ignition condition (see Fig. 15b). Therefore, R47 could be 5. Conclusions
competitive with R4 under ignition conditions. In addition, under
ignition conditions, HO2 and NO2 become important radicals. The In this work, a comprehensive kinetic model for NH3/syngas/
concentration of NO2 is much higher than HO2 in the oxidation of NH3/ CH3OH/CH4 was proposed based on our ongoing modeling studies for
CH4 mixture (see Fig. 15b), while the rate constants between CH3 + HO2 pure NH3 and small hydrocarbon fuels. Critical cross reactions between

12
X. Zhang et al. Fuel 341 (2023) 127676

C- and N-containing species were evaluated based on previous experi­ [4] Li Y, Zhou C-W, Somers KP, Zhang K, Curran HJ. The oxidation of 2-butene: A
high pressure ignition delay, kinetic modeling study and reactivity comparison
mental, modeling and theoretical calculation studies. The proposed
with isobutene and 1-butene. Proc Combust Inst 2017;36(1):403–11.
model was tested against previous experimental data for both NH3/C1 [5] Zhang X, Moosakutty SP, Rajan RP, Younes M, Sarathy SM. Combustion
and NOx/C1 fuels, covering temperatures of 473–2000 K, pressures of chemistry of ammonia/hydrogen mixtures: Jet-stirred reactor measurements and
0.04–100 atm, and equivalence ratios of 0.04–116. In general, the pre­ comprehensive kinetic modeling. Combust Flame 2021;234:111653.
[6] Glarborg P, Alzueta MU, Dam-Johansen K, Miller JA. Kinetic modeling of
sent model adequately predicts all the selected validation targets. hydrocarbon/nitric oxide interactions in a flow reactor. Combust Flame 1998;115
Compared with the previous Glarborg 2018 model [2], the present (1):1–27.
model can better predict laminar flame speeds and ignition delay times [7] Miller JA, Bowman CT. Mechanism and modeling of nitrogen chemistry in
combustion. Prog Energy Combust Sci 1989;15(4):287–338.
obtained in RCMs. The present model can not only be used to predict the [8] Dagaut P, Lecomte F, Chevailler S, Cathonnet M. Experimental and detailed
combustion of NH3/C1 fuel mixtures, but is also capable to predict the kinetic modeling of nitric oxide reduction by a natural gas blend in simulated
mutual interaction of NOx/C1 fuels. It is found that the rate constants of reburning conditions. Combust Sci Technol 1998;139(1):329–63.
[9] Glarborg P, Kristensen PG, Dam-Johansen K, Alzueta MU, Millera A, Bilbao R.
critical C/H cross reactions are generally faster than (or close to) those of Nitric oxide reduction by non-hydrocarbon fuels. Implications for reburning with
C/N cross reactions. Under flame conditions, H2-related radicals (i.e. H, gasification gases. Energy Fuels 2000;14(4):828–38.
O, OH) have higher concentrations than reactive N-related radicals (NH, [10] Rasmussen CL, Hansen J, Marshall P, Glarborg P. Experimental measurements
and kinetic modeling of CO/H2/O2/NOx conversion at high pressure. Int J Chem
NH2, NO2). Therefore, under these conditions, C-containing species are Kinet 2008;40(8):454–80.
more likely to react with H2-related radicals rather than N-related ones, [11] Rasmussen CL, Rasmussen AE, Glarborg P. Sensitizing effects of NOx on CH4
making the C/N cross reactions less competitive. However, under low- oxidation at high pressure. Combust Flame 2008;154(3):529–45.
[12] Mendiara T, Glarborg P. Reburn chemistry in oxy-fuel combustion of methane.
and intermediate-temperature conditions, the concentrations of H and O
Energy Fuels 2009;23(7):3565–72.
radicals become lower, while those of NH2 and NO2 become higher, [13] Frassoldati A, Faravelli T, Ranzi E. Kinetic modeling of the interactions between
making cross reactions between C- and N-containing species possible to NO and hydrocarbons at high temperature. Combust Flame 2003;135(1):97–112.
compete with C/H cross reactions. The present model can be used as the [14] Faravelli T, Frassoldati A, Ranzi E. Kinetic modeling of the interactions between
NO and hydrocarbons in the oxidation of hydrocarbons at low temperatures.
base chemistry subset for NH3/larger hydrocarbon fuels. Combust Flame 2003;132(1):188–207.
