Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Fuel 365 (2024) 131179

Contents lists available at ScienceDirect

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

Full Length Article

Chemical kinetic study of gasoline surrogate with ammonia on combustion:


Iso-octane modeling
Zechang Liu a, b, Xu He a, b, *, Guangyuan Feng a, b, Chengyuan Zhao a, b, Xiaoran Zhou a, b,
Zhi Wang c, Qingchu Chen c
a
School of Mechanical Engineering, Beijing Institute of Technology, Beijing 100081, China
b
Collaborative Innovation Center of Electric Vehicles in Beijing, Beijing 100081, China
c
School of Vehicle and Mobility, Tsinghua University, Beijing 100084, China

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

Keywords: To facilitate the transition towards a carbon–neutral society, it is necessary to undertake comprehensive research
iC8H18/NH3 on the utilization of ammonia in gasoline engine applications. In the present work, a new chemical kinetic
Laminar burning velocity mechanism for iso-octane/ammonia (iC8H18/NH3) was proposed and widely validated with the experiment data
Kinetic model
of ignition delay time (IDT) and laminar burning velocity (SL ). The iC8H18 model of Cai [Combust. Flame, 162,
Flame chemistry
2015] and NH3 model of Zhang [Combust. Flame, 234, 2021] was chosen as the based model, and they con­
nected through the chemical coupling between the iC8H18 and the NH2 chemistry. In addition, the influence of C-
N interaction on the prediction ability of iC8H18/NH3 model was also analyzed and found that the H-abstraction
reaction between NH2 and iC8H18 is the key reaction affecting the IDT and SL of mixture. Present work was also
focused on the chemistry kinetic of iC8H18/NH3 mixture. The reaction pathway analysis shows that the addition
of NH3 slightly enhanced the H-abstraction reaction of iC8H18 in tertiary site (cC8H17) and slightly inhibited the
H-abstraction reaction of primary site (aC8H17, bC8H17, and dC8H17). In the end, the mole fraction of NO in
iC8H18/NH3 flames is analyzed. The results show that NH3 addition increased the concentration of NO, but with
the gradual increase of equivalence ratio, the increasing trend of NO concentration gradually decreases. A higher
equivalence ratio can be selected to effectively control the increase of NOx concentration. Through the analysis
of NO formation and reduction reaction pathway, it was found that NH3 addition increases the mole fraction of
NH2 radical, the NO formation pathway was stronger than NO reduction pathway due to the sufficient OH
radical, and finally leading to the increase of NO concentration. However, with the increase of equivalence ratio,
the concentration of NH2 radical exceeds OH, and the number of OH is insufficient in satisfying the NH3-related
NO reaction pathway, resulting in a decrease of the influence of NH3 addition on NO concentration.

ammonia (NH3) offers a notable advantage over hydrogen owing to its


superior energy density. Taking into account the lower heating value
1. Introduction
loss, the overall energy requirements for producing and delivering
ammonia to end-users are estimated to be lower than those for trans­
The combustion of fossil fuels leads to substantial emissions of pol­
porting liquid hydrogen. However, in comparison to conventional hy­
lutants, giving rise to a range of environmental and health issues. As a
drocarbons, the practical utilization of ammonia does come with certain
result, there has been a growing focus on the exploration of hybrid
drawbacks. These drawbacks encompass a slower laminar flame speed,
power devices, innovative combustion techniques, cleaner alternative
higher ignition energy requirements, limited chemical reactivity, and
fuels, and sustainable energy solutions. This encompasses advancements
increased emissions of nitrogen-based pollutants. Previous research has
in storage technology as well [1–8].
suggested using flame enhancement (including hydrogen [10,11],
In recent years, hydrogen has gained significant attention as an
various hydrocarbon fuels like methane [12–17], dimethyl ether [17], n-
efficient and environmentally friendly energy source. However, the
heptane [18,19], aromatic hydrocarbons [14,15], and such oxygen-
inherent physical and chemical properties of hydrogen present chal­
containing fuels as dimethoxymethane (DMM) [16] and methanol
lenges related to safety, storage, and transportation [9]. In contrast,

* Corresponding author.
E-mail address: hhexxu@bit.edu.cn (X. He).

https://doi.org/10.1016/j.fuel.2024.131179
Received 21 November 2023; Received in revised form 9 January 2024; Accepted 6 February 2024
Available online 15 February 2024
0016-2361/© 2024 Elsevier Ltd. All rights reserved.
Z. Liu et al. Fuel 365 (2024) 131179

