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

Available online at www.sciencedirect.

com

Proceedings of the Combustion Institute 38 (2021) 2163–2174


www.elsevier.com/locate/proci

An experimental and modeling study of ammonia with


enriched oxygen content and ammonia/hydrogen
laminar flame speed at elevated pressure and temperature
Krishna Prasad Shrestha a, Charles Lhuillier b,c, Amanda Alves Barbosa b,
Pierre Brequigny b, Francesco Contino e, Christine Mounaïm-Rousselle b,
Lars Seidel d,∗, Fabian Mauss a
a Thermodynamics and Thermal Process Engineering, Brandenburg University of Technology, Siemens-Halske-Ring 8,
03046 Cottbus, Germany
b University Orléans, INSA-CVL, PRISME, EA 4229, F45072 Orléans, France
c Thermo and Fluid Dynamics (FLOW), Vrije Universiteit Brussel, Belgium
d LOGE Deutschland GmbH, Burger Chaussee 25, 03044 Cottbus, Germany
e Institute of Mechanics, Materials and Civil Engineering (iMMC), UCLouvain, Belgium

Received 7 November 2019; accepted 7 June 2020


Available online 27 August 2020

Abstract

Laminar flame speeds of ammonia with oxygen-enriched air (oxygen content varying from 21 to 30 vol.%)
and ammonia-hydrogen-air mixtures (fuel hydrogen content varying from 0 to 30 vol.%) at elevated pressure
(1–10 bar) and temperature (298–473 K) were determined experimentally using a constant volume combus-
tion chamber. Moreover, ammonia laminar flame speeds with helium as an inert were measured for the first
time. Using these experimental data along with published ones, we have developed a newly compiled kinetic
model for the prediction of the oxidation of ammonia and ammonia-hydrogen blends in freely propagating
and burner stabilized premixed flames, as well as in shock tubes, rapid compression machines and a jet-stirred
reactor. The reaction mechanism also considers the formation of nitrogen oxides, as well as the reduction of
nitrogen oxides depending on the conditions of the surrounding gas phase. The experimental results from the
present work and the literature are interpreted with the help of the kinetic model derived here. The experi-
ments show that increasing the initial temperature, fuel hydrogen content, or oxidizer oxygen content causes
the laminar flame speed to increase, while it decreases when increasing the initial pressure. The proposed
kinetic model predicts the same trends than experiments and a good agreement is found with measurements
for a wide range of conditions. The model suggests that under rich conditions the N2 H2 formation path
is favored compared to stoichiometric condition. The most important reactions under rich conditions are:
NH2 +NH=N2 H2 +H, NH2 +NH2 =N2 H2 +H2 , N2 H2 +H=NNH+H2 and N2 H2 +M=NNH+H+M. These

∗ Corresponding author.
E-mail addresses: shrestha@b-tu.de (K.P. Shrestha), Lars.Seidel@logesoft.com (L. Seidel).

https://doi.org/10.1016/j.proci.2020.06.197
1540-7489 © 2020 The Combustion Institute. Published by Elsevier Inc. All rights reserved.
2164 K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174

reactions were also found to be among the most sensitive reactions for predicting the laminar flame speed for
all the cases investigated.
© 2020 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