[15] Deng Y, Sun Z, Yuan W, Yang J, Zhou Z, Qi F. Exploring NH3 and NOx interaction
chemistry with CH4 and C2H4 at moderate temperatures and various pressures.
CRediT authorship contribution statement Front Energy Res 2022;10.
[16] Mei B, Zhang Y, Li W, Li Y. Characterizing methane and nitric oxide interaction in
Xiaoyuan Zhang: Conceptualization, Methodology, Validation, oxygen-free outwardly propagating spherical flame. Proc Combust Inst 2022.
[17] Yu L, Zhou W, Feng Y, Wang W, Zhu J, Qian Y, et al. The effect of ammonia
Data curation, Writing – original draft. Kiran K. Yalamanchi: Meth­ addition on the low-temperature autoignition of n-heptane: An experimental and
odology. S. Mani Sarathy: Writing – review & editing, Funding modeling study. Combust Flame 2020;217:4–11.
acquisition. [18] Wu X, Wu M, Hou Q, Zhang F. Theoretical investigation on the reaction kinetics
of NO2 with CH3OH and HCHO under combustion conditions. Proc Combust Inst
2022.
[19] Fuller ME, Goldsmith CF. On the relative importance of HONO versus HNO2 in
Declaration of Competing Interest low-temperature combustion. Proc Combust Inst 2019;37(1):695–702.
[20] Glarborg P, Andreasen CS, Hashemi H, Qian R, Marshall P. Oxidation of
methylamine. Int J Chem Kinet 2020;52(12):893–906.
The authors declare that they have no known competing financial [21] Mei B, Ma S, Zhang Y, Zhang X, Li W, Li Y. Exploration on laminar flame
interests or personal relationships that could have appeared to influence propagation of ammonia and syngas mixtures up to 10 atm. Combust Flame 2020;
the work reported in this paper. 220:368–77.
[22] Han X, Wang Z, He Y, Zhu Y, Cen K. Experimental and kinetic modeling study of
laminar burning velocities of NH3/syngas/air premixed flames. Combust Flame
Data availability 2020;213:1–13.
[23] Wang Z, Han X, He Y, Zhu R, Zhu Y, Zhou Z, et al. Experimental and kinetic study
on the laminar burning velocities of NH3 mixing with CH3OH and C2H5OH in
No data was used for the research described in the article.
premixed flames. Combust Flame 2021;229:111392.
[24] Okafor EC, Naito Y, Colson S, Ichikawa A, Kudo T, Hayakawa A, et al.
Acknowledgments Experimental and numerical study of the laminar burning velocity of
CH4–NH3–air premixed flames. Combust Flame 2018;187:185–98.
[25] Mendiara T, Glarborg P. Ammonia chemistry in oxy-fuel combustion of methane.
This paper is based on work supported by the Saudi Aramco Research Combust Flame 2009;156(10):1937–49.
and Development Center FUELCOM3 Program under Master Research [26] Arunthanayothin S, Stagni A, Song Y, Herbinet O, Faravelli T, Battin-Leclerc F.
Ammonia–methane interaction in jet-stirred and flow reactors: An experimental
Agreement Number 6600024505/01. FUELCOM (Fuel Combustion for and kinetic modeling study. Proc Combust Inst 2021;38(1):345–53.
Advanced Engines) is a collaborative research undertaking between [27] Ma S, Zhang X, Dmitriev A, Shmakov A, Korobeinichev O, Mei B, et al. Revisit
Saudi Aramco and KAUST, intended to address the fundamental aspects laminar premixed ethylene flames at elevated pressures: A mass spectrometric
and laminar flame propagation study. Combust Flame 2021;230:111422.
of hydrocarbon fuel combustion in engines, and develop fuel/engine [28] Zhang X, Ye L, Li Y, Zhang Y, Cao C, Yang J, et al. Acetaldehyde oxidation at low
design tools suitable for advanced combustion modes. This work is also and intermediate temperatures: An experimental and kinetic modeling
supported by King Abdullah University of Science and Technology investigation. Combust Flame 2018;191:431–41.
[29] Zhang X, Wang G, Zou J, Li Y, Li W, Li T, et al. Investigation on the oxidation
(KAUST) with funds allocated to the Clean Combustion Research Center.
chemistry of methanol in laminar premixed flames. Combust Flame 2017;180:
20–31.