suppression, combustion characteristics, thermal efficiency, and emis­


Nomenclature sion performance of gasoline. Zhang et al. [41] measured the IDT of
iC8H18/NH3 mixtures in the rapid compressed machine (RCM) and
SL Laminar flame speed developed a simplified mechanism for iC8H18/NH3 blends. However,
P Initial pressure their study shows that the simplified mechanism struggled to accurately
Ti Initial temperature of SL capture the IDT trend. Furthermore, Lubrano et al. [18] examined the SL
XNH3 ammonia mole fraction and chemical kinetic characteristics of n-heptane/NH3 and iC8H18/NH3
IDT Ignition delay time mixtures. They utilized the CRECK model [42], but our previous study
φ Equivalence Ratio [43] found that CRECK model tended to overestimate the SL .
T Initial temperature of IDT In summary, there remains a lack of widely validated kinetic
mechanism for iC8H18/NH3 mixture that accounts for various combus­
tion properties such as SL, IDT, and stable species concentrations. It is
crucial to advance our understanding of the combustion behavior of
[20]) as a viable approach to overcome these hurdles in ammonia these blends and their potential as alternatives in ICE. Therefore, the
combustion. In the context of gas-turbine power generation, methane objectives of the present study are threefold: (1) developing a new ki­
(CH4) has been extensively explored as a flame enhancement fuel and netic mechanism for iC8H18/NH3 mixtures, ensuring that it accurately
numerous studies have been dedicated to this area of research [12–16]. reproduces experimental data; (2) conducting an in-depth analysis of the
When considering the transition towards a carbon-free society, it’s flame chemistry of iC8H18/NH3 blends to shed light on how the intro­
crucial to acknowledge that the exploration of ammonia as a potential duction of NH3 affects the combustion behavior of iC8H18; (3) investi­
fuel in internal combustion engines (ICE), especially in the context of gating the influence of NH3 addition on the concentration of NO and
gasoline engine applications, remains relatively limited. provide a detailed examination of the chemical kinetics responsible for
One vital problem is that the exhaust temperature of gasoline engine NO emissions.
is high⋅NH3, as a nitrogen-containing fuel, when used in the field of
gasoline engine, may cause external pollutant emission. For NOx after 2. Modeling
treatment, currently, the most effective methods involve employing
three-way catalytic (TWC) converters coated with rhodium (Rh), plat­ 2.1. Kinetic modeling
inum (Pt), and palladium (Pd), and selective catalytic reduction (SCR)
converters based on small-pore zeolites (Cu/SSZ-13) [21]. Grannell et al. In this study, the chemical kinetics model is composed of three key
[22] examined the efficacy of employing a TWC for mitigating pollut­ components: the TRF (Toluene Reference Fuel) mechanism representing
ants in an ammonia-gasoline engine, but lean-burn must be absolutely gasoline surrogate fuel, an NH3 mechanism, and the chemical coupling
avoided. They suggest that the TWC can proficiently convert CO, HC, between the iC8H18 and the NH2 chemistry. Detailed mechanisms and
NH3, and NO, maintaining an air–fuel ratio within the range of stoi­ thermodynamic, transport data can be accessed in the supplementary
chiometric and slightly rich conditions of 0.2 %. In addition, for lean materials.
combustion scenarios, the study of Qi et al. [23] pointed out that the The TRF mechanism employed in this study adopts a recently opti­
most substantiated and efficient approach currently verified involves the mized model by Cai and Pitsch [44], in which the rate rules assigned by
use of ammonia slip catalyst (ASC). This catalyst selectively oxidizes Curran et al. [45] were calibrated to match the IDT for iC8H18 in [46],
unburned or secondarily produced NH3 in the exhaust, converting it into which limits the agreement of the model to a broader set of target data.
benign N2. Therefore, the challenge of pollutant emissions from The Cai model has been subjected to comprehensive validation to ensure
ammonia-gasoline blended fuels can be mitigated by employing the its accuracy and reliability.
TWC and ASC methodologies. In the context of the ammonia mechanism, several kinetic models for
Meanwhile, it’s important to note that the composition of base gas­ ammonia combustion have emerged in recent years, each aimed at
oline can vary significantly due to diverse crude oil sources and various elucidating the oxidation of ammonia as a primary fuel component.
processing techniques employed in its production. Therefore, re­ These models, as referenced in the literature, include the works of Han
searchers adopted surrogate fuels for the research of gasoline, such as n- et al. [47], Mei et al. [48], Stagni et al. [49], Okafor et al. [12], Wang
heptane, iso-octane, and toluene [24–32]. In recent years, TRF (n-hep­ et al. [15], Glarborg et al.[50], Zhang et al. [51], and Dai et al. [52],
tane, isooctane, and toluene) has been considered the most suitable Shrestha et al. [53], Konnov et al. [54]. They have been developed to
surrogate fuel of gasoline, and many experimental studies on the capture the intricacies of ammonia combustion and its associated re­
laminar burning velocity of TRF were conducted under different initial actions. In a comparative study by Yin et al. [55], various NH3 models
conditions [33–39]. were assessed for their performance concerning SL , IDT, and the con­
However, it is also should note that within the composition of gas­ centration of stable species. The results revealed that the model pro­
oline, iC8H18 stands out as one of the major constituents and plays a posed by Zhang et al. [51] exhibited superior performance, leading to its
critical role in determining the fuel’s octane rating, which reflects its selection as the NH3 sub-model in the present study. Furthermore, for
resistance to detonation. As a result, studying the behavior of mixtures the interaction mechanisms involving NH3 and C1-C2 species, the
containing iC8H18 and NH3 serves as a fundamental investigation, laying mechanism from Glarborg et al. [50] was adopted and integrated into
the groundwork for more complex gasoline/NH3 blends. Moreover, the present model.
gasoline is a highly complex mixture that comprises a wide range of One of the key objectives of this study is to establish a chemical
hydrocarbon compounds. By initially focusing on the reaction between coupling between the iC8H18 and the NH2 chemistry, including H-
single components, such as iC8H18, the complexity of the problem is abstraction reactions between NH2 and alkanes or alkenes. H-abstrac­
reduced. This simplification makes it more feasible to gain a compre­ tion reactions (R1-R4) play a significant role in influencing the IDT and
hensive understanding of the fundamental chemical reaction kinetics the concentrations of species in iC8H18/NH3 mixtures. It’s noteworthy
and mechanisms involved in these interactions. Such research is vital for that, as of now, there haven’t been any reported reaction rate constants
advancing our knowledge of ammonia’s potential as an alternative fuel available in the literature for reactions R1-R4. In previous studies, Mebel
in the context of gasoline engines as we work towards a sustainable, et al. [56] calculated the reaction rate constants for the primary and
carbon-free future. Limited research has been conducted on the funda­ tertiary H-abstraction reactions of NH2 with iso-butane (i-C4H10) using
mental combustion properties of iC8H18/NH3 blends. Liu et al. [40] the G2M energy, molecular, and transition state parameters within the
explored the impact of NH3 addition on various aspects, including knock framework of transition state theory with tunneling corrections. The rate

2
Z. Liu et al. Fuel 365 (2024) 131179

constants for reactions R1-R4 were estimated by unifying the reactions investigations, a recent study conducted by Jasper et al. [65] employed
of iC8H18 and i-C4H10 to the correct degeneracy level, and then first-principle calculations to develop a novel transport database
employing the rate constant analogy estimation using K2 = K1 (d1 d2 1.5
) , K2, explicitly tailored for direct integration into Chemkin. Konnov et al. [66]
K1 and d2, d1 are the rate constants and chemical degeneracy of iC8H18 incorporated this comprehensive database into a kinetic model of H2
and i-C4H10, respectively. Subsequently, rate constants for reactions R1- combustion and observed significant enhancements in the model per­
R4 were scaled to complete the experiment data, and the final rate formance due to the utilization of updated transport properties. Given
constant of R1-R4 are shown in Table 1. Although the IDT is highly the fact that the H2-O2 oxidation mechanism constitutes a crucial aspect
sensitive to reactions R1-R4, the sensitivity of the pre-exponential factor influencing the flame velocity of hydrocarbon fuels, we also adopted the
is relatively small. Therefore, uncertainties associated with the analogy new transport properties of Konnov model [66] in our present model.
method are expected to have small impact on model predictions [57].
Furthermore, to enhance the accuracy of the model, H-abstraction 2.2. Simulation setup
reactions for small alkanes/alkenes were incorporated from the work of
Thoren et al. [57]. Yin et al. [58] developed a mechanism for C3H8/NH3 In this study, a steady-state one-dimensional laminar premixed
mixtures and validated it using experimental data on species concen­ flame, propagating freely, was simulated utilizing the Chemkin Pro
trations. The H-abstraction reactions of C3H8 attacked by NH2 radicals software [67]. The computational domain was established with a length
related reactions were also integrated into the current mechanism to of 10 cm, employing a maximum of 500 grid points. The adaptive grid
ensure a comprehensive and reliable representation of the chemistry was controlled based on the solution gradient and curvature, with the
involved. control parameters set at 0.1 and 0.1, respectively. The simulation
In the present work, the NOx of TRF/NH3 mixture also needs to be incorporated the Soret effect and utilized mixture-averaged transport
discussed. Recently, Fang et al. [59] established a coupled chemical models. The computation of IDT was conducted using the Close Ho­
kinetics model for TRF/NOx based on IDT experimental data of iC8H18/ mogeneous Batch Reactor module within the Chemkin Pro software. The
NO. This model includes a subset of reactions such as the H-abstraction shock tube was treated as a zero-dimensional adiabatic reactor charac­
of iC8H18 attached by NO or NO2 (R5-R8). These reactions are also terized by constant volume. The ignition point is defined as the time
incorporated into the present model. when the initial temperature rises by 400 K. Additionally, the oxidation
simulation of the jet-stirred reactor was performed using the Perfectly
iC8H18 + NO/NO2 = aC8H17 + HNO/HONO (R5) Stirred Reactor (PSR) module available in the Chemkin Pro software
[31,68].
iC8H18 + NO/NO2 = bC8H17 + HNO/HONO (R6)

