Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Fuel 343 (2023) 127991

Contents lists available at ScienceDirect

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

Full Length Article

Experimental and numerical study on the ignition delay times of methane/


propane mixtures diluted in carbon dioxide
Wenxiang Xia , Chun Zou *, Yi Yuan , Rui Fu , Jiacheng Liu , Chao Peng
State Key Laboratory of Coal Combustion, Huazhong University of Science and Technology, Wuhan 430074, PR China

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

Keywords: Pressurized oxy-fuel combustion has gained great interest owing to its low emission and high efficiency. Methane
Oxy-fuel combustion and propane are important composition of natural gas, which is a clean fossil fuel and widely utilized for
CH4/C3H8 mixture combustion applications including gas turbines. In this work, the auto-ignition properties of CH4/C3H8 mixtures
Ignition delay time
in ratios from 90/10% to 70/30% diluted in CO2 were measured at pressures of 1.75 and 10 atm; equivalence
Chemical kinetics
ratios of 0.5, 1.0, and 2.0; and temperatures in the range of 1240–1520 K in a shock tube. Four key reactions in
the Oxymech 2.0 model were updated and the modified model shows reasonable performance in both the
ignition delay time and the laminar flame speed predictions in O2/N2 and O2/CO2 atmospheres. Two kinetic
models (NUIG 1.0 and AramcoMech 3.0) were evaluated by present data and the laminar flame speed from the
literature. The results of the experiment and calculation show that the addition of propane promotes the ignition
significantly in the C3H8 proportion range from 0 to 30% in both cases of CO2 and N2, but has slight effects on the
ignition with the C3H8 proportion more than 30%. The effects of CO2 on the promoting effects of C3H8 addition
are slight. The ignition delay times increase with the equivalence ratio at 1.75 atm but no statistical difference is
observed between the measured data of different equivalence ratios at 10 atm. The influence of C3H8 addition
and the equivalence ratio on the ignition delay of CH4/C3H8 mixtures was also analyzed in detail.

1. Introduction Ar or N2.
Petersen et al. [10] measured the IDTs of fuel mixtures consisting of
The emission of carbon dioxide from the combustion of fossil fuels, as CH4/C3H8 in ratios ranging from 90%/10% to 60%/40% at gas turbine
a cause of global warming, has gained great interest in recent years pressures by shock tube and found that increasing the blend ratios of
[1–3]. Oxy-fuel combustion, in which the combustion occurs in the O2/ propane from 10% to 30% can decrease the IDTs. They also proposed a
CO2 atmosphere, is a promising carbon capture technology due to the CH4/C3H8 oxidation mechanism to simulate the experimental results
high CO2 concentration product gas for easier carbon capture [4,5]. The and mentioned that the reactions involving CH3O, CH3O2, and CH3 +
direct semi-closed oxy-combustion cycle using natural gas, the Allam O2/HO2 were important in reproducing the correct kinetic behavior.
cycle, is capable of capturing CO2 with a net efficiency up to 59% LHV Then Healy et al. [11] extended the experiments to intermediate and low
[6,7]. Therefore, the pressurized oxy-fuel combustion of natural gas has temperature conditions using a rapid compression machine and updated
attracted much attention. the kinetic model of CH4/C3H8 [10]. The model reproduced correctly
Natural gas is primarily composed of methane with some heavier the ignition at conditions of various equivalence ratios and pressures.
alkanes ranging from ethane to heptane depending on its origin, Recently, El-Sabor Mohamed et al. [18] measured the IDTs of natural gas
extraction, and transport processes [8]. Although the fractional contri­ mixtures comprising C1-C7 n-alkanes in air using a rapid compression
bution of propane to the composition of conventional natural gas is machine at gas turbine operating conditions and developed a compre­
small, the contribution of propane to the combustion of natural gas hensive chemical kinetic model (NUIG 1.0), which can successfully
cannot be ignored. There are various experimental and kinetic modeling predict the experiments. Moreover, Lowry et al. [14] measured the LFSs
studies for ignition delay time (IDT) [9–13] and laminar flame speed of 80 %CH4/20 %C3H8 mixtures in air at 1 and 5 atm using a constant-
(LFS) [14–17] of a wide range of methane/propane mixtures diluted in volume vessel. Then Nilsson et al. [15] investigated the effects of

* Corresponding author.
E-mail address: zouchun@hust.edu.cn (C. Zou).