Appendix A. Supplementary data [30] Zhang X, Mei B, Ma S, Pan H, Wang H, Li Y. Experimental and kinetic modeling
investigation on laminar flame propagation of CH4/CO mixtures at various
pressures: Insight into the transition from CH4-related chemistry to CO-related
Supplementary data to this article can be found online at https://doi. chemistry. Combust Flame 2019;209:481–92.
org/10.1016/j.fuel.2023.127676. [31] Zhang X, Li Y, Cao C, Zou J, Zhang Y, Li W, et al. New insights into propanal
oxidation at low temperatures: An experimental and kinetic modeling study. Proc
Combust Inst 2019;37(1):565–73.
References [32] Zhang X, Lailliau M, Cao C, Li Y, Dagaut P, Li W, et al. Pyrolysis of butane-2,3-
dione from low to high pressures: Implications for methyl-related growth
chemistry. Combust Flame 2019;200:69–81.
[1] Li M, He X, Hashemi H, Glarborg P, Lowe VM, Marshall P, et al. An experimental
[33] Klippenstein SJ, Sivaramakrishnan R, Burke U, Somers KP, Curran HJ, Cai L, et al.
and modeling study on auto-ignition kinetics of ammonia/methanol mixtures at
HȮ2 + HȮ2: High level theory and the role of singlet channels. Combust Flame
intermediate temperature and high pressure. Combust Flame 2022;242:112160.
2022;111975.
[2] Glarborg P, Miller JA, Ruscic B, Klippenstein SJ. Modeling nitrogen chemistry in
[34] Burke U, Metcalfe WK, Burke SM, Heufer KA, Dagaut P, Curran HJ. A detailed
combustion. Prog Energy Combust Sci 2018;67:31–68.
chemical kinetic modeling, ignition delay time and jet-stirred reactor study of
[3] Klippenstein SJ. From theoretical reaction dynamics to chemical modeling of
methanol oxidation. Combust Flame 2016;165:125–36.
combustion. Proc Combust Inst 2017;36(1):77–111.

13
X. Zhang et al. Fuel 341 (2023) 127676

[35] Troe J, Ushakov VG. The dissociation/recombination reaction CH4 (+M) ⇔ CH3 [66] Miller JA, Melius CF, Glarborg P. The CH3+NO rate coefficient at high
+ H (+M): A case study for unimolecular rate theory. J Chem Phys 2012;136(21): temperatures: Theoretical analysis and comparison with experiment. Int J Chem
214309. Kinet 1998;30(3):223–8.
[36] Srinivasan NK, Michael JV, Harding LB, Klippenstein SJ. Experimental and [67] Hennig G, Wagner HGG. Investigation of the CH3 + NO reaction in shock waves.
theoretical rate constants for CH4 + O2 → CH3 + HO2. Combust Flame 2007;149 Ber Bunsen-Ges Phys Chem 1994;98(5):749–53.
(1):104–11. [68] Masaki A, Tsunashima S, Washida N. Rate constant for the reaction of CH3O2
[37] Chen X, Fuller ME, Franklin GC. Decomposition kinetics for HONO and HNO2. with NO. Chem Phys Lett 1994;218(5):523–8.
React Chem Eng 2019;4(2):323–33. [69] Villalta PW, Huey LG, Howard CJ. A temperature-dependent kinetics study of the
[38] Kroupnov AA, Pogosbekian MJ. DFT calculation-based study of the mechanism CH3O2 + NO reaction using chemical ionization mass spectrometry. J Phys Chem
for CO2 formation in the interaction of CO and NO2 molecules. Chem Phys Lett 1995;99(34):12829–34.
2018;710:90–5. [70] Scholtens KW, Messer BM, Cappa CD, Elrod MJ. Kinetics of the CH3O2 + NO
[39] Milks D, Adams TN, Matula RA. Single pulse shock tube study of the reaction reaction: Temperature dependence of the overall rate constant and an improved
between nitrogen dioxide (NO2) and carbon monoxide (CO). Combust Sci Technol upper limit for the CH3ONO2 branching channel. J Phys Chem A 1999;103(22):
1979;19(3–4):151–9. 4378–84.
[40] Freund H, Palmer HB. Shock-tube studies of the reactions of NO2 with NO2, SO2, [71] Bacak A, Bardwell MW, Raventos MT, Percival CJ, Sanchez-Reyna G,
and CO. Int J Chem Kinet 1977;9(6):887–905. Shallcross DE. Kinetics of the reaction of CH3O2 + NO: A temperature and
[41] Brown FB, Crist RH. Further studies on the oxidation of nitric oxide; the rate of pressure dependence study with chemical ionization mass spectrometry. J Phys
the reaction between carbon monoxide and nitrogen dioxide. J Chem Phys 1941;9 Chem A 2004;108(48):10681–7.
