1 s2.0 S0016236122001703 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Fuel 316 (2022) 123301

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 dimethyl


ether/propane mixtures in O2/CO2 atmospheres
Jinling Yang , Chun Zou *, Wenyu Li *, Qianjin Lin , Lixin Lu , Wenxiang Xia
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: This work measured the ignition delay times (IDTs) of dimethyl ether (DME)/propane mixtures in O2/CO2 at­
Dimethyl ether (DME) mosphere in a shock tube at an equivalence ratio of 0.5, a temperature range of 1157–1380 K, pressures of 2 and
Propane 10 atm, and a DME blending ratio range of 0–100%. A detailed kinetic model referred to as the DME-C3H8 model
Ignition delay time (IDT)
was constructed based on the OXYMECH 2.0 model and the DME sub-mechanism of Aramco 3.0. The DME-C3H8,
O2/CO2 atmosphere
Aramco 3.0, NUIG 1.0, and Dames models were evaluated through the IDTs obtained in the present study, and
from literature in O2/N2 and O2/Ar atmosphere. The impact of the DME blending ratio was investigated using the
DME-C3H8 model, and results show that the IDTs in O2/CO2 atmosphere at pressures of 2 and 10 atm decrease
linearly with increased DME blending ratio; there exists a clear synergy effect between DME and propane in the
O2/N2 atmosphere at the pressure of 2 atm, which is significantly weakened at the pressure of 10 atm. The
impact of CO2 and pressure on the ignition were discussed in detail.

1. Introduction logarithmic IDTs decrease linearly with increased DME blending ratio at
high temperatures. Dames et al. [26] measured the IDTs of stoichio­
In the past few decades, the emission of CO2 has steadily increased, metric DME/C3H8/air mixture in a rapid compression machine at a
causing serious environmental concerns. Pressurized oxy-fuel combus­ temperature range of 550–2000 K and pressures of 10–50 atm. Dames
tion (POC) is a new generation of oxy-fuel combustion that has lower proposed a model named Dames-model for simulating the measured
cost, reduced emissions, and higher efficiency [1–6]. In POC, the com­ IDTs. They found that the IDTs of the DME/propane mixture are very
bustion primarily takes place at elevated pressures and in O2/CO2 at­ sensitive to the rates of C3H8 + OH = products, especially at higher DME
mosphere. As reported in the World Energy Outlook [7], POC combined concentrations.
with bioenergy can realize “negative” CO2 emissions. In addition, According to the authors’ knowledge, there are limited studies on
exhaust gas recirculation (EGR) is widely used to reduce NOx emissions, DME/C3H8 IDTs in O2/CO2 atmosphere. Thus, in this work, one goal is
in which combustion occurs with higher CO2 concentrations [8–12]. to measure the IDTs of DME/C3H8 mixtures in O2/CO2 atmosphere and
Therefore, it is necessary to study fuel ignition characteristics in the rich to evaluate the kinetic model of DME/C3H8 combustion.
CO2 atmosphere. Lu et al. [27] measured the IDTs of DME/C2H6 mixture in O2/CO2
In recent years, dimethyl ether (DME) has been considered as an atmosphere in a shock tube (ST) at an equivalence ratio of 0.5, a tem­
alternative fuel and an ignition enhancer due to its high reactivity, high perature range of 1126–1449 K, pressures of 0.8, 2, and 10 atm, and
cetane number, and wide availability [13–15]. Moreover, DME is DME fractions range from 0 to 100%. They found that, at the pressure of
considered a second-generation biofuel [16] as it can be economically 10 atm and the temperature of 1300 K, in O2/CO2 atmosphere, the IDTs
produced from biomass on a large scale [17]. As a result, alkane fuel first decrease with increased DME fraction from 0 to 20%, and then
mixed with DME has attracted the growing attention of many re­ increase as the DME fraction continues to grow whereas the IDTs
searchers [16,18–24]. monotonously increase with increased DME fraction in O2/N2 atmo­
Hu et al. [25] measured the IDTs of DME/C3H8 mixture in O2/Ar sphere. The impact of CO2, mainly contributing through the third-body
atmosphere in a shock tube at 20 bar, temperatures between 1100 and reactions CH3OCH3 (+M) <=> CH3 + CH3O (+M) and H2O2 (+M) <=>
1500 K, and equivalent ratios of 0.5, 1, and 2. In their study, the 2OH (+M), is responsible for their observations. Djordjevic et al. [28]

* Corresponding authors.
E-mail addresses: zouchun@hust.edu.cn (C. Zou), wenyuli@hust.edu.cn (W. Li).