https://doi.org/10.1016/j.fuel.2023.127991
Received 27 July 2022; Received in revised form 12 February 2023; Accepted 24 February 2023
Available online 3 March 2023
0016-2361/© 2023 Elsevier Ltd. All rights reserved.
W. Xia et al. Fuel 343 (2023) 127991

hydrogen addition on the LFSs of 80 %CH4/20 %C3H8 mixtures and As shown in Fig. 1, the IDT is defined as the time interval between the
pointed out that the Aramco 1.3 model [19] slightly over-predicted the arrival of the reflected shocks and the onset of ignition, which is
experimental results. determined by extrapolating the maximum slope of the normalized OH*
For the oxy-fuel combustion atmospheres, Hargis and Petersen [20] emission linearly back to the zero baseline. Moreover, the shock bifur­
measured the IDTs of CH4 diluted in 75 %CO2 in shock tubes and pointed cation due to the large addition of CO2 could result in the difficult
out that CH3 + H (+M) ⇔ CH4 (+M) and HCO + M ⇔ H + CO + M were determination of the main reflected wave arriving. Laich et al. [24] and
significant third-body reactions for the CH4 ignition at high tempera­ Ninnemann et al. [25] have studied the shock bifurcation under CO2
ture, while 2CH3 (+M) ⇔ C2H6 (+M) was significant at low temperature. dilution in detail. In this study, the arrival time of the reflected shock, tA ,
In our previous study [21–23], the IDTs of CH4, C2H6, and C3H8 diluted was determined using the empirical formula proposed by Petersen and
in CO2 were measured and a kinetic model named Oxymech 2.0 was Hanson [26]:
developed, which was validated by experimental IDTs and LFSs for C1-
D
C3 alkanes in oxy-fuel and air combustion over a wide range of pressure. tA = ti + + ΔtAO (1)
2VR
To our knowledge, the IDTs of CH4/C3H8 mixtures in oxy-fuel com­
bustion are still lacking, and Oxymech 2.0 also needs validation by the where ti denotes the time when the initial pressure rises,VR denotes the
IDTs of CH4/C3H8 mixtures. velocity of the reflected shock, and D denotes the diameter of the
CO2 is not inert in oxy-fuel combustion, and the reaction rates of pressure transducer (Kistler 603B1). The ΔtAO can be expressed as:
third-body reactions in O2/CO2 atmospheres are enhanced due to the
chaperon effects of CO2. Liu et al. [21] found that the chemical effects of (2)
0.57 − 7.1 0.66
ΔtAO (μs) = 4.6M γ 2 MS
CO2 suppress the ignition at lower temperature mainly due to the
chaperon effects of 2CH3 (+M) ⇔ C2H6 (+M) while promoting the where M denotes the molecular weight of the mixture, γ2 denotes the
ignition of CH4 at higher temperature mainly due to the chaperon effects specific heat ratio upstream of the bifurcation, and Ms denotes the Mach
of HCO + M ⇔ H + CO + M at 0.8 and 1.75 atm. The kinetic analysis by number of the incident shocks. The temperature (T5) and pressure (P5)
Xia et al. [23] indicated that C3H8 (+M) ⇔ C2H5 + CH3 (+M) plays an behind the reflected shock waves were calculated by the chemical
important role in third-body reactions for the C3H8 ignition. Obviously, equilibrium program, Gaseq [27]. The uncertainty in T5 was estimated
the influence of CO2 on CH4/C3H8 ignition needs further investigation. within ± 24 K, and the total uncertainty in the IDT measurements was
In this study, the IDTs of CH4/C3H8 mixtures with ratios ranging within 20%. A detailed analysis of the uncertainty can be obtained in the
from 90%/10% to 70%/30% diluted in CO2 were measured at pressures Supplementary Material.
of 1.75 and 10 atm; equivalence ratios of 0.5, 1.0, and 2.0; and tem­
peratures in the range of 1240–1520 K in a shock tube. The modified 2.2. Experimental facilities and methods validation
Oxymech 2.0 model by updating four key reactions was validated by the
IDTs and LFSs of CH4/C3H8 mixtures in both oxy-fuel and air combus­ In our previous work [21], the IDTs of CH4 in O2/CO2 atmosphere
tion. Two kinetic models (NUIG 1.0, and AramcoMech 3.0) were eval­ were measured and compared with those from Hargis and Petersen [20]
uated using the experimental IDTs and the LFSs from the literature. The at the equivalence ratio of 0.5 and the pressure of 1.75 atm and our data
effects of the CH4/C3H8 ratio and the equivalence ratio on the IDTs of conform fairly well to the results of Hargis and Peterson. In this study,
CH4/C3H8 mixtures were also analyzed in detail. the IDTs of the CH4/C3H8 mixture in O2/N2 atmosphere were measured
and compared with those from the Petersen et al. [10] at the equivalence
2. Experiments ratio of 0.5 and the pressure of approximately 10 atm As shown in Fig. 2,