iC8H18 + NO/NO2 = cC8H17 + HNO/HONO (R7) 3. Results and discussion

iC8H18 + NO/NO2 = dC8H17 + HNO/HONO (R8) 3.1. Ignition delay time validation
Fig. 1 shows the main reaction schemes described in the current ki­
netics model. It includes typical high-temperature and low-temperature To effectively demonstrate the performance of the current kinetic
reactions of iC8H18, as well as interactions between NH3 and fuel spe­ mechanism, it is essential to validate it against experimental IDT data for
cies. The main high-temperature reaction pathway is shown in the green iC8H18/NH3 mixtures. In Fig. 2, IDT measurements for a 20 mol%
box, and the main low-temperature reaction pathway is shown in the iC8H18/80 % NH3/O2/N2/Ar mixture, conducted by Zhang et al. [41],
blue box. The red arrow indicates the crossed reactions between NH3 are presented. These experiments covered a temperature range of T =
and iC8H18. 630–840 K, pressure range of P = 15–20 bar, and equivalence ratio ϕ =
In addition, previous studies have demonstrated the vital role of 1.0. The experimental measurement error is within 15 %. The solid line
transport characteristics in the simulation of flame propagation, along in Fig. 2 represents the numerical results obtained from the present
with the substantial uncertainty associated with transport parameters model. For comparative purposes, we also compared two other kinetic
[60,61]. Conventional methods currently employed for estimating mechanisms: the Zhang model [41] and the Fang model [59]. The Zhang
transport characteristics commonly rely on the utilization of the 12–6 model combines the iso-octane model developed by Curran et al. [45]
Lennard-Jones (LJ) potential, computed through the implementation of with the ammonia model from Glarborg et al. [50], though it doesn’t
combination rules for similar molecular pairs [62]. However, it has been consider C-N cross reactions. On the other hand, the Fang model is
highlighted by Paul and Warnatz [63] that the repulsive potential of LJ specifically designed for iC8H18/NOx interactions and includes C-N cross
(12–6) tends to be consistently steeper than the actual potential, thereby and NH3-related reactions but excludes the NH2 + iC8H18 reactions.
resulting in inaccurate determination of the temperature-dependent It is evident that both the Zhang and Fang models tend to exhibit
transport properties. In recent years, certain studies have utilized ab certain limitations. They generally underestimate the IDT experimental
initio potentials for calculating the transport properties of classical data at T = 710–770 K while overestimating it at temperatures below
collision pairs (such as H-O2, H2O-H, H-CO, etc.) [64]. Under typical 710 K and it failed to capture various trends of the experimental data. In
flame temperatures, these calculations reveal a substantial deviation contrast, the present model, while still underestimating the experi­
from the conventional 12–6 Lennard-Jones (LJ) potential computations, mental data at P = 15 bar, reasonably captures the overall IDT trend.
thereby highlighting the enhanced diffusion characterization of small The current model’s underestimation of experimental values may pri­
entities such as OH, O, and H. Building upon the aforementioned marily stem from two reasons: (1) lack of experimental data on reactor
volume history. Zhang et al. [41] did not provide experimental data on
the volume history of the RCM. Our simulation directly adopts an iso­
Table 1 choric reactor, which introduces a difference compared to the actual
The main added and updated reactions in the chemical kinetic model. The rate experimental conditions. It is important to note that the use of an iso­
constants are in the form of k = A Tn exp (− Ea/(RT)). Units are mole (mol), choric reactor in simulations might not fully capture the complexities
centimeter (cm), Kelvin temperature (K), second (s), and calorie (cal).
introduced by the dynamic volume changes in the actual experimental
No Reaction k n Ea Source setup. The lack of information on the volume history hampers the
R1 iC8H18 + NH2 = aC8H17 + NH3 1.02E + 02 3.61 3342 pw model’s ability to accurately account for heat losses and other volume-
R2 iC8H18 + NH2 = bC8H17 + NH3 1.05E + 01 3.61 3342 pw related effects; (2) the present model’s underestimation of experi­
R3 iC8H18 + NH2 = cC8H17 + NH3 9.69E + 02 3.16 1925 pw mental data may be caused by Cai model’s underestimation of experi­
R4 iC8H18 + NH2 = dC8H17 + NH3 0.19E + 01 3.61 3342 pw
mental data of iC8H18-air flame.

3
Z. Liu et al. Fuel 365 (2024) 131179

Fig. 1. Ic8H18/NH3 reaction pathway diagram in the present model. The main high-temperature reaction path is shown in the green box, and the main low-
temperature reaction path is shown in the blue box. (For interpretation of the references to colour in this figure legend, the reader is referred to the web version
of this article.)

Fig. 2. The idt data of 20 % iC8H18/80 %NH3 mixture at T = 630–800 K, P = 15–20 bar, ϕ=1.0. Symbol: experiment data of Zhang et al. [41]; Solid line: present
model; Dash line: Fang model; Dot line: Zhang model.

To gain further insights into the behavior of the present model, Fig. 3 addition, at the T = 646 K, R1533: NH2 + HO2 = H2NO + OH can
presents numerical results from the Cai model for the IDT of iC8H18/air promote ignition, because R1533 forms H2NO at low temperature, and
mixtures at T = 640 K- 900 K. And the experiment data were obtained then the overall reaction activity is improved by the rapid oxidation of
from the studies of Atef et al. [69]. Notably, the Cai model demonstrates H2NO →HNO→ NO [49]. It is interesting to note that R2337: iC8H18 +
an underestimation of the IDT of pure iC8H18/air mixtures. NH2 = cC8H17 + NH3 and R2335: iC8H18 + NH2 = aC8H17 + NH3 have
Another important purpose of the present study is to investigate how opposite sensitivity in the T = 795 K. This is mainly because the H-
the C-N cross-reaction affect the iC8H18/NH3 mixture, and to provide abstraction from the primary site (aC8H17) increases the overall reac­
guidance for the development of reduced mechanisms. Fig. 4 shows the tivity, while the H-abstraction from the tertiary site (cC8H17) decreases
sensitivity analysis of the iC8H18/NH3 mixture at P = 15 bar, ϕ = 1.0, T reactivity due to the lack of low temperature chain branching paths and
= 646 K (a) and 795 K (b). First, it can be seen that R1632: NH3 + OH = the production of relatively inactive alkenes [69], so the sensitivity of
NH2 + H2O reaction has the largest negative sensitivity coefficient at all the H-abstraction of iC8H18 attacked by NH2 at different sites is also
temperatures, because R1632 is the main NH3 consumption reaction, different.
consuming a large number of OH radicals and inhibiting ignition. In In conclusion, in C-N cross-reactions, only R2337 and R2335 showed

4
Z. Liu et al. Fuel 365 (2024) 131179

the above analysis of C-N cross interaction on IDT is effective in all


temperature zones.

3.2. Laminar burning velocity validation

In this section, the validation of the present kinetic model for iC8H18/
NH3, is presented. For iC8H18/NH3 mixtures, the SL considered in the
validation varies over ranges of initial temperature, pressure, and
equivalence ratio.