(12):840–6. [72] Xing J, Nagai Y, Kusuhara M, Miyoshi A. Reactions of methyl- and ethylperoxy
[42] Poskrebyshev GA. Calculating the rate constant for the NH2•+ CO ⇄ NH2CO• ⇄ radicals with NO studied by time-resolved negative ionization mass spectrometry.
H + NHCO reactions and thermodynamic properties of NH2CO•. Kinet Catal J Phys Chem A 2004;108(47):10458–63.
2015;56(3):245–60. [73] Nguyen HT, Mai TVT, Huynh LK. Detailed kinetic mechanism for CH3OO+NO
[43] Baulch DL, Bowman CT, Cobos CJ, Cox RA, Just T, Kerr JA, et al. Evaluated reaction – An ab initio study. Comput Theo Chem 2017;1113:14–23.
kinetic data for combustion modeling: Supplement II. J Phys Chem Ref Data [74] Atkinson R, Baulch DL, Cox RA, Hampson RF, Kerr JA, Troe J. Evaluated kinetic
2005;34(3):757–1397. and photochemical data for atmospheric chemistry: Supplement IV. IUPAC
[44] Li QS, Lü RH. Direction dynamics study of the hydrogen abstraction reaction subcommittee on gas kinetic data evaluation for atmospheric chemistry. J Phys
CH2O + NH2 → CHO + NH3. J Phys Chem A 2002;106(41):9446–50. Chem Ref Data 1992;21(6):1125–568.
[45] Anastasi C, Hancock DU. NO2 kinetic studies using laser-induced fluorescence. [75] Atkinson R, Baulch DL, Cox RA, Crowley JN, Hampson RF, Hynes RG, et al.
J Chem Soc, Faraday Trans 1988;84(10):1697–706. Evaluated kinetic and photochemical data for atmospheric chemistry: Volume II –
[46] Koda S, Tanaka M. Ignition of premixed methanol/air in a heated flow tube and gas phase reactions of organic species. Atmos Chem Phys 2006;6(11):3625–4055.
the effect of NO2 addition. Combust Sci Technol 1986;47(3–4):165–76. [76] Tyndall GS, Cox RA, Granier C, Lesclaux R, Moortgat GK, Pilling MJ, et al.
[47] Xiao C-X, Yan N, Zou M, Hou S-C, Kou Y, Liu W, et al. NO2-catalyzed deep Atmospheric chemistry of small organic peroxy radicals. J Geophys Res: Atmos
oxidation of methanol: Experimental and theoretical studies. J Mol Catal A: Chem 2001;106(D11):12157–82.
2006;252(1):202–11. [77] Chemkin-Pro ANSYS. A chemical kinetics package for analysis of gas-phase
[48] Dean AM, Bozzelli JW. Combustion chemistry of nitrogen. In: Gardiner WC, chemical kinetics. Reaction Design 2019:R2.
editor. Gas-phase combustion chemistry. New York: New York, NY: Springer; [78] Yalamanchi KK, Tingas E-A, Im HG, Sarathy SM. Screening gas-phase chemical
2000. p. 125–341. kinetic models: Collision limit compliance and ultrafast timescales. Int J Chem
[49] Chai J, Goldsmith CF. Rate coefficients for fuel+NO2: Predictive kinetics for Kinet 2020;52(9):599–610.
HONO and HNO2 formation. Proc Combust Inst 2017;36(1):617–26. [79] Dayma G, Dagaut P. Effects of air contamination on the combustion of hydrogen-
[50] Demissy M, Lesclaux R. Kinetics of hydrogen abstraction by amino radicals from effect of NO and NO2 addition on hydrogen ignition and oxidation kinetics.
alkanes in the gas phase. A flash photolysis-laser resonance absorption study. Combust Sci Technol 2006;178(10–11):1999–2024.
J Am Chem Soc 1980;102(9):2897–902. [80] Mueller MA, Yetter RA, Dryer FL. Flow reactor studies and kinetic modeling of the
[51] Hack W, Kurzke H, Rouveirolles P, Wagner HG. Direct measurements of the H2/O2/NOX and CO/H2O/O2/NOX reactions. Int J Chem Kinet 1999;31(10):
reaction NH2+CH4→NH3+CH3 in temperature range 743≤T/K≤1023. Proc 705–24.