Keywords: Ammonia; Laminar flame speed; Kinetic modeling; Ammonia-hydrogen; NOx

1. Introduction vestigated mechanisms from literature could not


accurately predict the laminar burning velocity.
Recently, Mei et al. [2] performed laminar flame
Ammonia is a carbon free energy carrier. Its speed experiments for oxygen enriched NH3 /O2 /N2
technical application is slightly hindered by the blends. They modified our previous work [4] to
comparable low laminar flame speed. To consider reach an agreement with their experiments. Their
ammonia (NH3 ) as single or blend fuel, it is es- selection of the base model was founded on a pre-
sential to have a deep understanding of the de- vious comparative study, including work from dif-
tailed chemical processes of fuel oxidation to N2 ferent groups.
and H2 O and further oxidation to emission forma- In general, the discrepancies between the kinetic
tion, mainly NO and NO2 . Moreover, the knowl- model predictions are expected as the experimental
edge on NH3 kinetics is also needed to improve database for ammonia oxidation is not yet broad
the general NOx formation model. Ammonia is the enough. Uncertainties are found in the rate kinetic
smallest nitrogen containing fuel molecule. Several data as well as in the thermochemistry. Further ex-
NH3 –NOx mechanisms have been developed in re- perimental data are needed to allow further im-
cent years, each with different conditions of interest provement of the kinetic models. The conclusion
[1–3]. Shrestha et al. [4] published their NH3 model above includes that the further oxidation of inter-
recently which covers a broad range of experimen- mediates to NOx is not yet fully understood. Sev-
tal conditions available from the literature, includ- eral researchers have already put their effort into
ing NOx formation and reduction in dependence the development of an NH3 –NOx kinetic model
on the surrounding environment. Mathieu and Pe- but still, a well-established and reliable model is not
tersen [1] established an ammonia oxidation mech- yet available. As discussed in [2] our recent model
anism, based on experiments in a Shock Tube (ST). [4] was not validated for laminar flames but can be
They compared their work against nine different modified to include these experiments.
literature models to underline the large discrepan- In this study, we present new experimental flame
cies between the ammonia mechanisms. Hayakawa speed data together with a modification of our re-
et al. [5] measured laminar flame speeds of ammo- cent work to include these experiments in our vali-
nia – air in a cylindrical combustion chamber under dation database.
elevated pressure conditions. They compared their The aim of the present work is thus twofold:
data against the prediction of five detailed mecha-
nisms from the literature. Most mechanisms failed
to predict laminar flame experiments. (1) to extend laminar flame speed data already
Xiao and Valera-Medina [6] performed a com- published in [11], in the case of NH3 /H2
parative study on the performance of 12 different blends at higher pressure than atmospheric
NH3 –NOx kinetic mechanisms to identify the best one, and also in the case of neat NH3 with en-
model. They used NH3 /H2 laminar flame speed riched O2 conditions. Moreover, additional
data from [7] and NH3 ignition delay time data NH3 laminar flame speed measurement with
from Mathieu and Petersen [1] to rate the model He as an inert will be provided for the first
predictions. The majority of the kinetic models time to investigate the influence of third body
could not reproduce the experimental data for the coefficients in pressure dependent reactions.
investigated range of equivalence ratios. Recently, (2) to improve our previous model [4], based on
Pochet et al. [8] and He et al. [9] studied autoigni- the improved and broadened experimental
tion of neat NH3 and NH3 /H2 blends in a rapid database.
compression machine (RCM). Both studies as-
sessed the performance of the various NH3 kinetic This work is an ongoing effort to develop a
models including our previous work [4] against reliable and comprehensive mechanism for NH3
their experimental data and concluded that none and NH3 /H2 blends. Although the focus of this
of the mechanisms was able to reproduce the ex- study is on laminar flame speed, the derived model
periments satisfactorily. Han et al. [10] performed is critically tested against the previous validation
an experimental and kinetic modeling investiga- database; i.e. ignition delay times (IDT), speciation
tion on laminar flame speed for NH3 /air and dif- in jet-stirred reactors (JSR), and burner stabilized
ferent NH3 /H2 /CO/CH4 air blends. The three in- flames (BSF).
K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174 2165