https://doi.org/10.1016/j.fuel.2022.123301
Received 14 October 2021; Received in revised form 5 January 2022; Accepted 13 January 2022
Available online 21 January 2022
0016-2361/© 2022 Elsevier Ltd. All rights reserved.
J. Yang et al. Fuel 316 (2022) 123301

measured the IDTs of DME containing high amounts of CO2 in a shock Table 1
tube at pressures of 15,35 and 50 bar, a temperature range of 744–1316 Experimental conditions.
K, and equivalence ratios of 0.5–2.0. The results showed that CO2 Mixtures DME blending ratio XDME XC3 H8 XO2 XCO2 (%)
dilution has a great influence on IDTs in the negative temperature co­ (%) (%) (%) (%)
efficient region due to the high heat capacity of CO2, while the effects of
DME00 0 0.00 2.06 20.58 77.37
CO2 are less pronounced at low and high temperature ranges. Addi­ DME20 20 0.45 1.79 20.54 77.23
tionally, the influence of CO2 on third-body reactions leads to a slight DME50 50 1.28 1.28 20.47 76.97
acceleration of ignition. Liu et al. [29] measured the DME flame speed at DME80 80 2.40 0.60 20.38 76.62
equivalent ratios of 0.8 and 1.63, pressures from 1 to 20 bar, and CO2 DME100 100 3.38 0.00 20.30 76.32

fractions of 0 and 20%. The results showed that both the physical and
chemical properties of CO2 reduced the flame speeds. Moreover, Xia
et al. [30] measured the IDTs of propane in O2/CO2/Ar atmosphere
using a shock tube at pressures of 1 and 10 bar, temperatures from 1200
to 1600 K, equivalence ratios of 0.5, 1.0, and 1.5, and high CO2 fractions
of 60% and 80%. They proposed OXYMECH 2.0 based on the work of Liu
et al. [31,32] and found that the physical effects of CO2 are much
stronger than its chemical effects, both result in the longer IDTs in CO2
cases than those in N2 cases. The chaperon effects of CO2 promote the
ignition at low temperatures and are very weak at high temperatures
and at P > 2 bar. The effects of reactions involving CO2 do not vary
considerably with increased pressures and temperatures. Clearly, high
CO2 concentration has a significant influence on the ignition of DME/
C3H8 mixtures in O2/CO2 atmosphere. Thus, another aim of this work is
to analyze the effects of CO2 on the IDTs of the DME/C3H8 mixtures in
detail.
The Aramco 3.0 [33] model predicts well the low and high temper­
ature DME IDTs (Jiang et al. [22] and Li et al. [34]) and the OXYMECH
2.0 model has been validated against the IDTs, flame speeds, and species
profiles of propane in O2/CO2 atmosphere and O2/N2 atmosphere. Thus,
a reaction model for DME/propane mixtures, named the DME-C3H8 Fig. 1. Ignition delay time measure of DME00 at the pressure of 2.04 atm, the
model, was constructed based on the Aramco 3.0 and OXYMECH 2.0 temperature of 1333 K and the equivalence ratio of 0.5.
models in the present study.
In this work, the IDTs of the DME/C3H8/O2/CO2 mixture were XDME and XC3 H8 are the mole fraction of DME and propane, respectively.
measured using a shock tube (ST) at an equivalence ratio of 0.5, a The equivalence ratio, defined as φ = (F/O)real /(F/O)stoic , was set to 0.5.
temperature range of 1157–1380 K, pressures of 2 and 10 atm, and a Table 1 shows the experimental conditions. High purity (>99.999%)
DME blending ratio range of 0–100%. The DME-C3H8 model, Aramco helium and nitrogen blends were used as the driving gases. The purities
3.0 model, NUIG 1.0 model [35], and Dames-model [26] were evaluated of DME, C3H8, CO2, and O2 are greater than 99.5%, 99.99%, 99.999%,
using the experimental data measured in this study and from literature. and 99.999%. The pressure (P5) and temperature (T5) behind the re­
The impact of DME addition on the ignition of DME/C3H8 mixtures in flected shock waves were calculated by Gaseq [36] chemical equilibrium
O2/CO2 atmosphere was compared to that of O2/N2 atmosphere at program.
various pressures and temperatures. The effects of CO2 addition and
pressure on the ignition of DME/C3H8 mixtures in O2/CO2 atmosphere
2.2. Definition of IDTs
were analyzed in detail.
As shown in Fig. 1, the IDT is defined as the time interval between the
2. Experiment and numerical simulation
arrival of the reflected shocks and the onset of ignition, which is
determined by extrapolating the maximum slope of the normalized OH*
2.1. Experimental methods
emission linearly back to the zero baselines. The shock bifurcation was
caused by the large addition of CO2 to the mixture, resulting in a slight
The IDTs of the DME/C3H8/O2/CO2 mixtures were measured in a
increase in the pressure during combustion. The shock bifurcation in O2/
shock tube with an inner diameter of 0.1 m. The shock tube facility has
CO2 atmosphere has been investigated in detail by Ninnemann et al.
been described in detail in our previous work [31] and only a brief
[37] and Laich et al. [38]. They have found that the momentum defi­
introduction is given here. The lengths of the driven and driver sections
ciency in the boundary layer is the main reason for the reflected-shock
are 4 and 8 m, respectively. The driven section and driver section are
bifurcation. The bifurcation brings difficulty in determining the arrival
separated by polyester terephthalate (PET) diaphragms. The incident
of the main reflected wave. Thus, the arrival time of the reflected waves
shock velocity was measured by five piezoelectric pressure transducers
was determined using the empirical expression proposed by Petersen
(PCB 111A24) placed along the sides of the tube with an interval of 0.2
and Hanson [39]:
m. A piezoelectric pressure transducer (Kistler 603B1) located 0.02 m
from the end wall was used to capture the pressure, while the OH* ΔtAO (μs) = 4.6MS0.66 γ −2 7.1 M
0.57