2.1. Facilities and methods

In this study, the ignition delay time measurements of CH4/C3H8


mixtures under O2/CO2 atmospheres were conducted in a shock tube,
with 4 m driver section, 8 m driven section, and an inner diameter of 0.1
m, as reported in our previous work in detail [21]. The tested conditions
are listed in Table 1. The purities of two fuels were 99.99% while the
purities of oxygen and dilution gases were all 99.999%. All the mixtures
required resting in a 300 L mixing tank for more than 12 h while pre­
paring. The mixing tank and the shock tube were evacuated to less than
0.1 Pa before operation. Five piezoelectric pressure transducers (PCB
111A24) were placed alongside the shock tube at intervals of 20 cm to
record the incident shock velocity. The pressure trace was captured
using a piezoelectric pressure transducer (Kistler 603B1) placed 2 cm
upstream the endwall, while a photomultiplier with a band pass filter of
307 ± 10 nm was placed 2 cm upstream the endwall to monitor the OH*
Fig. 1. Typical signals of the ignition delay time measurement for mix-1 at T =
emission. 1361 K and P = 10.2 atm.

Table 1
Composition of the experimental mixtures and test conditions.
Mixture CH4 C3H8 O2 CO2 N2 CH4/C3H8 Φ Pressure (atm)

Mix-1 3.93% 0.44% 20.09% 75.54% 0 9/1 0.5 1.75, 10


Mix-2 3.11% 0.78% 20.19% 75.92% 0 8/2 0.5 1.75, 10
Mix-3 2.45% 1.05% 20.27% 76.23% 0 7/3 0.5 1.75, 10
Mix-4 4.73% 2.03% 19.59% 73.66% 0 7/3 1.0 1.75, 10
Mix-5 8.86% 3.80% 18.35% 69.00% 0 7/3 2.0 1.75, 10
Mix-6 3.93% 0.44% 20.09% 0 75.54% 9/1 0.5 10

2
W. Xia et al. Fuel 343 (2023) 127991

hydrocarbon kinetic model and is widely utilized. These kinetic models


are evaluated by the experimental data in this study. The closed ho­
mogeneous batch reactor in the CHEMKIN-PRO software [33] was
adopted for the numerical calculation of IDTs. A pressure rise about
16%/ms owing to non-ideal effects appeared from 1100 μs behind the
reflected shock wave, as shown Fig. S1 in the supplemental material.
Considering the relatively short IDT, the pressure rise resulting from
non-ideal effect was considered negligible during simulation [34,35],
thus the constant U-V assumption was adopted without modification.
The calculated IDT is defined as the time interval between zero and the
moment of the maximum rate of temperature rise (max dT/dt).

4. Results and discussion

4.1. Model evaluation

Fig. 2. Comparison between the experimental data in this work and those from Fig. 3 displays the measured IDT data for different CH4/C3H8 mix­
Petersen et al. [10]. tures diluted in CO2 with different equivalence ratios at 1.75 atm and 10
atm and the simulated IDTs using the unmodified and modified Oxy­
the measured data of the present work are consistent with the literature mech 2.0, NUIG 1.0, and AramcoMech 3.0. As shown in Fig. 3, NUIG 1.0
data. These confirm the reliability of present facilities and methods for and AramcoMech 3.0 over-predict nearly all of the IDTs. Oxymech 2.0
the IDT measurement. The measured data are provided in the supple­ and the modified model show reasonable performance in most of the IDT
mental material. predictions. The absolute relative error (ARE) was introduced to eval­
uate the four models quantitatively, and defined as follows:
3. Modeling ⃒ ⃒
⃒τsim − τexp ⃒
ARE = ⃒⃒ ⃒ × 100%
⃒ (3)
The Oxymech 2.0 model proposed in our previous study [23] per­ τexp
forms good agreement with the IDT data of CH4/C3H8 mixtures but over-
where τsim and τexp are the simulated and measured IDTs, respectively.
predicted most of the LFS data in fuel-rich conditions, as discussed in
The boxplots are utilized to report and compare the ARE data distribu­
next section. To improve the performance of Oxymech 2.0 on LFS pre­
tions of the four models, as displayed in Fig. 3(f). The average and the
dictions, four reactions were updated, which show high sensitivity to
median ARE values (marked by the rectangular symbol and the central
LFS but not sensitive to IDT, as listed in Table 2. Baulch et al. [28]
line in the box, respectively) of NUIG 1.0 and AramcoMech 3.0 are much
provided the preferred values of the kinetic data of CH4 + H ⇔ CH3 + H2
larger than those of Oxymech 2.0 and the modified model. Moreover,
(R43) based on the previous investigations. Jasper et al. [29] calculated
the distribution of the ARE data for Oxymech 2.0 and the modified
the rate coefficients of the pressure-dependent reaction CH3 + OH ⇔
model is more concentrated owing to the smaller box length. The eval­
CH2OH + H (R94) and only the low pressure limit chemical activation
uation of the models using the literature IDT data of different CH4/C3H8
rate constant expression was adopted in Oxymech 2.0. It is necessary to
mixtures diluted in N2 is shown in the Supplementary Material and the
apply the completed rate coefficients for different pressures. The rate
modified model also shows good performance.
constants of C2H4 + H (+M) ⇔ C2H5 (+M) (R207) in AramcoMech 3.0
Fig. 4(a) displays the experimental LFS data from Nilsson et al. [15]
sourced from Miller and Klippenstein [30] were also adopted. The re­
and the simulated results for 80 %CH4/20 %C3H8 mixtures at 1 atm and
action C2H3 + H ⇔ C2H2 + H2 (R300) can affect the LFSs of methane/
298 K and Fig. 4(b) displays the boxplots of the ARE data distributions of
propane mixtures significantly and was also updated using the results
the four models. More simulated results for CH4/C3H8 mixtures diluted
adopted in AramcoMech 3.0 [31].
in N2 are shown in the Supplementary Material. It is observed that the
Moreover, the NUIG 1.0 model [18] was developed with updated
performance of the modified model on LFS predictions is much better
reaction rates based on recent theoretical and experimental studies
than the Oxymech 2.0 model. Moreover, the LFS simulations using
together with the heptane mechanism developed by Zhang et al. [32]
Modified Oxymech 2.0 for CH4 diluted in CO2 are also shown in the
and showed excellent agreement with the IDT measurements for C1-C7
Supplementary Material, in which the model performs good agreement
n-alkanes. The AramcoMech 3.0 [31] model is a representative
with the experimental LFS data. Based on the results of IDTs and LFSs,
the modified model is utilized for the subsequent kinetic analysis.
Table 2
Key elementary reactions updated in the present studya.
Number Reaction A b Eα Source
4.2. Effects of C3H8 addition and O2/CO2 atmosphere