2. Experimental set-up and processing dr


K = r2 dtf and Lb the Markstein length for burned
f
gases.
The experiments were performed in the 4.2 L This methodology was applied for flame diame-
(inner diameter 200 mm) spherical stainless steel ter greater than 6.5 mm to avoid any ignition effect
combustion chamber, fully described in [12,13]. and lower than 25 mm corresponding to a burned
The fresh gases are heated up to a maximum ini- gas volume less than 1.6% of sphere volume, to con-
tial temperature of 473 K by means of a heating sider constant pressure chamber and avoid confine-
wire. Before filling the sphere, a vacuum pump al- ment effect.
lows the evacuation of residual gases until a pres- Then the fundamental unstretched laminar
sure <0.009 bar. Then all reactant gases are injected burning velocity SL 0 is obtained with the expan-
by a thermal flow meter. Before being introduced sion factor, ρ b /ρ u as SL0 = ρρub Vs0 , with ρ b and ρ u
in the sphere, the inlet valve heats the mixture up
to the required temperature with a maximum of the burned and the unburned gases densities respec-
2 K uncertainty. To guarantee a perfectly homoge- tively.
neous mixture, a fan runs during few seconds, but For low unstretched laminar burning velocity
is stopped 10 s before the ignition to avoid any per- (<20 cm.s−1 ), radiative heat loss induced error can
turbation. The maximum deviation between the ef- be of importance. Those effects were already ad-
fective initial pressure inside the combustion cham- dressed in [11]. In the present work, using the corre-
ber and the required initial pressure was estimated lation of Yu et al. [16], the worst case is obtained at
about 0.5%. Two tungsten electrodes (1 mm diam- 473 K, 3 bar where an underestimate of about 12%
eter) with a conventional capacitive discharge ig- is calculated for neat ammonia corresponding to an
nition system are used, with an ignition coil time absolute error of 1 cm/s. The error decreases down
charge of 3 ms (i.e. discharge energy of less than to 1% for the fastest flames.
100 mJ). The electrodes gap is adjusted to ensure a
sufficient energy deposit to sustain the flame.
Experiments were carried out at various ini- 3. Kinetic modeling
tial pressure, temperature, global equivalence ra-
tio from 1 to 10 bar, 298–473 K and 0.8–1.4 respec- The kinetic model used in this work follows our
tively. The volume content of H2 , xH2 , was varied previous publication [4] and was revised to predict
from 0 to 30% (in fuel). Synthetic air (20.9% O2 the new experimental data determined in this work
/79.1% N2 ) was used as ambient gas, but experi- and in our recent work [11] for the wide range of
ments were also performed for the neat ammonia conditions. In this work, the reaction rate parame-
cases with a variation of oxygen content, xO2 (23, ter of several reactions was updated based on a lit-
25, 27 and 30% vol, in oxidizer). Last, a set of ex- erature study. The changes were made in the NH3 ,
periments was also conducted by using He as inert NH2 , NH, N2 H2 , and H2 NO sub-mechanisms and
gases for 27 and 30% vol. O2 cases. For each con- modified reactions are listed in Table 1. The model
dition, at least 3 tests were carried out in order to developed in this work is validated against 104 sets
obtain average and standard deviation values, rep- (approximately 500 data points) of laminar flame
resented by error bars in the following figures. speed data (covering φ = 0.7- 1.7, T = 298–473 K,
To measure laminar flame speeds, the Schlieren P = 1–10 bar). The complete kinetic scheme is pro-
technique was implemented thanks to two oppo- vided as supporting information (SI).
site and transparent windows (diameter 70 mm).
The Schlieren set-up is composed by a continu-
ous LED, coupled with a 0.8 mm pinhole and 2
4. Results and discussion
convex mirrors (864 mm focal length). At the fo-
cal point of the second mirror, a 0.5 mm dot is
The symbols in figures are experimental data
placed and two lenses (250 mm and 400 mm) al-
from this work or from the literature, lines are
low to focus the Schlieren images on the Phantom
model prediction. All the simulations are per-
v1210 high-speed CMOS ship. Images of 640 × 800
formed with the LOGEresearch v1.10 [24] soft-
pixels were recorded with a magnification ratio of
ware suit. The model from this work is validated
0.11 mm/pixel. The frame rate was adapted from 5
against 104 data sets of laminar flame speed mea-
to 8 kHz. The temporal evolution of the expand-
surements from different sources (32 data are from
ing spherical flame was then processed to get the
this work and 40 data (for NH3 /H2 ) are from our
flame front radius, rf . The laminar flame propaga-
recent work [11] and others are from published lit-
tion speed, Vs was calculated from the time deriva-
erature [2,5,7,10,11,25–32]).
tive of rf . The non-linear methodology as sug-
gested by Kelley and Law [14] and Halter et al.
[15] was used to estimate the unstretched laminar 4.1. Laminar flame speed of NH3 and NH3 /H2
flame velocity Vs 0 , as ( VVs0 )2 ln ( VVs0 )2 = −2 LVb 0K , with
s s s
Figure 1a displays the laminar flame speeds of
K the stretch rate impacting the flame, estimated by
NH3 /air at 1 atm and 298 K as a function of the
2166 K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174

Table 1
List of reactions with updated rate parameter in the detailed model, found in the SI.
Reactions Source of adopted rate
NH2 +H(+M)=NH3 (+M) Altinay and Macdonald [17]
NH3 +H=NH2 +H2 Nguyen and Stanton [18]
NH2 +H=NH+H2 Samu et al. [19]
NH2 +O=HNO+H Miller et al. [20] /2
NH2 +HO2 =NH3 +O2 Sumathi and Peyerimhoff [21]
NH2 +HO2 =H2 NO+OH Sumathi and Peyerimhoff [21]
NH2 +HO2 =HNO+H2 O Sumathi and Peyerimhoff [21]
NH2 +HO2 =HON+H2 O Sumathi and Peyerimhoff [21]
N2 H2 +M=NNH+H+M Mei et al. [2] x 0.75
N2 H2 +H=NNH+H2 Zheng et al. [22] x 2
NH2 +NH=N2 H2 +H Klippenstein et al. [23] x 2
NH2 +NH2 =N2 H2 +H2 Klippenstein et al. [23] x 2
NH+OH=HNO+H Klippenstein et al. [23]
NH+OH=N+H2 O Klippenstein et al. [23]
H2 NO+O2 =HNO+HO2 Song et al. [3] x 0.78