3.2.1. Model validation against the SL data of iC8H18/NH3 mixtures


Fig. 6 compares the measured SL of iC8H18/NH3-air mixtures at P = 1
atm, XNH3 = 0–70 %, and Ti = 400 K (a) and 470 K (b) with the simulated
results from the present model. Ti is different from T shown above, Ti
represents the initial temperature of SL , and T represents the initial
temperature of IDT. It is apparent that the model closely predicts the
measurements forXNH3 = 0–70 % in all of ϕ with high fidelity within the
bounds of experimental uncertainty. This shows that the present model
(model I) can effectively capture the SL of iC8H18/NH3 mixtures at
Fig. 3. IDT profiles of iC8H18 measured experimentally at the UCONN RCM
facility and the corresponding simulations. Solid lines represent simulations
different initial temperature and XNH3 .
using the Cai model. For SL measurements, the spherical flame technique has been widely
considered a highly reliable method, particularly for examining flames
at high pressures. Therefore, the SL of iC8H18/NH3 mixtures at P = 2–4
significant sensitivity coefficients to IDT. In order to more clearly
atm and XNH3 = 0–50 % was validated with the kinetic model (model I).
explain the influence of C-N cross-reaction on IDT, we divided the
As depicted in Fig. 7, the present model (model I) accurately captures
mechanism into three parts: model I: NH3 reactions + iC8H18 reactions
the measured SL at ϕ = 0.8 and 1.0 within the experimental uncertainty,
+ all C-N cross-reactions; II: NH3 reactions + iC8H18 reactions + R1-R4
while slightly underestimating the measured SL at ϕ = 1.2. This may be
reactions; III: NH3 reactions + iC8H18 reactions and removed all the C-N
attributed to the fact that, with the increase of initial pressure, the flame
cross-reactions. The contrast between the different mechanisms can
surface begins to exhibit cellular instability, leading to a higher value of
illustrate the effect of the different C-N cross-reaction on IDT. In addi­
measured SL . In general, the present model can well capture experi­
tion, it should be noted that the experimental data of IDT for iC8H18/
mental data under high pressure.
NH3 by Zhang et al. [41]. was mainly in the middle and low tempera­
Although there are some discrepancies between the simulation and
ture, and the influence of C-N cross-reaction on IDT at high temperature
experimental data for the iC8H18 and NH3 binary blends, the present
was not clear. As shown in Fig. 5, the effect of C-N interaction on the
model accurately captures the experimental trends, demonstrating its
middle and low temperature region is obviously greater than that on the
potential in characterizing flame kinetics and serving as a foundation for
high temperature region, and the effect on IDT in the high temperature
further improvements.
region is basically negligible. Fig. 5 compares the numerical calculation
In addition to IDT, the influence of C-N cross reaction on flame ve­
results obtained by mechanisms I, II, and III of iC8H18/NH3 mixture at P
locity should also be discussed here to guide the development of
= 15 bar, T = 650 K-1100 K, ϕ= 1.0. It can be seen that the C-N inter­
simplified mechanisms. In a previous study by Han et al.[70], it was
action reaction significantly shortens the IDT of the mixture. The
indicated that the influence of C-N cross-reactions on the SL of NH3 bi­
mechanisms I and III both can capture the changing trend of IDT well,
nary mixture could be disregarded. However, Han’s research only
indicating that only the H-abstraction reaction between NH2 and iC8H18
considered the interaction between NH3 and small molecules of C1-C2,
needs to be included in the construction of the reduced mechanism and
without taking into account the impact of the chemical coupling set

Fig. 4. The sensitivity analysis of 20 % iC8H18/80 %NH3 mixture at T = 646 K (a) and 795 K (b), P = 15 bar andϕ = 1.0.

5
Z. Liu et al. Fuel 365 (2024) 131179

30–50 % and ϕ = 1.0 are shown in Fig. 9. The percentage of a species


consumed by a specific route is also indicated on each arrow. The red,
and blue colors indicate XNH3 = 50 %, and 30 %, respectively.
The main iC8H18 high-temperature reaction pathway can be
concluded as: RH → R →β-scission products → fragment. The initial
oxidation of iC8H18 by H-abstraction reaction plays a crucial role in the
overall reaction path. In the H-abstraction reactions, the formation of
cC8H17 is the predominant pathway due to the lower bond energy of
cC8H17. H-abstraction is mainly attacked by H radicals.
With NH3 addition, the reaction pathways for iC8H18 oxidation
remain unaltered, while their relative importance is altered. It is inter­
esting that, as ammonia was added, the proportion of aC8H17, bC8H17
and dC8H17 formation channels in the H-abstraction reaction slightly
decreased, but the proportion of cC8H17 reaction channels increased.
This reaction path is the main channel of iC8H18 consumption, which
seems to indicate that the addition of NH3 introduces a large number of
NH2 radicals to participate in the H-abstraction reaction of iC8H18, and
promotes the initial oxidation. And then, cC8H17 rapidly generates iC4H8
and tC4H9 through beta-sicission in the subsequent high-temperature
reaction, and these species consume the active free radicals, thereby
Fig. 5. The idt data of 20 %iC8H18/80 %NH3 mixture at T = 630–800 K, P = reducing SL , so this may be one of the reasons why the addition of NH3
15 bar, ϕ = 1.0. Symbol: experiment data of Zhang et al. [41]; Solid line: model reduces the flame speed of iC8H18/NH3 mixture.
I; Dash line: model III; Dash dot line: model II.
Fig. 10 (a) and (b) show the NH3 reaction paths ofXNH3 = 30 % (a)
and 70 % (b) atϕ = 1.0 (black) and 1.3 (red), respectively. It can be seen
between NH2 and hydrocarbon fuels. Therefore, in Fig. 8, we compared that the primary oxidation consumption of NH3 is mainly through the H-
the simulated result of models I, II, and III for the SL of iC8H18/NH3 abstraction reaction of OH to generate NH2. In order to better explain
mixtures under different conditions. We divided the mechanism into the subsequent reaction path of NH2, it is divided into five pathways: (i)
three parts: I: NH3 reactions + iC8H18 reactions + all C-N cross- it combines with HO2 to generate H2NO, and H2NO is then converted
reactions; II: NH3 reactions + iC8H18 reactions + R1-R4 reactions; III: into HNO and finally to generate NO; (ii) Combined with iC8H18 to
NH3 reactions + iC8H18 reactions and removed all the C-N cross- generate NH3; (iii) Combined with NO to generate NNH or directly
reactions. It can be observed that, although C-N cross-reactions do generate N2; (iv) converted to NH, and then, NH is then combined with
have an effect on SL , given the experimental errors associated with NO/CH3 to form N2O/CH2NH, N2O is converted to N2 in subsequent
laminar flame speeds themselves, this particular influence can be reactions, and CH2NH is further converted to HCN; (v) combined with O
considered negligible. to generate HNO and finally converted to NO.
Firstly, the influence of XNH3 on reaction pathway was analyzed. In
3.3. iC8H18/NH3 flame chemistry the comparison between Fig. 10 (a) and (b), it can be found that atφ =
1.0 %, the main channel of NH2 radical consumption is (ii), (iv) and (v),
3.3.1. Chemical effect of NH3 in which the reaction channel (iv) accounts for the largest proportion
Reaction path analysis enables a more quantitative tracking of (57.1 %). With the increase of XNH3 , the reaction channel (i) (6.82 %
iC8H18 oxidation and how NH3 addition influences the oxidation routes − 5.16 %) and (v) (26.1 % − 22.3 %) decreased and channel (v)
of iC8H18, further elucidating the kinetic effect. The results of a primary increased (57.1 %-59.6). In general, the proportion of each channel is
reaction pathway analysis of iC8H18/NH3 at Ti = 400 K, P = 1 bar, XNH3 =

Fig. 6. TheSL of iC8H18/NH3-air mixtures measured in constant-volume method at ϕ = 0.8–1.3, P = 1 atm and Ti = 400 K (a) and 470 K (b). Symbols are
experimental data from our previous work [43]. Solid lines show the numerical results for the present model (model I).