emission was monitored by a photomultiplier with a band-pass filter of


307 ± 10 nm placed 0.02 m from the end wall. where MS is the Mach number of the incident shocks, γ 2 is the specific
All the mixtures were prepared in a 300 L mixing tank, which was heat of the upstream mixture, and M. is the average molecular weight of
evacuated to below 0.1 Pa, using the Dalton’s law of partial pressure. the mixtures. The arrival time of the reflected shock waves, tA , is defined
The mixtures were then settled in the mixing tank for > 12 h. In this as follows:
study, the oxidizer is a CO2/O2 mixture, in which the ratio of CO2/O2 is tA =ΔtAO + 2VDR + ti
set to 79/21. The fuel is a mixture of C3H8 and DME and the blending where D is the diameter of the piezoelectric pressure transducer
ratio of DME is calculated as χ DME = XDME /(XDME + XC3 H8 ), in which (Kistler 6B1), VR is the reflected shock velocity and ti is the time of initial

2
J. Yang et al. Fuel 316 (2022) 123301

Fig. 2. Comparison of the experiment data in the present work and the simulated ones using the DME-C3H8 model, Aramco 3.0 model, NUIG 1.0 model, and
Dames2016 model.

pressure rise. is 1.35 m/s atT5 = 1376 K. The uncertainty of T5 can be calculated using
RRS as follows.

2.3. Uncertainty analysis ∂T5 δVs


δ T5 = δM = (190M + 112M − 3 ) √̅̅̅̅̅̅̅̅̅̅
∂M γRT1
The reflected temperature (T5 ) was calculated using a 1-D shock-tube
Vs
relation [40]. The definition of T5 is as follows: M = √̅̅̅̅̅̅̅̅̅̅
γRT1
[ ][ ]
T1 2(γ − 1)M 2 + (3 − γ) (3γ − 1)M 2 − 2(γ − 1)
T5 = The largest error of T5 was estimated to be ±17 K.
(γ + 1)2 M 2 The measured IDTs can be accurately represented using the
following equation:
where T1 is the initial temperature (290 K), γ is the mixture specific ratio
and M is the incident shock Mach number. For DME00, γ= 1.300 and the τ = APa exp( )
b
equation becomes: T
By performing multiple linear regression on lnτ, lnP, and 1/T of the
T5 = 95M 2 + 251 − 56M − 2
experimental data of DME00, we can get A =4.69 × 10− 5 , a = -0.738 and
According to the study of Zhang et al. [41] and Petersen et al. [42], b = 26640.05. Therefore, the uncertainty of IDT measurements can be
we adopted the standard root-sum-squares (RRS) method [41–43] for calculated as:
uncertainty analysis. The uncertainty of incident-shock velocity δVs is √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( b )2 ( b )2
described as follows: δτ = AeT aPa− 1 δp + AeT aPa δT
√̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅
( )2 ( ̅
1 − Δz )2 where δp is the uncertainty of pressure and was estimated to be 3%. The
δ Vs = δΔz + δΔt
Δt Δt2 error in determining the maximum OH* slope was estimated to be 3%. In
total, the uncertainty of measured IDTs was less than 22%.
Δz
Δt =
Vs
2.4. Modeling
where Δz is the distance between the two adjacent pressure transducers,
δΔz is the machining error of the distance between two adjacent pressure A detailed kinetic model referred to as the DME-C3H8 model was
transducers, δΔt represents the response error of the pressure transducers constructed based on the OXYMECH 2.0 model proposed by Xia et al.
during the measurement of time, and Δt represents the time for the [30] and the DME sub-mechanism of Aramco 3.0. Three other models
incident shock wave to pass through two adjacent pressure transducers. including DME and C3H8 in addition to the DME-C3H8 model, termed the
For the facilities used in this work, δΔt = 2 × 10− 7 s, δΔz = 1 × 10− 4 m, Aramco 3.0 [33], NUIG 1.0 [35], and Dames [26] models, were also
Δz= 0.2 m. Thus, for DME00, the largest error of incident-shock velocity evaluated using the newly measured experiment data. The simulations

3
J. Yang et al. Fuel 316 (2022) 123301

Fig. 2. (continued).