equivalence ratio (φ). Literature data, own mea- tions NNH+O2 =N2 +HO2 , NH2 +O=HNO+H
surements and calculations are shown. Good agree- and NH+OH=HNO+H are important for pre-
ment can be observed among the measured lam- dicting laminar flame speed, with increasing
inar flame speeds for fuel-lean and stoichiometric importance with increasing H2 content in the
conditions while discrepancies can be observed on blend.
the fuel-rich side. The model (from this work) cap- Figure 3 shows measured laminar flames speeds
tures the experimental trends from lean to stoichio- (symbols) compared against the model predictions
metric conditions very well. For lean and stoichio- using our previous work [4] and the model improve-
metric conditions, the model agrees with most mea- ments from this work for NH3 /H2 /air at 473 K and
surements. For fuel-rich conditions, the best agree- 1 bar [11] and 3 bar (present work). Simulations
ment is found with the older measurements from were performed using a detailed multicomponent
Ronney [31] and with the very recent experiments diffusion model to consider the strong thermo dif-
from Mei et al. [2]. Furthermore, particularly for fusion of H and H2 . It can be seen that as the hy-
fuel-rich conditions the model in this work shows drogen content in the mixture is increased lami-
an improved agreement against the measurements, nar flame speed is also increased. The experiments
compared to the model from our previous work show peak laminar flame speed at φ = 1.1, which
[4]. This improvement is mainly due to the update is well predicted by the models. It can be observed
in the rate parameters for the reactions involving that below 10% H2 the laminar flame speed does
N2 H2 (listed in Table 1). Reaction flow analysis not increase significantly, but does for H2 contents
(see Fig. 2, only major paths are shown here) re- in the blend above 10% (see Fig. 1b). For 1 bar
veals that reaction paths involving N2 H2 are fa- (Fig. 3a) at φ = 1.1 with 10% and 30% H2 , lami-
vored more at rich conditions. As can be seen in nar flame speed is increased by factor 1.4 and 2.8
Fig. 2 at rich environment (φ=1.4 – red numbers) respectively and for 3 bar (Fig. 3b) at same condi-
most of the NH2 reacts to N2 H2 reacting with NH tions by factor 1.3 and 2.0 respectively compared
or self-recombining. This route increases its flux by to neat ammonia flame. This leads to the conclu-
60% when compared to stoichiometric conditions. sion that the impact of H2 on the laminar flame
This increased importance of N2 H2 formation via speed of NH3 /air is reduced with increasing pres-
the NH2 route is compensated by the shrink in the sure. Fig. 3 also demonstrates that the model pre-
NH formation route. dictions from this work agree well with most mea-
Figure 1b shows the laminar flame speed of surements with a slight underestimation for 30% H2
NH3 /H2 /air blends at stoichiometric and standard (Fig. 3a) under lean to stoichiometric conditions.
condition (1 atm, 298 K) at varying H2 content (0– Figure 4 shows new laminar flame speed data
100%) in the fuel blend. Most measurements limit of NH3 /H2 /air at φ = 1.1, 473 K, and pressure of 5,
the hydrogen content to 60%. The measurements 7 and 10 bar respectively as a function of H2 con-
from Kumar and Meyer [32] at 60% H2 and above tent in the fuel blend. It can be observed that the
diverge from other experimental measurements. It model predicts slightly higher flame speeds com-
can be observed that the updated model from this pared to measurements but remain within the error
work (solid line) accurately predicts the laminar margins of experiments. There are limited number
flame speed. Calculations with our mechanism of experimental data in literature at higher pressure
from prior work [4] is also shown (dashed line). (we believe this study is first to report data at 7 and
It is found that for NH3 /H2 /air blends the reac- 10 bar).
K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174 2167

Fig. 1. Laminar flame speed at 1 atm and 298 K for NH3 /air (a) and for stoichiometric conditions for NH3 /H2 /air (b).
Symbols: measurements published literature [2,5,10,11,25,27–33]. Dashed lines: previous model [4], solid lines: this work.