6
Z. Liu et al. Fuel 365 (2024) 131179

Fig. 7. TheSL of iC8H18/NH3/air mixture measured in constant-volume method for ϕ = 0.8–1.2, Ti = 400 K and P = 2 atm (a) and P = 4 atm (b). Symbols are
experimental data from our previous work [43]. Solid lines show the numerical results from the present model (model I).

basically unchanged.
Secondly, the influence of equivalence ratio on NH3 reaction
pathway was analyzed. When the XNH3 = 70 % and ϕ = 1.0–1.3, it can be
found that NH3 generation is still mainly through the H-abstraction re­
action of OH to NH2 (94.4 %-94.8 %) with the ϕ increase from 1.0 to 1.3.
However, with the increase of the ϕ, the proportion of (iv) in reaction
channels increased, and the proportion of channels (v) and (i) decreased
rapidly. NH2-NH-N2 becomes the main path of NH3 consumption at high
ϕ. In addition, it should also be noted that as the ϕ increases, the pro­
portion of the reaction path where NH combines with CH3 to form
CH2NH also increases, which is an important reaction path for NOx
generation in hydrocarbon fuels.

3.3.2. Stable species concentration validation


Ammonia contains a nitrogen atom in its molecule; therefore,
ammonia combustion may result in the production of large quantities of
nitric oxides through fuel NOx chemistry [71]. Comprehending the
oxidized and reduced states of NO can assist in overcoming the chal­
lenges of high fuel NOx production in iC8H18/NH3 mixtures.
So far, no experiments have been carried out on the species con­
centration of iC8H18/NH3 mixture, and the discussion in the present
Fig. 8. TheSL data of iC8H18/NH3 mixture at Ti = 400 K, P = 1 bar, ϕ = 1.0. model is mainly carried out through the kinetic model, so it is necessary
Symbol: experiment data; Solid line: model I; Dash line: model II; Dash dot line: to verify the concentration of NO to fully illustrate the validity of the
model III. discussion. Considering that the construction of the mechanism is based
on the stratification theory, it is also valid to verify the species

Fig. 9. Reaction pathway analysis of ic8H18 for Ti = 400 K, P = 1 bar, XNH3 = 30–50 % and ϕ = 1.0.

7
Z. Liu et al. Fuel 365 (2024) 131179

Fig. 10. Reaction pathway analysis of nh3 for Ti = 400 K, P = 1 bar, XNH3 = 30–70 % and ϕ = 1.0 and 1.3.

concentration based on the existing experimental data on H2/NH3, and (1a)NH3 ̅̅̅̅̅→ NH2 ̅̅̅̅→ H2NO ̅̅̅̅̅→ HNO ̅̅̅̅̅→ NO;
+OH,H,O +HO2 +H,O,OH +H,O,OH

C3H8/NH3. Figs. 11 and 12 shows the species concentration distribution


(2a)NH3 ̅̅̅̅̅→ NH2 ̅̅̅̅̅→ NH ̅̅̅→ NO; (3a)NO2 → NO + OH(4a)
+OH,H,O +OH,H,O +O2 +H
of C3H8, NH3, H2 and NO in JSR at different initial conditions. In general,
the current mechanism can better capture the changing trend of species CH3 ̅̅̅̅̅̅̅→ NO.
+HNO,NO2

concentration, especially the prediction of NO concentration. And the main reduction of NO are as follows:
(1b) NO ̅̅̅̅→ NO2 + OH;(2b) NH ̅→ N2O;(3b) R10:NH2 + NO
+HO2 +NO
3.3.3. Effects of NH3 blending on NOx concentration
= NNH + OH and R11:NH2 + NO = N2 + H2O. R10 and R11 are the
In Fig. 13, the mole fraction of NOx (NO + NO2 + N2O) is displayed
dominant NO reduction step under all conditions, especially in lean
for iC8H18/NH3 mixtures at Ti = 400 K, P = 1 bar, and XNH3 = 0–70 %.
flames.
As expected, NH3 addition increases the NOx mole fraction of iC8H18/
Based on the ROP analysis of NO, it can be concluded that channels
NH3 mixtures. However, it can be observed that in the case of a rich-
(3a) and (1b) are the primary generation and consumption reactions of
flame, as ϕ increases, the sensitivity of NOx concentration to XNH3
NO. In addition, in the NO generation channels (1a)-(3a), various steps
gradually decreases. For example, at XNH3 = 50 % and XNH3 = 70 %, the
such as the initial decomposition of NH3, the rapid conversion of NH2 to
NOx mole fraction has almost the same value as that at ϕ = 1.3. This
NO via different pathways, and the conversion of NO2 to NO all require
indicates that increasing the blending ratio while increasing the equiv­
H, O, OH radicals. Furthermore, Miller et al.’s research [72] indicates
alent ratio can effectively reduce NOx concentration. Additionally, as
that NO reduction is promoted by the influence of the radical pool
observed from the graph, NO is the predominant species in NOx, with
concentration, which is affected by the branching ratio α = K10/(K10 +
concentrations of species such as N2O and NO2 significantly lower than
K11) and the residence time of NNH. Therefore, in summary, active
that of NO. Consequently, the ensuing discussion will primarily focus on
radicals such as H, O, OH, and NH2 are crucial species influencing NO
the concentration of NO.
concentration and need to be further discussed.
In this section, we aim to elucidate kinetic insights into the signifi­
Due to the nonlinear variation in NO concentration with increasing
cant variation of NO concentration. Through the analysis of the rate of
equivalence ratio, the discussion is divided into two parts. Fig. 14 il­
production (ROP) of NO, the main kinetic reaction pathway of NO was
lustrates the mole fractions of H, O, OH, and NH2 radicals for ϕ = 1.0 (a)
determined. The relevant paths of NO can be divided into NO generation
and 1.3 (b), withXNH3 = 0–70 %. From Fig. 14(a), it can be observed that
pathway and NO reduction pathway.
as the XNH3 increases, the concentrations of H, O, and OH radicals
The main pathways of NO generation are as follows:
gradually decrease, while the NH2 radical increases. Glarborg et al. [50]

Fig. 11. The stable species concentration for c3H8/NH3 mixture in JSR atϕ = 1.0, P = 1 atm, τ=1s and XC3H8 = 20 % (a) and 50 % (b). Symbol: the experiment data
[55]; Line: present model.