were carried out using the closed zero-dimensional reactor model in the the simulated ones using the four models. It is observed that the DME-
CHEMKIN-PRO software [44]. According to our previous study [31] and C3H8 model consistently provides the best agreement between simula­
the studies of [45–48], since the ignition delay times measured in this tions and the data. The Aramco 3.0 model only produces a satisfactory
work were shorter than 2000 μs, the pressure rise resulting from non- agreement between its simulations and the data for pure DME (DME100)
ideal effects can be considered negligible during simulation of the and overpredicts the IDTs of other DME/C3H8 mixtures. The NUIG 1.0
relatively short IDTs of interest here. Therefore, a constant-volume, model predicts well the IDTs of pure DME (DME100) at 10 atm and
zero-dimensional chemistry model (U and V assumption) was used overpredicts the IDTs of other DME/C3H8 mixtures. The Dames model
without modification in the current study [49,50]. The Definition of the gives good predictions for pure C3H8 at 10 atm and overpredicts the IDTs
IDT in the simulation is the same as that in the experiment. of other mixtures. The IDTs of DME/C3H8 mixtures in O2/Ar atmosphere
measured by Hu et al. [25] were used to evaluate these four models and
3. Results and discussion the results are shown in Fig. 3. As presented in Fig. 3, all the models
generate overall good predictions of DME/C3H8 IDTs in O2/Ar atmo­
3.1. Experimental results and model predictions sphere. The IDTs of pure DME in O2/N2 atmosphere measured by Jiang
et al. [20], Burke et al. [16], and Li et al. [34] were also used to evaluate
The IDTs of DME/C3H8 mixtures in O2/CO2 atmosphere at the DME the four models and the results are shown in Figs. 4–6, respectively. As
blending ratios of 0, 20, 50, 80, and 100% were measured at an equiv­ shown in Figs. 4 and 5, all the models except for the Dames model
alence ratio of 0.5, a temperature range of 1157–1380 K, and pressures predict well the low-temperature IDTs measured by Jiang and Burke. For
of 2 and 10 atm. The experimental data can be obtained from Supple­ the high-temperature IDTs measured by Li, the DME-C3H8 and Aramco
mental Materials. 3.0 models generate overall good predictions. The results demonstrate
Fig. 2(a)–(j) demonstrate the comparison of the measured IDTs and the reliability of the DME-C3H8 model, which was used for the following

4
J. Yang et al. Fuel 316 (2022) 123301

Fig. 3. Comparison between the experiment date of Hu et al. [25] and the simulated ones calculated by the DME-C3H8 model, Aramco 3.0 model, NUIG 1.0 model,
and Dames model.

Fig. 4. Comparison of the experiment date of Jiang et al. [20] and the simulated ones calculated by the DME-C3H8 model, and Aramco 3.0 model.

kinetic analysis. fraction of C3H5-A is much higher in the DME-C3H8 model, as shown in
The brute force sensitivity of the simulated DME20 IDTs at the Fig. 9(a). As a result, the net reaction rate of R569 in the DME-C3H8
pressure of 2 atm and the temperature of 1300 K was calculated based on model is faster than that in the Aramco 3.0 model, as shown in Fig. 8 (b),
the DME-C3H8 and Aramco 3.0 models. Fig. 7 (a) and (b) show the top 15 leading to enhanced OH production. Moreover, reaction H2O2 (+M)
reactions with the largest sensitivity magnitude in the DME-C3H8 model <=> 2OH (+M) (R2572) ranks 10th in DME-C3H8 model but is absent in
and Aramco 3.0 model, respectively. As shown in Fig. 7, reaction C3H5- the top 15 sensitive reactions in the Aramco 3.0 model. The rate of
A + HO2 <=>C3H5O + OH (R569) ranks 5th and 7th in the DME-C3H8 production analysis of H2O2 showed that reaction CH2O + HO2 <=>
and Aramco 3.0 model, respectively. Since the reaction rate constants of H2O2 + HCO (R2632) is the main H2O2 production reaction. The reac­
C3H8 + OH <=> IC3H7 + H2O (R341) in the DME-C3H8 model are larger tion rate constants of R2632 in the DME-C3H8 model and the Aramco 3.0
than in Aramco 3.0, as shown in Table 2, the net reaction rates of R341 are shown in Table 2. As shown in Table 2 and Fig. 8(c), the reaction rate
in the DME-C3H8 model are faster than that in Aramco 3.0, as shown in constants of R2632 are much higher in the DME-C3H8 than in the Ara­
Fig. 8(a). Therefore, the reaction pathway C3H8 → IC3H7 → C3H6 → mco 3.0 model, leading to faster H2O2 accumulation, as shown in Fig. 9
C3H5-A , which is the main generation pathway of the C3H5-A in both (b). As a result, R2572 is faster in the DME-C3H8 model as shown in
models, has been promoted in the DME-C3H8 model. Thus, the mole Fig. 8(d) due to higher H2O2 concentrations. Therefore, the

5
J. Yang et al. Fuel 316 (2022) 123301

Fig. 5. Comparison between the experiment date from Burke et al. [16] and the simulated ones calculated by the DME-C3H8 model, Aramco 3.0 model, and
Dames model.