Fig. 2. Integrated N-atom mass flux analysis at 1 atm and 298 K for φ = 1.0 and φ = 1.4 NH3 /air blends. Numbers are the
percentage of the reactant forming the respective product. Major changes in flow highlighted by red for φ = 1.4.
2168 K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174

Fig. 3. Laminar flame speeds of NH3 /H2 /air at 473 K and 1 bar (a) and 3 bar (b) at varying H2 content. Symbols: mea-
surements from [11] and this work. Dashed lines: previous model [4], solid lines: this work.

Fig. 4. Laminar flame speed of NH3 /H2 /air at 473 K and varying H2 content at φ = 1.1 and 5–10 bar. Symbols: present
measurements. Lines: simulations from this work.
K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174 2169

Fig. 5. Reaction sensitivity analysis for laminar flame speed of NH3 /H2 /air at φ = 1.0, 1 bar and 473 K with varying H2
content in the fuel blend.

To outline important reactions with regard to Furthermore, for a comparison study simula-
NH3 /H2 /air laminar flame speeds a reaction sen- tions have also been performed using three differ-
sitivity analysis is conducted which is presented ent mechanisms from literature (Glarborg et al.
in Fig. 5 for φ = 1.0 at 1 bar and 473 K and dif- [34], Otomo et al. [35] and Stagni et al. [36]).
ferent H2 contents (0–30%) in the fuel blend. The performance of these mechanisms against
In Fig. 5 positive sensitivity means reaction pro- the experimental data (shown in Figs. 1, 3 and
motes the reactivity (increased flame speed) and 4) are provided in SI (Fig. S12-S15). Among
negative sensitivity means reaction retards the re- these, the model from Glarborg [34] overesti-
activity. The chain branching reaction from the mates the laminar flame speed at all condi-
hydrogen chemistry (O2 +H=OH+O) is also for tions, Otomo’s [35] underestimates mainly for
ammonia the most sensitive reaction under all con- NH3 /H2 blends, and the Stagni model [36]
ditions. Its sensitivity increases as the H2 content performance is similar as our mechanism from this
is increased and explains the accelerating lami- work.
nar flame speed with increasing H2 content. Other
most sensitive reactions involve the NNH radical, 4.2. Laminar flame speed of NH3 with an enriched
which is a key radical in the ammonia oxidation oxygen content
chemistry. The thermal dissociation NNH=N2 +H
competes with the oxidation NNH+O2 =N2 +HO2 . Figure 6 shows laminar flame-speed compar-
The former reaction promotes the reactivity by isons between model predictions and new measure-
producing more reactive H-atoms while the latter ments for NH3 /O2 /N2 with varying O2 content (21–
retards the system reactivity by producing less re- 30%) at 298 K (a), 323 K (b) and 373 K (c) as a func-
active HO2 radicals. Sensitivity increases with in- tion of φ. The experiments show that increasing the
creasing H2 content in the blend for both reac- O2 content increases the laminar flame speed and
tions. The reactions, NH2 +NO=NNH+OH and the same trend is predicted by the model. The sen-
N2 H2 +M=NNH+H+M are chain branching re- sitivity on O2 is slightly overestimated, which is seen
actions promoting the reactivity of the system. under lean conditions. The model follows the exper-
With respect to the increasing H2 content, the sen- imental trend very well for all the conditions. For
sitivity of the former reaction increases while sen- completeness it is also stated that the agreement is
sitivity of later reaction decreases. A further im- better for 298 and 373 K than at 323 K. The lami-
portant reaction is NH2 +NH=N2 H2 +H, with the nar flame speed depends almost linearly on the O2
highest contribution to the formation of N2 H2 content in the oxidizer (see also Fig. S17 in SI). The
and consuming the important key radicals (NH2 influence of O2 is mainly explained by the increased
and NH). The sensitivity of this reaction also in- adiabatic flame temperature. The influence of the
creases with increase in H2 content. Similar ob- chemistry is discussed in Fig. 7. It should be noted
servations can be drawn for other reactions as that with increasing H2 in fuel blend the growth is
well. Additional model validation of NH3 /air and exponential while this does not hold true for O2 (see
NH3 /H2 /air laminar flame speeds for wide range Figs. 1b and S17).
of experimental condition is shown in SI (see To explore the important reactions for predict-
Fig. S1 – S8). ing the laminar flame speeds at varying O2 content
2170 K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174