Combust Inst 1988;21(1):905–11. [81] Glarborg P, Østberg M, Alzueta MU, Dam-Johansen K, Miller JA. The
[52] Song S, Golden DM, Hanson RK, Bowman CT, Senosiain JP, Musgrave CB, et al. recombination of hydrogen atoms with nitric oxide at high temperatures. Proc
A shock tube study of the reaction NH2 + CH4 → NH3 + CH3 and comparison Combust Inst 1998;27(1):219–26.
with transition state theory. Int J Chem Kinet 2003;35(7):304–9. [82] Han X, Wang Z, Costa M, Sun Z, He Y, Cen K. Experimental and kinetic modeling
[53] Hennig G, Wagner HGG. A kinetic study about the reactions of NH2(X2B1) study of laminar burning velocities of NH3/air, NH3/H2/air, NH3/CO/air and
radicals with saturated hydrocarbons in the gas phase. Ber Bunsen-Ges Phys NH3/CH4/air premixed flames. Combust Flame 2019;206:214–26.
Chem 1995;99(6):863–9. [83] Wang S, Wang Z, Elbaz AM, Han X, He Y, Costa M, et al. Experimental study and
[54] Yu Y-X, Li S-M, Xu Z-F, Li Z-S, Sun C-C. An ab initio study on the reaction NH2+ kinetic analysis of the laminar burning velocity of NH3/syngas/air, NH3/CO/air
CH4→NH3+CH3. Chem Phys Lett 1998;296(1):131–6. and NH3/H2/air premixed flames at elevated pressures. Combust Flame 2020;
[55] Mebel AM, Lin MC. Prediction of absolute rate constants for the reactions of NH2 221:270–87.
with alkanes from ab Initio G2M/TST calculations. J Phys Chem A 1999;103(13): [84] Alzueta MU, Røjel H, Kristensen PG, Glarborg P, Dam-Johansen K. Laboratory
2088–96. study of the CO/NH3/NO/O2 system: Implications for hybrid reburn/SNCR
[56] Mebel AM, Moskaleva LV, Lin MC. Ab initio MO calculations for the reactions of strategies. Energy Fuels 1997;11(3):716–23.
NH2 with H2, H2O, NH3 and CH4: prediction of absolute rate constants and [85] Dagaut P, Lecomte F, Mieritz J, Glarborg P. Experimental and kinetic modeling
kinetic isotope effects. J Mol Struct 1999;461–462:223–38. study of the effect of NO and SO2 on the oxidation of CO H2 mixtures. Int J Chem
[57] Siddique K, Altarawneh M, Gore J, Westmoreland PR, Dlugogorski BZ. Hydrogen Kinet 2003;35(11):564–75.
abstraction from hydrocarbons by NH2. J Phys Chem A 2017;121(11):2221–31. [86] Dayma G, Ali KH, Dagaut P. Experimental and detailed kinetic modeling study of
[58] Yu Y-X, Li S-M, Xu Z-F, Li Z-S, Sun C-C. Direct dynamics study of the reaction path the high pressure oxidation of methanol sensitized by nitric oxide and nitrogen
and rate constants of NH2+C2H6→NH3+C2H5. Chem Phys Lett 1999;302(3): dioxide. Proc Combust Inst 2007;31(1):411–8.
281–7. [87] Moréac G, Dagaut P, Roesler JF, Cathonnet M. Nitric oxide interactions with
[59] Glarborg P, Bendtsen AB, Miller JA. Nitromethane dissociation: Implications for hydrocarbon oxidation in a jet-stirred reactor at 10 atm. Combust Flame 2006;
the CH3 + NO2 reaction. Int J Chem Kinet 1999;31(9):591–602. 145(3):512–20.
[60] Wollenhaupt M, Crowley JN. Kinetic Studies of the Reactions CH3 + NO2 → [88] Alzueta MU, Bilbao R, Finestra M. Methanol Oxidation and Its Interaction with
Products, CH3O + NO2 → Products, and OH + CH3C(O)CH3 → CH3C(O)OH + Nitric Oxide. Energy Fuels 2001;15(3):724–9.