Fig. 6. Comparison of the experiment date of Li et al. [34] and the simulated results calculated by the DME-C3H8 model, and Aramco 3.0 model.

underestimation in rates of R569, R2572, and R2632 is the main reason results are presented in Fig. 10.
for the overprediction made by the Aramco 3.0 model. As shown in Fig. 10(a) and (b), at the pressure of 2 atm and the
temperatures of 1200 K and 1300 K, the IDTs in O2/CO2 atmosphere
3.2. Impact of the DME blending ratio on the IDTs of DME/C3H8 mixture decrease linearly with increased DME blending ratio. The IDTs in O2/N2
atmosphere are noticeably longer than those obtained by linear inter­
In order to explore the impact of the DME blending ratio on the IDTs polating between pure DME and pure C3H8, indicating synergy effects
of the mixtures, various DME blending ratios were used at fixed tem­ between DME and C3H8 in O2/N2 atmosphere. Replacement of N2 by
perature and pressure. However, due to the intrinsic randomness asso­ CO2 in the dilution clearly weakens this synergy effect. Moreover, the
ciated with the shock tube facility, it is difficult to maintain the exact IDTs in CO2 cases are longer than those in N2 cases for the same tem­
temperature and pressure over repeated experiments. Therefore, the perature, pressure, and DME blending ratio due to the differences in
validated DME-C3H8 model was used in the simulation for calculating physical and chemical properties between CO2 and N2. Compared to N2,
the IDTs for different DME blending ratios at a preset pressure, tem­ CO2 is not chemically inert and can participate in chemical reactions as a
perature, and equivalent ratio. The simulation conditions are as follows: chemically active reactant in addition to as a third-body collider. The
pressures of 2 and 10 atm, temperatures of 1200 K and 1300 K, an effects of CO2 will be discussed in the next section.
equivalent ratio of 0.5, and in O2/CO2 and O2/N2 atmosphere. The In Fig. 10(c) and (d), at the pressure of 10 atm and the temperatures