Fig. 6. Laminar flame speed of NH3 /O2 /N2 at 1 bar and different temperatures with varying O2 content (21 – 30%). At
298 K (a), 323 K (b) and 373 K (c). Symbols: measurements from this work. Lines: model prediction from this work.

a reaction sensitivity analysis is performed. blend (Fig. 5), shows a decreasing sensitivity
Figure 7 displays the 15 most sensitive reactions at trend with an increase in O2 content. Simi-
φ = 1.0, 1 bar and 298 K. The highest sensitivity of larly, reaction N2 H2 +M=NNH+H+M and
the branching reaction O2 +H=OH+O is expected. N2 H2 +H=NNH+H2 show an increasing trend
Its sensitivity increases with increasing O2 content in sensitivity contrary to Fig. 5. However, other
in the oxidizer. Most of the sensitive reactions important reactions NH2 +NO=NNH+OH and
are unchanged when compared to Fig. 5 (H2 NH2 +NH=N2 H2 +H show a similar trend de-
addition to the fuel). Reaction NNH=N2 +H creasing and increasing. Hence the chemical effect
and NNH+O2 =N2 +HO2 which showed of O2 enrichment is more complex. It is, how-
an increased sensitivity trend with an in- ever governed by the chain branching reaction
crease in H2 content in the NH3 /H2 /air O2 +H=OH+O.
K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174 2171

Fig. 7. Reaction sensitivity analysis for laminar flame speed of NH3 /O2 /N2 at φ = 1.0, 1 bar and 298 K with varying O2
content (21 – 30%).

Fig. 8. Laminar flame speed of NH3 /O2 /He at 1 bar and 373 K and different O2 contents. Symbols: measurements. Lines:
simulation from this work.

The prediction of different models from litera- slightly higher laminar flame speed at lean and
ture [34–36] was compared against the present ex- stoichiometric conditions and shows much better
perimental data (see SI Fig. S22 – Fig. S24). The agreement on the rich side.
model from Glarborg [34] overestimates the lami-
nar flame speed at all conditions, Otomo’s [35] un- 4.3. Model validation against literature data for
derestimates mainly on the rich side, and Stagni IDT, speciation in BSF and JSR
model [36] performance is similar as our model
from this work. Finally, we demonstrate for completeness of this
Figure 8 shows the laminar flame speeds of paper that the modified model still predicts the val-
NH3 /O2 /He blend compared to the model predic- idation experiments from previous work. Figure 9a
tion for 27 and 30% O2 at 373 K and model pre- shows the comparison between the model predic-
dictions for 21% O2 . This experiment varies the tions against selected experimental data for igni-
heat capacity on the one hand, and the collision tion delay times in shock tube [1] and RCM [9] for
efficiency on the other hand of the inert gas in NH3 /O2 /Ar blend. The model predictions are in
the oxidizer. The strong increase in laminar flame good agreement with the experimental data for
speed compared to experiments with N2 /O2 oxi- a wide range of pressure over the whole temper-
dizer blends is mainly explained by the increased ature range investigated. Fig. 9 (b: speciation in
adiabatic flame temperature. The model predicts NH3 /NO/Ar, BSF (295 K, 7.2 kPa), (c, d): specia-
2172
K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174
Fig. 9. Comparison between model prediction and experimental data from literature for NH3 /O2 /Ar, IDT (in ST and RCM) (a); for NH3 /NO/Ar (φ = 1.46), BSF (b); and for
H2 /O2 /NO (220 ppm)/N2 (φ = 0.5, 10 atm, τ = 1.0 s), JSR (c, d). Symbols: measurements [1,37,38], lines: this work.
K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174 2173