R43 CH4 + H ⇔ CH3 + H2 6.14E + 2.5 9587 [28]


The addition of propane can promote the ignition process, thus
05
R94 CH3 + OH ⇔ CH2OH + H 5.60E + 0.73966 3971 [29] causing shorter IDTs, as shown in Fig. 5. The IDTs decrease significantly
10 with the proportion of C3H8 changing from 0 to 10% but decrease
(Pressure-dependent) slightly with the C3H8 proportion from 10 to 30% in O2/CO2 atmo­
R207 C2H4 + H (+M) ⇔ C2H5 9.57E + 1.463 1355 [30]
spheres, similar to literature results in O2/N2 atmospheres. To investi­
(+M) 08
Low pressure limit 1.42E + − 6.642 5769
gate the effects of C3H8 addition quantitatively, the IDTs of CH4/C3H8
39 mixtures in different ratios need to be obtained in the same condition.
a = -0.569, T3 = 299, T1 = -9147, T2 = 152.4 However, it is difficult to measure the IDTs at the same pressure and
R300 C2H3 + H ⇔ C2H2 + H2 1.70E + 0 0 [31] temperature in a shock tube owing to the variant shock wave attenua­
14
tion. Therefore, the IDTs of CH4/C3H8 mixtures in different ratios were
a
Rate constants are expressed as k = AT b exp(− Ea /Ru T) with units of calories, calculated using Modified Oxymech 2.0 at the same pressure (10 atm)
mole, cm3 and s. and temperature (1400 K) to investigate the effects of C3H8 addition in
both O2/N2 and O2/CO2 atmospheres, as depicted in Fig. 6. It is noted

3
W. Xia et al. Fuel 343 (2023) 127991

Fig. 3. Comparison between experimental and simulated IDTs of CH4/C3H8 diluted in CO2 in this study.

Fig. 4. Performance of four models on LFS predictions of CH4/C3H8 mixtures.

4
W. Xia et al. Fuel 343 (2023) 127991

Fig. 5. Experimental and simulated IDTs of CH4/C3H8 mixtures with different propane additions.

10 atm, T = 1400 K, and Φ = 0.5 in O2/CO2 atmospheres. The brute


force sensitivity in O2/N2 atmospheres is provided in the Supplementary
Material. For the cases in O2/CO2 atmospheres, the decomposition re­
action of propane, C3H8 (+M) ⇔ C2H5 + CH3 (+M) (R327), ranks the 3rd
among the ignition-promoting reactions with the addition of propane
but is absent in the pure methane conditions, as depicted in Fig. 7. Fig. 8
displays the mole fractions of methyl radicals 10 μs after the beginning
for different C3H8 proportions at 1400 K and 10 atm. It is noted that the
concentrations of methyl radicals are negligible in the pure methane
condition but increase significantly with the propane mole fraction
increasing to 30%. With more propane addition, the concentrations of
methyl radicals increase not so significantly, as shown in Fig. 8, which is

Fig. 6. Effects of C3H8 addition on the IDTs of fuel-lean CH4/C3H8 mixtures.

that the addition of 10% propane in the mixtures can significantly


reduce the IDTs. When the proportion of propane exceeds a certain limit
(about 30%), the IDTs of the CH4/C3H8 mixtures are nearly equal to
those of pure propane, i.e. the effects of C3H8 addition are slight, as
expected. The promoting effects of C3H8 addition in O2/CO2 atmo­
spheres are comparable with those in O2/N2 atmospheres. In another
word, the influence of CO2 on the promoting effects of C3H8 addition is
slight.
To further investigate the effects of C3H8 addition, the brute force
sensitivity [36] analysis was introduced. Fig. 7 shows the brute force
sensitivity with different propane proportions in the conditions of P = Fig. 8. Simulated mole fraction of CH3 at different C3H8 proportions.