6
J. Yang et al. Fuel 316 (2022) 123301

3.3. Effect of CO2 on the maximum IDT

As shown in Fig. 10, at a pressure of 2 atm and temperatures of 1200


and 1300 K, a synergy effect between DME and propane is observed in
the O2/N2 atmosphere and is weakened in the O2/CO2 atmosphere.
Fig. 14 presents the top 10 reactions with the largest sensitivity
magnitude at the temperature of 1300 K, the pressure of 2 atm, and the
equivalence ratio of 0.5 for DME20 and DME80 in O2/N2 atmosphere.
The results show that, besides reaction H + O2 <=> O + OH (R2556),
reaction CH3 + HO2 <=> OH + CH3O (R2655) has the strongest pro­
motion impact on the ignition. As the DME blending ratio increases,
reaction CH3OCH3 (+M) <=> CH3 + CH3O (+M) (R261) becomes more
important for enhancing ignition through producing CH3. However,
R261 is suppressed to a certain extent by the reaction C3H8 (+M) = CH3
+ C2H5(+M) (R334) due to the Le Chatelier’s principle. The reciprocal
inhibition between R334 and R261 results in the synergy effect between
C3H8 and DME. Therefore, the IDTs in O2/N2 atmosphere are higher
Fig. 7. The top 15 reactions in brute force sensitivity calculated by DME-C3H8 than those obtained by linear interpolating between pure DME and pure
model (yellow bar) and Aramco 3.0 model (green bar) at the temperature of C3H8.
1300 K, the pressure of 2 atm, and the equivalence ratio of 0.5 for DME20 in The effects of CO2 can be divided into the physical effects and
O2/CO2 atmosphere. (For interpretation of the references to colour in this figure
chemical effects, which can be further divided into the effects from re­
legend, the reader is referred to the web version of this article.)
actions involving CO2 and the chaperon effects of CO2. To investigate
these effects separately, three artificial substances, X, Y, and Z, are
Table 2
introduced into the DME-C3H8 model. All these three artificial sub­
The reaction rate constants of R341 and R2632 in the DME-C3H8 model and stances have the same thermochemical property of CO2. The species X is
Aramco 3.0 model. assumed to be chemically inert; the species Y participates in chemical
reactions but does not contribute as a third-body collider; the species Z
R341 C3H8 + OH model A n Ea Source
<=> IC3H7 + H2O only acts like a third-body collider with the same collision efficiency of
CO2. Thus, the differences in the IDTs of DME/C3H8 mixtures between
DME- 1.81E + 2.437 − 536 [51]
C3H8 05
the O2/X atmosphere (τX ) and O2/N2 atmosphere (τN2 ) represent the
Aramco 9.17E + 9.35E- 5.047E + [52] effects of the physical properties of CO2 on the ignition; the differences
3.0 09 01 02 in the IDTs between the O2/X atmosphere and O2/CO2 atmosphere
R2632 CH2O þ HO2 model A n Ea Source (τCO2 ) represent the effects of the chemical properties of CO2; the dif­
<¼> H2O2 þ HCO ferences in the IDTs between the O2/Y atmosphere (τY ) and O2/N2 at­
DME- 7.11E + 2.50 10,210 [53] mosphere represent the effects of the reactions involving CO2 on the
C3H8 4 ignition, while the differences in the IDTs between the O2/Z atmosphere
Aramco 1.88E + 2.70 11,520 [54]
3.0 4
(τZ ) and O2/N2 atmosphere represent the chaperon effects of CO2 on the
ignition. The IDTs of DME/C3H8 in X, Y, and Z cases at the pressure of 2
atm, the temperature of 1300 K, and the equivalence ratio 0.5 are shown
of 1200 and 1300 K, the IDTs in both O2/CO2 and O2/N2 atmosphere in Fig. 15. Like that in N2 cases, the IDTs in the O2/X atmosphere still
decrease linearly with increased DME blending ratio, indicating weak­ vary with the blending ratio in an approximately parabolic manner,
ened minimal synergy effect in the O2/N2 atmosphere. The effects of suggesting the physical effects of CO2 are not the reason for the weak­
pressure will be discussed in Section 3.4. ened synergy effect in O2/CO2 atmosphere. Similar trends are observed
Brute force sensitivity analysis of the IDTs for the five tested mixtures in the O2/Y atmosphere, indicating the effects of reactions involving CO2
was conducted at the temperature of 1300 K, the pressure of 2 atm, and are also not responsible for the reduced synergy effect. Only in O2/Z
the equivalence ratio of 0.5 in O2/CO2 atmosphere, and the results are atmosphere, the IDTs decrease almost linearly with increased blending
shown in Fig. 11. ratios, demonstrating that the chaperon effect of CO2 weakens the syn­
As shown in the Fig. 11, reactions H + O2 <=> O + OH (R2556) and ergy effect.
CH3 + HO2 <=> OH + CH3O (R2655) are the two most important To investigate the effects of CO2 quantitatively, the percentage
facilitating reactions. Reaction C3H8(+M) <=> CH3 + C2H5(+M) variation of the effects (PV) was introduced and is defined as follows:
(R334) promotes ignition for blending less than 50% by providing CH3
PV O = PV P + PV CH
for R2655 and CH3 + O2 <=> OH + CH2O (R2657) and C2H5 to C2H4 +
H <=> C2H5 (M) (R2781). As the DME blending ratio increases, reac­ τ X − τ N2
tion CH3OCH3 (+M) <=> CH3 + CH3O (+M) (R261) shows a growing PV P =
τX
contribution to facilitating ignition by providing CH3 to R2655 for OH
production, and to CH3OCH3 + CH3 <=> CH3OCH2 + CH4 (R267) for PV CH =
τCO2 − τX
CH3OCH2 production, which subsequently decomposes to CH3 and τX
CH2O through reaction CH3OCH2 <=> CH3 + CH2O (R278). As a result,
τY − τX
reactions CH2O + HO2 <=> HCO + H2O2 (R2641) and CH2O + OH PV R =
τX
<=> HCO + H2O (R2639) promote the ignition at blending ratios
higher than 50%. Moreover, reaction HCO + O2 <=> HO2 + CO τ Z − τX
PV CA =
(R2597) is responsible for>75% of HO2 production, which then converts τX
to OH by R2655. Therefore, the net reaction rate of R2655 and the OH
in which,PV O is the PV of overall effects, PV P is the PV of physical
mole fraction increase with increased blending ratio, as shown in
effects, PV CH is the PV of chemical effects,PV R is the PV of the effects of
Figs. 12 and 13, resulting in shorter IDTs.
reactions involving CO2, and PV CA is the PV of the chaperon effects of

7
J. Yang et al. Fuel 316 (2022) 123301

Fig. 8. The comparison on the net reaction rates of (a) R341, (b) R569, (c) R2632, and (d) R2572 in the DME-C3H8 model (red line), and Aramco 3.0 models (black
line) at the temperature of 1300 K, the pressure of 2 atm, and the equivalence ratio of 0.5 for DME20 in O2/CO2 atmosphere. (For interpretation of the references to
colour in this figure legend, the reader is referred to the web version of this article.)