tion during H2 /O2 /N2 oxidation in JSR doped with References


220 ppm of NO) compares the model prediction
against the experimental data from the literature [1] O. Mathieu, E.L. Petersen, Combust. Flame 162
and good agreement is found. Additional model (2015) 554–570, doi:10.1016/j.combustflame.2014.
validation for IDT, BSF, and JSR is shown in SI 08.022.
[2] B. Mei, X. Zhang, S. Ma, M. Cui, H. Guo, Z. Cao,
to offer a broad data base (Fig. S25 – S31).
Y. Li, Combust. Flame 210 (2019) 236–246, doi:10.
1016/j.combustflame.2019.08.033.
[3] Y. Song, H. Hashemi, J.M. Christensen, C. Zou,
P. Marshall, P. Glarborg, Fuel 181 (2016) 358–365,
5. Conclusion doi:10.1016/J.FUEL.2016.04.100.
[4] K.P. Shrestha, L. Seidel, T. Zeuch, F. Mauss, En-
New laminar flame speed database of NH3 in ergy & Fuels 32 (2018) 10202–10217, doi:10.1021/acs.
O2 -enriched air and NH3 /H2 blends have been energyfuels.8b01056.
studied experimentally at 1–10 bar, 298–473 K us- [5] A. Hayakawa, T. Goto, R. Mimoto, Y. Arakawa,
T. Kudo, H. Kobayashi, Fuel 159 (2015) 98–106,
ing a constant pressure spherical combustion ves-
doi:10.1016/j.fuel.2015.06.070.
sel. Based on the present measurements and pub- [6] H. Xiao, A. Valera-Medina, J. Eng. Gas Turbines
lished experimental data, a reaction mechanism for Power 139 (2017) 081504, doi:10.1115/1.4035911.
the oxidation of NH3 and NH3 /H2 blends is de- [7] J. Li, H. Huang, N. Kobayashi, Z. He, Y. Nagai, Int.
veloped. A number of published experiments have J. Energy Res. 38 (2014) 1214–1223, doi:10.1002/er.
been selected to demonstrate that important fea- 3141.
tures of the ammonia chemistry are well captured [8] M. Pochet, V. Dias, B. Moreau, F. Foucher, H. Jean-
by the mechanism that was modified to include mart, F. Contino, Proc. Combust. Inst. 37 (2019) 621–
laminar flame speed prediction. Overall, there is a 629, doi:10.1016/J.PROCI.2018.05.138.
[9] X. He, B. Shu, D. Nascimento, K. Moshammer,
good agreement between model predictions and ex-
M. Costa, R.X. Fernandes, Combust. Flame 206
perimental measurements for the wide range of ex- (2019) 189–200, doi:10.1016/j.combustflame.2019.
perimental conditions investigated. It is found that 04.050.
laminar flame speed of neat ammonia can be in- [10] X. Han, Z. Wang, M. Costa, Z. Sun, Y. He, K. Cen,
creased by increasing the oxygen content or using Combust. Flame 206 (2019) 214–226, doi:10.1016/j.
hydrogen as fuel blend. To rate the sensitivity of H2 combustflame.2019.05.003.
addition to the fuel blend or O2 enrichment in the [11] C. Lhuillier, P. Brequigny, N. Lamoureux, F. Con-
oxidizer we compared the data from Fig. 1b and tino, C. Mounaïm-Rousselle, Fuel (2020) 263, doi:10.
Fig. 6a at 1 atm, 298K and φ = 1.0: for 30% H2 (in 1016/j.fuel.2019.116653.
[12] P. Brequigny, G. Dayma, F. Halter, C. Mounaïm-
fuel stream) or 30% O2 (in oxidizer stream) the lam-
Rousselle, T. Dubois, P. Dagaut, Proc. Combust. Inst.
inar flame speed of NH3 is increased by factor 2.9 35 (2015) 703–710, doi:10.1016/j.proci.2014.06.126.
or 2.8 respectively. Hence, 9% increase of O2 (com- [13] B. Galmiche, F. Halter, F. Foucher, Combust. Flame
pared to air) content has the same effect as 30% H2 159 (2012) 3286–3299, doi:10.1016/j.combustflame.
in the fuel blend. 2012.06.008.
[14] A.P. Kelley, C.K. Law, Combust. Flame 156 (2009)
1844–1851, doi:10.1016/j.combustflame.2009.04.
004.
Declaration of Competing Interest [15] F. Halter, T. Tahtouh, C. Mounaïm-Rousselle, Com-
bust. Flame 157 (2010) 1825–1832, doi:10.1016/j.
combustflame.2010.05.013.
There are competing interests. [16] H. Yu, W. Han, J. Santner, X. Gou, C.H. Sohn,
Y. Ju, Z. Chen, Combust. Flame 161 (2014) 2815–
2824, doi:10.1016/j.combustflame.2014.05.012.
[17] G. Altinay, R.G. Macdonald, J. Phys. Chem. A. 119
Acknowledgements (2015) 7593–7610, doi:10.1021/acs.jpca.5b00917.
[18] T.L. Nguyen, J.F. Stanton, Int. J. Chem. Kinet. 51
The research leading to these results has re- (2019) 321–328, doi:10.1002/kin.21255.
ceived funding from the French Government’s [19] V. Samu, T. Varga, I. Rahinov, S. Cheskis, T. Turányi,
“Investissement d’Avenir” program: “Laboratoire Fuel 212 (2018) 679–683, doi:10.1016/J.FUEL.2017.
10.019.
d’Excellence CAPRYSSES” (Grant No ANR-11- [20] J.A. Miller, M.D. Smooke, R.M. Green, R.J. Kee,
LABX-0006–01). Combust. Sci. Technol. 34 (1983) 149–176, doi:10.
1080/00102208308923691.
[21] R. Sumathi, S.D. Peyerimhoff, Chem. Phys. Lett.
263 (1996) 742–748, doi:10.1016/S0009-2614(96)
Supplementary materials 01283-3.
[22] J. Zheng, R.J. Rocha, M. Pelegrini, L.F.A. Fer-
Supplementary material associated with this ar- rão, E.F.V. Carvalho, O. Roberto-Neto,
ticle can be found, in the online version, at doi:10. F.B.C. MacHado, D.G. Truhlar, J. Chem. Phys.
136 (2012) 184310, doi:10.1063/1.4707734.
1016/j.proci.2020.06.197.
2174 K.P. Shrestha et al. / Proceedings of the Combustion Institute 38 (2021) 2163–2174