Fig. 7. Top 10 reactions influencing the ignition in O2/CO2 atmospheres at C3H8 proportion of: (a) 0%; (b) 10%.

5
W. Xia et al. Fuel 343 (2023) 127991

in line with the case of Fig. 6. According to the rate of production, in the 4.3. Effects of equivalence ratio
case with C3H8 addition, the fast increase in the methyl radicals in the
initial stage is through R327 owing to the easier breaking of C-C bonds in Fig. 10 displays the measured IDTs of 70 %CH4/30 %C3H8 mixtures
propane, which enhances CH3 + HO2 ⇔ OH + CH3O (R2646), conse­ and simulated results by Modified Oxymech 2.0 at the pressure of 1.75
quently promoting the ignition. Therefore, the addition of propane in atm and 10 atm with the equivalence ratios increasing from 0.5 to 2.0.
the CH4/C3H8 mixtures can significantly reduce the IDTs of CH4/C3H8 The simulated results by NUIG 1.0 are shown in Fig. S12 in the Sup­
mainly through R327. plementary Material. As shown in Fig. 10 (a), the IDTs increase with the
As depicted in Fig. 6, the IDTs in O2/CO2 atmospheres are longer equivalence ratio at 1.75 atm and both the two models can predict the
than those in O2/N2 atmospheres. The influence of CO2 on the ignition dependence of IDT on the equivalence ratio. The brute force sensitivity
consists of physical effects and chemical effects. The former is related to was calculated using the modified model for different equivalence ratios
a high heat capacity of CO2 with respect to N2, which implies a low at the pressures of 1.75 atm to investigate the effects of the equivalence
adiabatic temperature, whereas the latter is mainly related to CO2 to ratio. Fig. 11 displays the top 15 reactions with high sensitivity co­
directly participate in reactions (effects of reactions with CO2) or indi­ efficients at 1.75 atm and1400 K. As shown in Fig. 11, CH4 + H ⇔ CH3 +
rectly participate in third-body reactions with higher third-body effi­ H2 (R2676) makes the largest inhibiting influence on the ignition. With
ciencies than N2 (chaperon effects) [37]. To isolate these effects of the the equivalence ratios increasing from 0.5 to 2.0, the order of C3H8 + H
properties of CO2, two artificial species, A and B, were added to the ⇔ n-C3H7 + H2 (R348) rises from the 11th to 4th, while reaction C3H8 +
kinetic model. The thermochemical properties of species A are the same H ⇔ i-C3H7 + H2 (R332), also inhibiting the ignition, rises from the 16th
as those of CO2, and the chemical properties are inert. Species B has the to 9th (not shown in Fig. 11). It is also noted that the order of reaction
same thermochemical properties of CO2 and participates in the third- 2CH3 (+M) ⇔ C2H6 (+M) (R2653) rises from 5th to 2nd due to the
body reactions, but does not takes part in the other reaction. There­ increasing CH3 radicals with the equivalence ratios from 0.5 to 2.0.
fore, the difference in the IDTs between O2/A atmospheres and O2/N2 These indicate that the inhibition effects of the hydrogen-abstraction
atmospheres demonstrates the effects of the physical properties of CO2. reactions of fuels by H radicals and the recombination of CH3 radicals
Similarly, the difference between O2/B atmospheres and O2/A atmo­ were strengthened in fuel-rich conditions due to the higher fuel con­
spheres represents the chaperon effects of CO2, while the difference centration. The net reaction rates of the reactions above increase with
between O2/CO2 atmospheres and O2/B atmospheres represents the the equivalence ratio. Therefore, the IDTs of CH4/C3H8 mixtures diluted
effects of the reactions with CO2. Fig. 9 displays the percentages of the in CO2 increase with the equivalence ratio at the pressure of 1.75 atm.
effects of the physical and chemical properties of CO2 in the overall As shown in Fig. 10 (b), no statistical difference is observed between
effects on the ignition at 1400 K and 10 atm. The positive value repre­ the measured data of different equivalence ratios at 10 atm, which is
sents inhibiting effects. For the 90 %CH4/10 %C3H8 mixtures, the similar to the IDTs of C3H8 diluted in CO2 [23]. Modified Oxymech 2.0
physical effects of CO2 due to the higher volume-specific heat of CO2 and NUIG 1.0 can both reproduce the independence of IDT on the
compared to N2 are slightly larger than the sum of the chaperon effects equivalence ratio at 10 atm. Fig. 12 shows the brute force sensitivity for
and the effects of the reactions with CO2, and the chaperon effects are three equivalence ratios at 10 atm and 1300 K. As depicted in Fig. 12, H
very close to the effects of the reactions with CO2. Moreover, it is noted + O2 (+M) ⇔ HO2 (+M) (R2563), a pressure-dependent reaction, is
that the chaperon effects are comparable to the physical effects for the significantly enhanced at 10 atm. R2563 ranks 10th among the ignition-
pure methane case. According to Liu et al. [21], the inhibiting influence suppressing reactions in fuel-lean conditions, inhibiting the ignition,
of the chaperon effects mainly results from 2CH3 (+M) ⇔ C2H6 (+M) while it ranks 6th among the ignition-promoting reactions in stoichio­
(R2653) for the pure methane case. With the propane addition, the metric conditions, promoting the ignition due to the enhancement of
promoting effects of C3H8 (+M) ⇔ C2H5 + CH3 (+M) (R327) result in the CH3 + HO2 ⇔ OH + CH3O (R2646). Because the increase in the con­
significant decrease in the inhibiting influence of the chaperon effects. centration of O2 with the decreasing equivalence ratio from stoichio­
Therefore, for the CH4/C3H8 mixtures, the ratio of the chaperon effects metric to fuel-lean conditions promotes the ignition due to the
on the ignition is the lowest in the total effects of CO2. Compared to the hydrogen-abstraction reactions of fuels and the recombination of CH3
case of O2/N2 atmospheres, R327, which promotes the ignition, and radicals, which is comparable to the inhibition effect of R2563, the IDTs
R2653, which inhibits the ignition, are both enhanced in the case of O2/ of the CH4/C3H8 mixtures in fuel-lean conditions are very close to those
CO2 atmospheres, thus the positive and negative effects due to chaperon in stoichiometric conditions. From Fig. 12, R2563 ranks 4th among the
effects of CO2 approximately cancel each other out. Therefore, the ef­ ignition-promoting reactions in fuel-rich conditions, promoting the
fects of CO2 on the promoting effects of C3H8 addition are slight. ignition. Because the increase in the mole fraction of fuels with the
increasing equivalence ratio from stoichiometric to fuel-rich conditions
inhibits the ignition, which is comparable to the positive effect of R2563,
the IDTs of the CH4/C3H8 mixtures in fuel-rich conditions are very close
to those in stoichiometric conditions. Therefore, there is no statistical
difference observed between these measured data of different equiva­
lence ratios at 10 atm.