8
Z. Liu et al. Fuel 365 (2024) 131179

Fig. 12. The stable species concentration for h2/NH3 mixture in JSR atϕ = 1.0, P = 1 atm, τ=1s and XH2 = 30 % (a) and 10 % (b). Symbol: the experiment data [51];
Line: present model.

Consequently, with an increasing equivalence ratio, the sensitivity of NO


to the mole fraction of NH3 decreases.
The analysis above regarding NO suggests that compared with H, O
and NH2 radicals, OH appears to be the most critical radical influencing
NO concentration. To further illustrate the impact of radicals on NO
concentration, Fig. 15 displays the relationship between the maximum
NO concentration in the iC8H18/NH3-air flame and the maximum OH, H,
and NH2 radicals. It is evident that, under all XNH3 , H and NH2 radicals
are not particularly sensitive to NO concentration. NO concentration is
relatively sensitive to OH concentration, but with increasing NH3 con­
tent, the sensitivity of NO to OH concentration decreases (R2: 0.98 →
0.96). This decrease is likely mainly due to the rapid increase in NH2
radicals, resulting in an insufficient supply of OH radicals to meet the
overall NO reaction demands. Therefore, it can be concluded that at
lower mixing ratios, OH radical concentration is the primary active
species influencing NO concentration. Effective control of NO concen­
tration can be achieved by regulating the relative quantities of NH2 and
OH radicals.

4. Conclusions
Fig. 13. The mole fraction distribution of nox for ic8H18/NH3 mixture at Ti =
400 K, P = 1 atm and XNH3 = 0–70 %. Present study introduces a novel mechanism for iC8H18/NH3 mix­
tures, building upon the iC8H18 model by Cai and the NH3 model by
pointed out that the concentration of NO formed is determined by the Zhang. These models are interconnected through chemical coupling
competition between NH2 and NO (reduction pathway) or reactions between iC8H18 and NH2 chemistry. The newly developed chemical ki­
with the radical pool (generation pathway). Therefore, the increase in netic model has undergone extensive validation using literature data for
NO concentration with increasing XNH3 indicates that the NO generation iC8H18/NH3, encompassing laminar burning velocity, ignition delay
pathway is faster than its reduction pathway. This might be attributed to time, and stable species concentrations. Overall, our kinetic model
the fact that at low equivalence ratios, although the NH2 radical con­ demonstrates a robust ability to predict laminar burning velocity for
centration increases, OH radicals remain the most abundant radicals in iC8H18/NH3, well within experimental uncertainties. An in-depth anal­
the free radical pool. The abundant presence of OH can provide the ysis of C-N interaction reactions reveals that the H-abstraction reaction
requirements for the initial oxidation of NH3 to generate NH2 and the between NH2 and iC8H18 is pivotal, significantly impacting ignition
conversion of intermediate species like HNO to NO. delay time. This insight serves as a practical guide for constructing a
From Fig. 14(b), it can be seen that an increase in XNH3 still leads to a reduced mechanism.
decrease in the concentrations of H, O, OH, and other radicals, while the Another focus of this study involves analyzing the flame chemistry of
NH2 radical concentration increases. However, atXNH3 = 50 %, NH2 iC8H18/NH3. Our findings highlight that the addition of NH3 primarily
exceeds OH radical concentration, becoming the most abundant radical affects the laminar burning velocity of iC8H18/NH3 mixtures through
in the pool. In the initial oxidation of NH3, a substantial amount of OH is chemical effects⋅NH3 enhances the H-abstraction reaction of iC8H18 in
required. The remaining OH radicals may not be sufficient to meet the tertiary sites (cC8H17) while inhibiting the same reaction in primary sites
demands of reactions (1a)-(3a). This weakening of reaction pathways is (aC8H17, bC8H17, and dC8H17), resulting in a decrease in active H, O, and
evident in our earlier analysis of NH3 reaction pathways that with the OH radical concentrations and subsequently reducing flame velocity.
increase of XNH3 , the percent of these reaction channels decreases. Additionally, NH3 addition increases the concentration of NO, but

9
Z. Liu et al. Fuel 365 (2024) 131179

Fig. 14. The mole fractions of h, o, oh and nh2 in iC8H18/NH3-air mixtures for Ti = 400 K, P = 1 atm, XNH3 = 0–70 % and ϕ = 1.0 (a) and 1.3 (b).

Fig. 15. The relationship between the concentration of no and h (a), oh (c), nh2 (b) radicals for iC8H18/NH3 mixture at P = 1 bar and Ti = 400 K.