Fig. 9. The comparison on the mole fraction of (a) C3H5-A and (b) H2O2 in the DME-C3H8 model (red line) and Aramco 3.0 models (black line) at the temperature of
1300 K, the pressure of 2 atm, and the equivalence ratio of 0.5 for DME20 in O2/CO2 atmosphere. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)

CO2. other hand, only show a mild inhibiting impact on the ignition for
Fig. 16 shows the effects of CO2 at different DME blending ratios DME00 and become a promotion effect when DME is added. The pro­
calculated based on the DME-C3H8 model at the pressure of 2 atm, the motion effects are stronger at a higher DME blending ratio. Overall, the
temperature of 1300 K, and the equivalence ratio of 0.5. As shown in chaperon effects of CO2 dominate the chemical effects for DME blending
Fig. 16, the physical effects, PVP (orange bars in Fig. 16), significantly ratio>20%.
inhibit the ignition and its caused percentage variation is much larger The brute force sensitivity of the IDT was calculated at the temper­
than PVCH across all tested conditions. The chemical effects, PVCH (green ature of 1300 K, the pressure of 2 atm, and the equivalence ratio of 0.5
bars in Fig. 16), inhibit the ignition for pure propane (DME00) but for DME20 in O2/Z atmosphere. The top 12 reactions with the largest
promote the ignition once DME is added. The ignition enhancement sensitivity magnitude are shown in Fig. 17. The results show that, in O2/
grows with an increased blending ratio from 20 to 100%, leading to Z atmosphere, the third-body reaction 2CH3 (+M) <=> C2H6 (+M)
overall reduced inhibition effects of CO2. The effects of reactions (R2662) inhibits ignition whereas reactions C3H8 (+M) <=> CH3 +
involving CO2, PVR (yellow bars in Fig. 16), inhibit the ignition for all C2H5 (+M) (R334) and CH3OCH3 (+M) <=> CH3 + CH3O(+M) (R261)
tested cases and its inhibiting impacts slightly weaken at a higher DME promote the ignition.
blending ratio. The chaperon effects of CO2 (blue bars in Fig. 16), on the Furthermore, three artificial substances with the same

8
J. Yang et al. Fuel 316 (2022) 123301

Fig. 10. Impact of DME blending ratio on IDTs of DME/C3H8 mixtures.

Fig. 11. The top 10 reactions with largest brute force sensitivity magnitude on Fig. 12. The net reaction rates of R2655 at the temperature of 1300 K, the
ignition of DME00 - DME100 at the temperature of 1300 K, the pressure of 2 pressure of 2 atm, and the equivalence ratio of 0.5 of DME00 - DME100.
atm, and the equivalence ratio of 0.5.

mitigate the synergy effect. The IDTs in the O2/Z3 atmosphere, however,
thermochemical property of CO2, Z1, Z2, and Z3, were introduced to decrease almost linearly with increased DME blending ratios, exhibiting
investigate the effects of R2662, R334, and R261 separately. Z1, Z2, and substantially reduced synergy effect due to chaperon effects of CO2 in
Z3 were set to active third-body colliders with the same collision effi­ R261.
ciency of CO2 only for R2662, R334, and R261, respectively. The The percentage variation of the chaperon effects (PV) of CO2 in
simulated IDTs in O2/Z1, O2/Z2 and O2/Z3 atmosphere at the pressure of R2662, R334, and R261 are defined as follows:
2 atm, the temperature of 1300 K, and the equivalence ratio of 0.5 are in
τ Z1 − τ X
Fig. 18. The results show that the IDTs vary with DME blending ratio in PV R2662 =
τX
an approximately parabolic manner in O2/Z1 atmosphere and O2/Z2
atmosphere, indicating the chaperon effect in R2662 and R334 does not

9
J. Yang et al. Fuel 316 (2022) 123301

Fig. 13. The mole fraction of OH radicals at the temperature of 1300 K, the
pressure of 2 atm, and the equivalence ratio of 0.5 of DME00-DME100. Fig. 15. Comparison of IDTs of DME/C3H8 mixture at the pressure of 2 atm, the
temperature of 1300 K and the equivalence ratio of 0.5 in O2/CO2, O2/X, O2/Y
O2/Z and O2/N2 atmosphere.