5. Conclusion

The auto-ignition properties of methane/propane mixtures in ratios


from 90/10% to 70/30% diluted in CO2 were measured at equivalence
ratios of 0.5, 1.0, and 2.0; pressures of 1.75 and 10 atm; and tempera­
tures in the range of 1240–1520 K in a shock tube. Four key reactions in
the Oxymech 2.0 model were updated and the modified model shows
reasonable performance in both the IDT and the LFS predictions in O2/
N2 and O2/CO2 atmospheres.
The addition of propane promotes the ignition significantly in the
C3H8 proportion range from 0 to 30% in both cases of CO2 and N2 but
Fig. 9. Physical and chemical effects of CO2 on the ignition at 1400 K and has slight effects on the ignition with the C3H8 proportion more than
10 atm. 30%. The fast increase in the methyl radicals in the initial stage is

6
W. Xia et al. Fuel 343 (2023) 127991

Fig. 10. Experimental and calculated IDTs of CH4/C3H8 mixtures at different equivalence ratios.

Fig. 11. Top 15 reactions influencing the ignition in different equivalence ratios at 1400 K and 1.75 atm.

through C3H8(+M) ⇔ C2H5 + CH3 (+M) owing to the easier breaking of effects of the propane addition. The positive (from R327) and negative
C-C bonds in propane, which enhances CH3 + HO2 ⇔ OH + CH3O, effects (from R2653) due to chaperon effects of CO2 approximately
consequently promoting the ignition. The physical effects of CO2 due to cancel each other out.
the higher volume-specific heat of CO2 compared to N2 are slighter The IDTs increase with the equivalence ratio at the pressure of 1.75
larger than the chemical effects of CO2, and the chaperon effects are very atm but there is no statistical difference observed between these
close to the effects of the reactions with CO2. The effects of CO2 on the measured data of different equivalence ratios at 10 atm. The negative
promoting influence of the propane addition are slight. The chaperon effects of H-abstraction reactions by H radicals of the fuel molecules
effects of CO2 sourced from R327 and R2653 impact the promoting increase with the equivalence ratio at 1.75 atm. H + O2(+M) ⇔

7
W. Xia et al. Fuel 343 (2023) 127991

Fig. 12. Top 18 reactions influencing the ignition in different equivalence ratios at 1300 K and 10 atm.