10
Z. Liu et al. Fuel 365 (2024) 131179

with an escalating equivalence ratio, the rate of increase in NO con­ [12] Okafor EC, Naito Y, Colson S, Ichikawa A, Kudo T, Hayakawa A, et al. Experimental
and numerical study of the laminar burning velocity of CH4–NH3–air premixed
centration gradually diminishes. This phenomenon is closely tied to the
flames. Combust Flame 2018;187:185–98.
balance of active radicals between OH and NH2 in the fuel. When suf­ [13] Okafor EC, Naito Y, Colson S, Ichikawa A, Kudo T, Hayakawa A, et al.
ficient OH free radicals are available, NH3 addition leads to increase NO Measurement and modelling of the laminar burning velocity of methane-ammonia-
concentration. However, at high equivalence ratios, where the OH air flames at high pressures using a reduced reaction mechanism. Combust Flame
2019;204:162–75.
concentration is inadequate to meet the demands of NH2 consumption [14] Lubrano Lavadera M, Pelucchi M, Konnov AA. The influence of ammonia on the
and generation, the impact of NH3 addition on NO concentration be­ laminar burning velocities of methylcyclohexane and toluene: An experimental and
comes limited. Importantly, for iC8H18/NH3 mixtures, selecting a higher kinetic modeling study. Combust Flame 2022;237.
[15] Wang D, Wang Z, Zhang T, Zhai Y, Hou R, Tian Z-Y, et al. A comparative study on
equivalence ratio proves effective in controlling the increase of NO the laminar C1–C4 n-alkane/NH3 premixed flame. Fuel 2022;324.
concentration. [16] Elbaz AM, Giri BR, Issayev G, Shrestha KP, Mauss F, Farooq A, et al. Experimental
and kinetic modeling study of laminar flame speed of dimethoxymethane and
ammonia blends. Energy Fuels 2020;34(11):14726–40.
CRediT authorship contribution statement [17] Yin G, Li J, Zhou M, Li J, Wang C, Hu E, et al. Experimental and kinetic study on
laminar flame speeds of ammonia/dimethyl ether/air under high temperature and
Zechang Liu: Writing – review & editing, Writing – original draft, elevated pressure. Combust Flame 2022;238.
[18] Lubrano Lavadera M, Han X, Konnov AA. Comparative effect of ammonia addition
Methodology, Investigation. Xu He: . Guangyuan Feng: Data curation. on the laminar burning velocities of methane, n-heptane, and iso-octane. Energy
Chengyuan Zhao: Formal analysis, Data curation. Xiaoran Zhou: Data Fuels 2020;35(9):7156–68.
curation. Zhi Wang: Funding acquisition. Qingchu Chen: Funding [19] Dong S, Wang B, Jiang Z, Li Y, Gao W, Wang Z, et al. An experimental and kinetic
modeling study of ammonia/n-heptane blends. Combust Flame 2022;246.
acquisition. [20] 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
premixed flames. Combust Flame 2021;229.
Declaration of competing interest
[21] Koebel M, Elsener M, Kleemann M. Urea-SCR: a promising technique to reduce
NOx emissions from automotive diesel engines. Catal. Today 2000;59(3):335–45.
The authors declare that they have no known competing financial [22] Grannell S, Assanis D, Gillespie D, Bohac S. Exhaust Emissions From a
interests or personal relationships that could have appeared to influence Stoichiometric, Ammonia and Gasoline Dual Fueled Spark Ignition Engine. 2009.
[23] Y. Qi, W. Liu, S. Liu, W. Wang, Y. Peng, Z. Wang, A review on ammonia-hydrogen
the work reported in this paper. fueled internal combustion engines, eTransportation 2023;18.
[24] H. Chu, L. Xiang, X. Nie, Y. Ya, M. E.J. Gu, Laminar burning velocity and pollutant
Data availability emissions of the gasoline components and its surrogate fuels: a review. Fuel 2020;
269:117451.
[25] Zhang C, Chen L, Ding S, Xu H, Li G, Consalvi J-L, et al. Effects of soot inception
No data was used for the research described in the article. and condensation PAH species and fuel preheating on soot formation modeling in
laminar coflow CH4/air diffusion flames doped with n-heptane/toluene mixtures.
Fuel 2019;253:1371–7.
Acknowledgment [26] Zhou L, Xiong G, Zhang M, Chen L, Ding S, de Goey LPH. Experimental study of
Polycyclic Aromatic Hydrocarbons (PAHs) in n-Heptane laminar diffusion flames
The research was financially supported by the National Natural from1.0 to 3.0bar. Fuel 2017;209:265–73.
[27] Liang S, Li Z, Gao J, Ma X, Xu H, Shuai S. PAHs and soot formation in laminar
Science Foundation of China (Grant No. 52376092) and the Ningbo partially premixed co-flow flames fuelled by PRFs at elevated pressures. Combust
Major Research and Development Plan Project (Grant No.2022Z151). Flame 2019;206:363–78.
[28] Chu H, Ren F, Xiang L, Dong S, Qiao F, Xu G. Numerical investigation on
combustion characteristics of laminar premixed n-heptane/air flames at elevated
Appendix A. Supplementary material initial temperature and pressure. J Energy Inst 2019;92(6):1821–30.
[29] Dong W, Hu J, Xiang L, Chu H, Li Z. Numerical investigation on combustion
Supplementary data to this article can be found online at https://doi. characteristics of laminar premixed n-heptane/hydrogen/air flames at elevated
pressure. Energy Fuel 2020;34(11):14768–75.
org/10.1016/j.fuel.2024.131179. [30] Dong W, Jin T, Qiu B, Chu H. Effects of carbon dioxide on the combustion
characteristics of the laminar premixed n-heptane/air flames at elevated pressures.
References J Energy Inst 2021;99:127–36.
[31] Xu H, Liu F, Sun S, Zhao Y, Meng S, Chen L, et al. Flame attachment and kinetics
studies of laminar coflow CO/H2 diffusion flames burning in O2/H2O. Combust
[1] Ma X, Ma Y, Wang Z, Mao J, Liu H, Su F, et al. Optical study on multi-time ignition
Flame 2018;196:147–59.
mixed-mode combustion with gasoline and PODE. Fuel 2023;335:126910.
[32] Hu X, Yu Z, Chen L, Huang Y, Zhang C, Salehi F, et al. Morphological and
[2] Ding H, Zhao J, Zhang Z, Xu K, Fu L, He X. A numerical study on the interaction of
nanostructure characteristics of soot particles emitted from a jet-stirred reactor
droplet collisions and air flow impact in cross-impinging spray. Energy 2023;277.
burning aviation fuel. Combust Flame 2022;236:111760.
[3] Zhao J, Fu L, Ding H, Bai B, Zhang D, Liu J, et al. Numerical simulation of working
[33] Zhang S, Lee TH, Wu H, Pei J, Wu W, Liu F, et al. Experimental and kinetic studies
process and gas-liquid interaction mechanism of air assisted nozzle. Int J
on laminar flame characteristics of acetone-butanol-ethanol (ABE) and toluene
Multiphase Flow 2023;164.
reference fuel (TRF) blends at atmospheric pressure. Fuel 2018;232:755–68.
[4] Zhao J, Ding H, Fu L, Bai B, Zhang D, Liu J, et al. Energy analysis and research on
[34] Sileghem L, Alekseev VA, Vancoillie J, Van Geem KM, Nilsson EJK, Verhelst S,
injection control parameter influence mechanism of air assisted spray system. Fuel
et al. Laminar burning velocity of gasoline and the gasoline surrogate components
2023;346.
iso-octane, n-heptane and toluene. Fuel 2013;112:355–65.
[5] Ma Y, Cui L, Ma X, Wang J. Optical study on spray combustion characteristics of
[35] Liu F, Liu Z, Sang Z, He X, Liu F, Liu C, et al. Kinetic study of the effects of hydrogen
PODE/diesel blends in different ambient conditions. Fuel 2020;272:117691.
blending to toluene reference fuel (TRF)/air mixtures on laminar burning velocity
[6] Zhang W, Zhang Z, Ma X, Awad OI, Li Y, Shuai S, et al. Impact of injector tip
and flame structure. Fuel 2020;274:117850.
deposits on gasoline direct injection engine combustion, fuel economy and
[36] Di Lorenzo M, Brequigny P, Foucher F, Mounaïm-Rousselle C. Validation of TRF-E
emissions. Appl Energy 2020;262:114538.
as gasoline surrogate through an experimental laminar burning speed
[7] Kamil M, Ramadan KM, Olabi AG, Al-Ali EI, Ma X, Awad OI. Economic, technical,
investigation. Fuel 2019;253:1578–88.
and environmental viability of biodiesel blends derived from coffee waste. Renew
[37] Liao Y-H, Roberts WL. Laminar flame speeds of gasoline surrogates measured with
Energy 2020;147:1880–94.
the flat flame method. Energy Fuel 2016;30:1317–24.
[8] Liu F, Liu Z, Sang Z, He X, Sjöberg M, Vuilleumier D, et al. Numerical study and
[38] Hu E, Xu Z, Gao Z, Xu J, Huang Z. Experimental and numerical study on laminar
cellular instability analysis of E30-air mixtures at elevated temperatures and
burning velocity of gasoline and gasoline surrogates. Fuel 2019;256:115933.
pressures. Fuel 2020;271:117458.
[39] He X, Liu Z, Jiang H, Liu F, Yang Q, Jiang Z, et al. The reduction in laminar burning
[9] Kojima Y, Yamaguchi M. Ammonia as a hydrogen energy carrier. Int J Hydrogen
velocity and Markstein length extrapolation uncertainty for TRF-air mixtures. Fuel
Energy 2022;47(54):22832–9.
2023;335:127052.
[10] Wang S, Wang Z, Elbaz AM, Han X, He Y, Costa M, et al. Experimental study and
[40] Liu S, Lin Z, Zhang H, Lei N, Qi Y, Wang Z. Impact of ammonia addition on knock
kinetic analysis of the laminar burning velocity of NH3/syngas/air, NH3/CO/air
resistance and combustion performance in a gasoline engine with high
and NH3/H2/air premixed flames at elevated pressures. Combust Flame 2020;221:
compression ratio. Energy 2023;262.
270–87.
[41] Zhang R, Liu W, Zhang Q, Qi Y, Wang Z. Auto-ignition and knocking combustion
[11] Shrestha KP, Lhuillier C, Barbosa AA, Brequigny P, Contino F, Mounaïm-
characteristics of iso-octane-ammonia fuel blends in a rapid compression machine.
Rousselle C, et al. An experimental and modeling study of ammonia with enriched
Fuel 2023;352.
oxygen content and ammonia/hydrogen laminar flame speed at elevated pressure
and temperature. Proc Combust Inst 2021;38(2):2163–74.