CH3, over a Range of Temperature and Pressure. J Phys Chem A 2000;104(27): [89] Hoener M, Kasper T. Nitrous acid in high-pressure oxidation of CH4 doped with
6429–38. nitric oxide: Challenges in the isomer-selective detection and quantification of an
[61] Srinivasan NK, Su MC, Sutherland JW, Michael JV. Reflected shock tube studies elusive intermediate. Combust Flame 2022;112096.
of high-temperature rate constants for OH + CH4 → CH3 + H2O and CH3 + NO2 [90] Bendtsen AB, Glarborg P, Dam-Johansen KIM. Low temperature oxidation of
→ CH3O + NO. J Phys Chem A 2005;109(9):1857–63. methane: the influence of nitrogen oxides. Combust Sci Technol 2000;151(1):
[62] Sahu AB, Mohamed AAE-S, Panigrahy S, Saggese C, Patel V, Bourque G, et al. An 31–71.
experimental and kinetic modeling study of NOx sensitization on methane [91] Song Y, Marrodán L, Vin N, Herbinet O, Assaf E, Fittschen C, et al. The sensitizing
autoignition and oxidation. Combust Flame 2022. 238. 111746. effects of NO2 and NO on methane low temperature oxidation in a jet stirred
[63] Matsugi A, Shiina H. Thermal Decomposition of Nitromethane and Reaction reactor. Proc Combust Inst 2019;37(1):667–75.
between CH3 and NO2. J Phys Chem A 2017;121(22):4218–24. [92] Dagaut P, Nicolle A. Experimental and kinetic modeling study of the effect of
[64] Zaslonko IS, Petrov YP, Smirnov VN. Thermal decomposition of nitromethane in sulfur dioxide on the mutual sensitization of the oxidation of nitric oxide and
shock waves: The effect of pressure and collision partners. Kinet Catal 1997;38 methane. Int J Chem Kinet 2005;37(7):406–13.
(3):321–4. [93] Dagaut P, Nicolle A. Experimental study and detailed kinetic modeling of the
[65] Weng J-J, Tian Z-Y, Zhang K-W, Ye L-L, Liu Y-X, Wu L-N, et al. Experimental and effect of exhaust gas on fuel combustion: mutual sensitization of the oxidation of
kinetic investigation of pyrolysis and oxidation of nitromethane. Combust Flame nitric oxide and methane over extended temperature and pressure ranges.
2019;203:247–54. Combust Flame 2005;140(3):161–71.

14
X. Zhang et al. Fuel 341 (2023) 127676

[94] Lamoureux N, Marschallek-Watroba K, Desgroux P, Pauwels J-F, Sylla MD, [99] Xiao H, Lai S, Valera-Medina A, Li J, Liu J, Fu H. Experimental and modeling
Gasnot L. Measurements and modelling of nitrogen species in CH4/O2/N2 flames study on ignition delay of ammonia/methane fuels. Int J Energy Res 2020;44(8):
doped with NO, NH3, or NH3+NO. Combust Flame 2017;176:48–59. 6939–49.
[95] Knyazkov DA, Shmakov AG, Dyakov IV, Korobeinichev OP, De Ruyck J, [100] Shu B, He X, Ramos CF, Fernandes RX, Costa M. Experimental and modeling study
Konnov AA. Formation and destruction of nitric oxide in methane flames doped on the auto-ignition properties of ammonia/methane mixtures at elevated
with NO at atmospheric pressure. Proc Combust Inst 2009;32(1):327–34. pressures. Proc Combust Inst 2021;38(1):261–8.
[96] Mathieu O, Cooper SP, Alturaifi SA, Petersen EL. Assessing NO2-Hydrocarbon [101] Wang S, Wang Z, Chen C, Elbaz AM, Sun Z, Roberts WL. Applying heat flux
Interactions during Combustion of NO2/Alkane/Ar Mixtures in a Shock Tube method to laminar burning velocity measurements of NH3/CH4/air at elevated
Using CO Time Histories. Fuels 2022;3(1):1–14. pressures and kinetic modeling study. Combust Flame 2022;236:111788.
[97] Zhang X, Ye W, Shi JC, Wu XJ, Zhang RT, Luo SN. Shock-induced ignition of [102] Okafor EC, Naito Y, Colson S, Ichikawa A, Kudo T, Hayakawa A, et al.
methane, ethane, and methane/ethane mixtures sensitized by NO2. Energy Fuels Measurement and modelling of the laminar burning velocity of methane-
2017;31(11):12780–90. ammonia-air flames at high pressures using a reduced reaction mechanism.
[98] Sun Z, Deng Y, Song S, Yang J, Yuan W, Qi F. Experimental and kinetic modeling Combust Flame 2019;204:162–75.
study of the homogeneous chemistry of NH3 and NOx with CH4 at the diluted
conditions. Combust Flame 2022;112015.

15

You might also like