HO2(+M) suppresses the ignition at the fuel-lean condition but promotes Appendix A. Supplementary data
the ignition in stoichiometric and fuel-rich conditions at 10 atm.
Supplementary data to this article can be found online at https://doi.
CRediT authorship contribution statement org/10.1016/j.fuel.2023.127991.

Wenxiang Xia: Investigation, Formal analysis, Writing – original References


draft. Chun Zou: Methodology, Supervision, Writing – review & editing.
Yi Yuan: Investigation, Data curation, Validation. Rui Fu: Data cura­ [1] Buhre BJP, Elliott LK, Sheng CD, Gupta RP, Wall TF. Oxy-fuel combustion
technology for coal-fired power generation. Prog Energy Combust Sci 2005;31(4):
tion. Jiacheng Liu: Data curation. Chao Peng: Validation. 283–307.
[2] Normann F, Andersson K, Leckner B, Johnsson F. Emission control of nitrogen
Declaration of Competing Interest oxides in the oxy-fuel process. Prog Energy Combust Sci 2009;35(5):385–97.
[3] Toftegaard MB, Brix J, Jensen PA, Glarborg P, Jensen AD. Oxy-fuel combustion of
solid fuels. Prog Energy Combust Sci 2010;36(5):581–625.
The authors declare that they have no known competing financial [4] Chen L, Yong SZ, Ghoniem AF. Oxy-fuel combustion of pulverized coal:
interests or personal relationships that could have appeared to influence Characterization, fundamentals, stabilization and CFD modeling. Prog Energy
Combust Sci 2012;38(2):156–214.
the work reported in this paper. [5] Mendiara T, Glarborg P. Reburn chemistry in oxy-fuel combustion of methane.
Energy Fuel 2009;23(7):3565–72.
Data availability [6] Allam RJ, Fetvedt JE, Forrest BA, Freed DA. The Oxy-Fuel, Supercritical CO2 Allam
Cycle: New Cycle Developments to Produce Even Lower-Cost Electricity from Fossil
Fuels without Atmospheric Emissions. Proceedings of ASME Turbo Expo 2014:
No data was used for the research described in the article. Turbine Technical Conference and Exposition 2014:GT2014-26952.
[7] Sleiti AK, Al-Ammari W, Ahmed S, Kapat J. Direct-fired oxy-combustion
Acknowledgments supercritical-CO2 power cycle with novel preheating configurations
-thermodynamic and exergoeconomic analyses. Energy 2021;226:120441.
[8] Faramawy S, Zaki T, Sakr AAE. Natural gas origin, composition, and processing: a
This work was supported by the general program (52276111) of the review. J Nat Gas Sci Eng 2016;34:34–54.
National Natural Science Foundation of China. [9] Huang J, Bushe WK. Experimental and kinetic study of autoignition in methane/
ethane/air and methane/propane/air mixtures under engine-relevant conditions.
Combust Flame 2006;144(1–2):74–88.