[23] S.J. Klippenstein, L.B. Harding, B. Ruscic, R. Sivara- [31] P.D. Ronney, Combust. Sci. Technol. 59 (1988) 123–
makrishnan, N.K. Srinivasan, M. Su, J.V. Michael, 141, doi:10.1080/00102208808947092.
J. Phys. Chem. A. 113 (2009) 10241–10259, doi:10. [32] P. Kumar, T.R. Meyer, Fuel 108 (2013) 166–176,
1021/jp905454k. doi:10.1016/j.fuel.2012.06.103.
[24] http://logesoft.com/loge-software/ [33] Y. Li, M. Bi, B. Li, W. Gao, Combust. Sci. Technol.
[25] K. Takizawa, A. Takahashi, K. Tokuhashi, 190 (2018) 1804–1816, doi:10.1080/00102202.2018.
S. Kondo, A. Sekiya, J. Hazard. Mater. 155 (2008) 1473859.
144–152, doi:10.1016/j.jhazmat.2007.11.089. [34] P. Glarborg, J.A. Miller, B. Ruscic, S.J. Klippenstein,
[26] J.H. Lee, S.I. Lee, O.C. Kwon, Int. J. Hydrog. Energy Prog. Energy Combust. Sci. 67 (2018) 31–68, doi:10.
35 (2010) 11332–11341, doi:10.1016/j.ijhydene.2010. 1016/j.pecs.2018.01.002.
07.104. [35] J. Otomo, M. Koshi, T. Mitsumori, H. Iwasaki,
[27] A. Ichikawa, A. Hayakawa, Y. Kitagawa, K. Yamada, Int. J. Hydrogen Energy 43 (2018) 3004–
K.D. Kunkuma Amila Somarathne, T. Kudo, 3014, doi:10.1016/j.ijhydene.2017.12.066.
H. Kobayashi, Int. J. Hydrog. Energy 40 (2015) [36] A. Stagni, C. Cavallotti, S. Arunthanayothin,
9570–9578, doi:10.1016/j.ijhydene.2015.04.024. Y. Song, O. Herbinet, F. Battin-Leclerc, T. Far-
[28] T. Jabbour, D.F. Clodic, ASHRAE Trans 110 (2004) avelli, React. Chem. Eng. 5 (2020) 696–711,
522–533. doi:10.1039/c9re00429g.
[29] V.F. Zakaznov, L.A. Kursheva, Z.I. Fedina, Com- [37] J. Vandooren, J. Bian, P.J. Van Tiggelen, Combust.
bust. Explos. Shock Waves 14 (1978) 710–713, doi:10. Flame 98 (1994) 402–410, doi:10.1016/0010-2180(94)
1007/BF00786097. 90178-3.
[30] U.J. Pfahl, M.C. Ross, J.E. Shepherd, [38] G. Dayma, P. Dagaut, Combust. Sci. Technol. 178
K.O. Pasamehmetoglu, C. Unal, Combust. Flame (2006) 1999–2024, doi:10.1080/00102200600793171.
123 (2000) 140–158, doi:10.1016/S0010-2180(00)
00152-8.

You might also like