11
Z. Liu et al. Fuel 365 (2024) 131179

[42] http://creckmodeling.chem.polimi.it/menu-kinetics/menu-kinetics-detailed- [58] Yin G, Xiao B, Zhan H, Hu E, Huang Z. Chemical kinetic study of ammonia with
mechanisms. propane on combustion control and NO formation. Combust Flame 2023;249:
[43] Liu Z, He X, Jiang Z, Feng G, Zhao C, Yang Q. Study on the laminar combustion 112617.
characteristics and kinetic of IC8H18/NH3 premixed flames. Fuel 2024;356: [59] Fang R, Saggese C, Wagnon SW, Sahu AB, Curran HJ, Pitz WJ, et al. Effect of nitric
129633. oxide and exhaust gases on gasoline surrogate autoignition: iso-octane experiments
[44] Cai L, Pitsch H. Optimized chemical mechanism for combustion of gasoline and modeling. Combust Flame 2022;236.
surrogate fuels. Combust Flame 2015;162(5):1623–37. [60] Mishra MK, Yetter R, Reuven Y, Rabitz H, Smooke MD. On the role of transport in
[45] Curran HJ, Gaffuri P, Pitz WJ, Westbrook CK. A comprehensive modeling study of the combustion kinetics of a steady-state premixed laminar CO + H2 + O2 flame.
iso-octane oxidation. Combust Flame 2002;129(3):253–80. Int. J. Chem. Kinet. 1994;26(4):437–53.
[46] Fieweger K, Blumenthal R, Adomeit G. Self-ignition of S.I. engine model fuels: A [61] Esposito G, Chelliah HK. Effect of binary diffusion and chemical kinetic parameter
shock tube investigation at high pressure. Combust Flame 1997;109(4):599–619. uncertainties in simulations of premixed and non-premixed laminar hydrogen
[47] Han X, Wang Z, He Y, Liu Y, Zhu Y, Konnov AA. The temperature dependence of flames. Combust Flame 2012;159(12):3522–9.
the laminar burning velocity and superadiabatic flame temperature phenomenon [62] Dagdigian PJ. Combustion simulations with accurate transport properties for
for NH3/air flames. Combust Flame 2020;217:314–20. reactive intermediates. Combust Flame 2015;162(6):2480–6.
[48] Mei B, Ma S, Zhang Y, Zhang X, Li W, Li Y. Exploration on laminar flame [63] Paul P, Warnatz J. A re-evaluation of the means used to calculate transport
propagation of ammonia and syngas mixtures up to 10 atm. Combust Flame 2020; properties of reacting flows. Symp (Int) Combust 1998;27(1):495–504.
220:368–77. [64] Dagdigian PJ, Alexander MH. Exact quantum scattering calculations of transport
[49] Stagni A, Cavallotti C, Arunthanayothin S, Song Y, Herbinet O, Battin-Leclerc F, properties for the H2O–H system. J. Chem. Phys. 2013;139(19):194309.
et al. An experimental, theoretical and kinetic-modeling study of the gas-phase [65] Jasper AW, Kamarchik E, Miller JA, Klippenstein SJ. First-principles binary
oxidation of ammonia. React Chem Eng 2020;5(4):696–711. diffusion coefficients for H, H2, and four normal alkanes + N2. J. Chem. Phys.
[50] Glarborg P, Miller JA, Ruscic B, Klippenstein SJ. Modeling nitrogen chemistry in 2014;141(12):124313.
combustion. Prog Energy Combust Sci 2018;67:31–68. [66] Konnov AA. Yet another kinetic mechanism for hydrogen combustion. Combust
[51] X. Zhang, S.P. Moosakutty, R.P. Rajan, M. Younes, S.M. Sarathy, Combustion Flame 2019;203:14–22.
chemistry of ammonia/hydrogen mixtures: Jet-stirred reactor measurements and [67] Kee R, Grcar J, Smooke M, Miller J, Meeks E. PREMIX: a fortran program for
comprehensive kinetic modelling, Combust Flame 2021;234. modeling steady laminar one-dimensional premixed flames. Sandia Rep 1985;143.
[52] Dai L, Gersen S, Glarborg P, Levinsky H, Mokhov A. Experimental and numerical [68] Li G, Yang W, Tay KL, Yu W, Chen L. A reduced and robust reaction mechanism for
analysis of the autoignition behavior of NH3 and NH3/H2 mixtures at high toluene and decalin oxidation with polycyclic aromatic hydrocarbon predictions.
pressure. Combust Flame 2020;215:134–44. Fuel 2020;259:116233.
[53] Issayev G, Giri BR, Elbaz AM, Shrestha KP, Mauss F, Roberts WL, et al. Combustion [69] Atef N, Kukkadapu G, Mohamed SY, Rashidi MA, Banyon C, Mehl M, et al.
behavior of ammonia blended with diethyl ether. ProCombust Inst 2021;38(1): A comprehensive iso-octane combustion model with improved thermochemistry
499–506. and chemical kinetics. Combust Flame 2017;178:111–34.
[54] Konnov AA. Implementation of the NCN pathway of prompt-NO formation in the [70] Han X, Wang Z, He Y, Zhu Y, Cen K. Experimental and kinetic modeling study of
detailed reaction mechanism. Combust Flame 2009;156(11):2093–105. laminar burning velocities of NH3/syngas/air premixed flames. Combust Flame
[55] Yin G, Xiao B, Zhan H, Hu E, Huang Z. Chemical kinetic study of ammonia with 2020;213:1–13.
propane on combustion control and NO formation. Combust Flame 2023;249. [71] Valera-Medina A, Xiao H, Owen-Jones M, David WIF, Bowen PJ. Ammonia for
[56] Mebel AM, Lin MC. Prediction of absolute rate constants for the reactions of NH2 power. Prog Energy Combust. Sci. 2018;69:63–102.
with alkanes from ab initio G2M/TST calculations. Chem. A Eur. J. 1999;103(13): [72] Miller JA, Klippenstein SJ. Theoretical considerations in the NH2 + NO reaction.
2088–96. J Phys Chem A 2000;104:2061–9.
[57] Thorsen LS, Jensen MST, Pullich MS, Christensen JM, Hashemi H, Glarborg P, et al.
High pressure oxidation of NH3/n-heptane mixtures. Combust Flame 2023;254.

12

You might also like