8
W. Xia et al. Fuel 343 (2023) 127991

[10] Petersen EL, Kalitan DM, Simmons S, Bourque G, Curran HJ, Simmie JM. Methane/ [23] Xia W, Peng C, Zou C, Liu Y, Lu L, Luo J, et al. Shock tube and modeling study of
propane oxidation at high pressures: experimental and detailed chemical kinetic ignition delay times of propane under O2/CO2/Ar atmosphere. Combust Flame
modeling. Proc Combust Inst 2007;31(1):447–54. 2020;220:34–48.
[11] Healy D, Curran HJ, Dooley S, Simmie JM, Kalitan DM, Petersen EL, et al. [24] Laich AR, Baker J, Ninnemann E, Sigler C, Naumann C, Braun-Unkhoff M, et al.
Methane/propane mixture oxidation at high pressures and at high, intermediate Effects of high fuel loading and CO2 Dilution on oxy-methane ignition inside a
and low temperatures. Combust Flame 2008;155(3):451–61. shock tube at high pressure. J Energy Res Technol 2020;142(10).
[12] Zsély IG, Nagy T, Simmie JM, Curran HJ. Reduction of a detailed kinetic model for [25] Ninnemann E, Pryor O, Barak S, Neupane S, Loparo Z, Laich A, et al. Reflected
the ignition of methane/propane mixtures at gas turbine conditions using shock-initiated ignition probed via simultaneous lateral and endwall high-speed
simulation error minimization methods. Combust Flame 2011;158(8):1469–79. imaging with a transparent, cylindrical test-section. Combust Flame 2021;224:
[13] Pachler RF, Ramalingam AK, Heufer KA, Winter F. Reduction and validation of a 43–53.
chemical kinetic mechanism including necessity analysis and investigation of CH4/ [26] Petersen EL, Hanson RK. Measurement of reflected-shock bifurcation over a wide
C3H8 oxidation at pressures up to 120 bar using a rapid compression machine. Fuel range of gas composition and pressure. Shock Waves 2006;15(5):333–40.
2016;172:139–45. [27] Morley C. Gaseq v0.76; Available from: http://www.gaseq.co.uk.
[14] Lowry W, de Vries J, Krejci M, Petersen E, Serinyel Z, Metcalfe W, et al. Laminar [28] Baulch DL, Bowman CT, Cobos CJ, Cox RA, Just Th, Kerr JA, et al. Evaluated
flame speed measurements and modeling of pure alkanes and alkane blends at kinetic data for combustion modeling: supplement II. J Phys Chem Ref Data 2005;
elevated Pressures. J Eng Gas Turbines Power 2011;133(9). 34(3):757–1397.
[15] Nilsson EJK, van Sprang A, Larfeldt J, Konnov AA. The comparative and combined [29] Jasper AW, Klippenstein SJ, Harding LB, Ruscic B. Kinetics of the reaction of
effects of hydrogen addition on the laminar burning velocities of methane and its methyl radical with hydroxyl radical and methanol decomposition. J Phys Chem A
blends with ethane and propane. Fuel 2017;189:369–76. 2007;111(19):3932–50.
[16] Kuppa K, Goldmann A, Schöffler T, Dinkelacker F. Laminar flame properties of [30] Miller JA, Klippenstein SJ. The H + C2H2(+M) ⇄ C2H3(+M) and H + C2H2(+M)
C1–C3 alkanes/hydrogen blends at gas engine conditions. Fuel 2018;224:32–46. ⇄ C2H5(+M) reactions: electronic structure, variational transition-state theory,
[17] Huang Z, Zhang Y, Zeng K, Liu B, Wang Q, Jiang D. Measurements of laminar and solutions to a two-dimensional master equation. Phys Chem Chem Phys 2004;6
burning velocities for natural gas–hydrogen–air mixtures. Combust Flame 2006; (6):1192–202.
146(1–2):302–11. [31] Zhou C-W, Li Y, Burke U, Banyon C, Somers KP, Ding S, et al. An experimental and
[18] El-Sabor Mohamed AA, Panigrahy S, Sahu AB, Bourque G, Curran HJ. An chemical kinetic modeling study of 1,3-butadiene combustion: Ignition delay time
experimental and kinetic modeling study of the auto-ignition of natural gas blends and laminar flame speed measurements. Combust Flame 2018;197:423–38.
containing C1–C7 alkanes. Proc Combust Inst 2021;38(1):365–73. [32] Zhang K, Banyon C, Bugler J, Curran HJ, Rodriguez A, Herbinet O, et al. An
[19] Metcalfe WK, Burke SM, Ahmed SS, Curran HJ. A hierarchical and comparative updated experimental and kinetic modeling study of n-heptane oxidation. Combust
kinetic modeling study of C1–C2 hydrocarbon and oxygenated fuels. Int J Chem Flame 2016;172:116–35.
Kinet 2013;45(10):638–75. [33] CHEMKIN-PRO 15151. Reaction Design: San Diego, CA; 2015.
[20] Hargis JW, Petersen EL. Methane ignition in a shock tube with high levels of CO2 [34] Zhang J, Niu S, Zhang Y, Tang C, Jiang X, Hu E, et al. Experimental and modeling
dilution: consideration of the reflected-shock bifurcation. Energy Fuel 2015;29 study of the auto-ignition of n-heptane/n-butanol mixtures. Combust Flame 2013;
(11):7712–26. 160(1):31-9.
[21] Liu Y, Zou C, Cheng J, Jia H, Zheng C. Experimental and numerical study of the [35] Zhang D, Wang Y, Zhang C, Li P, Li X. Experimental and numerical investigation of
effect of CO2 on the ignition delay times of methane under different pressures and vitiation effects on the auto-ignition of n-heptane at high temperatures. Energy
temperatures. Energy Fuel 2018;32(10):10999–1009. 2019;174:922–31.
[22] Liu Y, Cheng J, Zou C, Lu L, Jing H. Ignition delay times of ethane under O2/CO2 [36] Ji W, Ren Z, Law CK. Evolution of sensitivity directions during autoignition. Proc
atmosphere at different pressures by shock tube and simulation methods. Combust Combust Inst 2019;37(1):807–15.
Flame 2019;204:380–90. [37] Sabia P, Lubrano Lavadera M, Giudicianni P, Sorrentino G, Ragucci R, de
Joannon M. CO2 and H2O effect on propane auto-ignition delay times under mild
combustion operative conditions. Combust Flame 2015;162(3):533–43.

You might also like