30
Inhibition

20
Percentage variation(%)

10

0
PVP
PVCH
-10 PVO
PVR
Fig. 14. The top 10 reactions in brute force sensitivity at the temperature of Promotion
PVCA
1300 K, the pressure of 2 atm, and the equivalence ratio of 0.5 for DME20 and -20
DME80 in O2/N2 atmosphere. -20 0 20 40 60 80 100 120
DME blending ratio
τ Z2 − τ X
PV R334 =
τX Fig. 16. Effects of CO2 on the IDTs of DME/C3H8 mixtures at various DME
blending ratios calculated by DME-C3H8 model.
τ Z3 − τ X
PV R261 =
τX temperature of 1300 K, the pressure of 2 atm, and the equivalence ratio
Fig. 19 presents the calculated PV R2662 , PV R334 , and PV R261 for of 0.5 for DME20 in O2/CO2 atmosphere. As shown in Fig. 20, the
different DME blending ratios using the DME-C3H8 model at the pressure suppression effects of reactions involving CO2 mainly come from the
of 2 atm, the temperature of 1300 K, and the equivalence ratio of 0.5. As reaction CO + OH <=> CO2 + H (R2589). However, the brute force
shown in Fig. 19, R2662 (yellow bars in Fig. 19) inhibits ignition and sensitivity coefficient of R2589 is relatively small, ranking 25th, so it has
PV R2662 slightly increases with increased DME blending ratio. Reaction fewer suppression effects on the mixture’s ignition, which is consistent
R334 (green bars in Fig. 19) promotes the ignition in the DME blending with Fig. 16.
ratio range from 0 to 80% but shows a minimal inhibiting impact for
pure DME due to the absence of C3H8. Reaction R261 (purple bars in
Fig. 19), which displays a very small inhibiting influence on the IDT for 3.4. Influence of pressure
DME00, becomes a considerable ignition enhancer when DME is added.
The promoting effect, PV R261 , increases with increased DME blending As discussed above, the synergy effect between DME and propane in
ratios. Among these three reactions, R261 displays the largest effect. O2/N2 atmosphere at the pressure of 2 atm is caused by the reciprocal
Therefore, in CO2 atmosphere, the enhanced R261 generates a larger inhibition between R261 (CH3OCH3 (+M) <=> CH3 + CH3O (+M)) and
CH3 production, which further promotes CH3 + HO2 = OH + CH3O R334 (C3H8 (+M) <=> CH3 + C2H5 (+M)), which significantly in­
(R2655) and generates a larger amount of OH, thus promoting ignition fluences the OH production reaction, R2655 (CH3 + HO2 <=> OH +
and weakening the synergy effect between DME and C3H8. CH3O).
Fig. 20 shows the top 26 reactions in the brute force sensitivity at the The brute force sensitivity was calculated at the temperature of 1300
K, the pressures of 2 and 10 atm, and the equivalence ratio of 0.5 for

10
J. Yang et al. Fuel 316 (2022) 123301

Fig. 17. The top 12 reactions in the brute force sensitivity at the temperature of
1300 K, the pressure of 2 atm and the equivalence ratio of 0.5 for DME20 in O2/
Z atmosphere.
Fig. 19. The chaperon effects sourced from R2662, R334, and R261 on the
IDTs of DME/C3H8 mixtures of various DME blending ration calculated by the
DME-C3H8 model.

Fig. 18. Comparison of IDTs of DME/C3H8 mixture at the pressure of 2 atm, the
Fig. 20. The top 26 reactions in the brute force sensitivity at temperature of
temperature of 1300 K, and the equivalence ratio of 0.5 in O2/Z1, O2/Z2, and
1300 K, pressure of 2 atm, and equivalence ratio of 0.5 for DME20 in O2/CO2
O2/Z3 atmosphere.
atmospheres.

DME50 in O2/N2, and the results are shown in Fig. 21. As shown in
Fig. 21, reactions related to HO2, i.e., CH3 + HO2 <=> OH + CH3O
(R2655), CH2O + HO2 <=> H2O2 + HCO (R2641), H2O2 (+M) <=>
2OH (+M) (R2581), CH3 + HO2 <=> O2 + CH4 (R2654), 2HO2 <=>
H2O2 + O2 (R2579) and H + HO2 <=> H2 + O2 (R2570), play an
important role on ignition at the pressure of 10 atm. One of the main
HO2 production reactions, H + O2(+M) <=> HO2 (+M) (R2569), is
pressure-dependent and strongly enhanced at 10 atm. Fig. 22 (a) shows
the net reaction rates of R2569 at pressures of 2 and 10 atm, respec­
tively. The net reaction rates of R2569 are much higher at the pressure of
10 atm than at the pressure of 2 atm, leading to faster HO2 production
and thus higher net reaction rates of R2655 as shown in Fig. 23 and
Fig. 22(b), respectively. The enhanced R2655 at 10 atm weakens the
synergy effect between DME and C3H8 in the O2/N2 atmosphere.

4. Conclusions

This work measured the IDTs of DME/C3H8 mixtures using a shock


tube (ST) at an equivalence ratio of 0.5, a temperature range of Fig. 21. The top 20 reactions in brute force sensitivity at the temperature of
1300 K, the pressures of 2 and 10 atm and the equivalence ratio of 0.5 for
1157–1380 K, pressures of 2 and 10 atm, and a DME blending ratio
DME50 in O2/N2 atmosphere.
range of 0–100% in O2/CO2 atmosphere. A detailed kinetic model

11
J. Yang et al. Fuel 316 (2022) 123301

Fig. 22. The net reaction rates of R2569 and R2655 at T = 1300 K, P = 2 and 10 atm, and φ = 0.5 for DME50.

higher pressures due to a higher HO2 production rate caused by the


enhanced pressure-dependent reaction H + O2 (+M) <=> HO2 (+M)
(R2569). The enhanced reaction R2655 weakens the synergy effect be­
tween DME and C3H8 in the O2/N2 atmosphere.

CRediT authorship contribution statement

Jinling Yang: Investigation, Formal analysis, Writing – original


draft. Chun Zou: Conceptualization, Methodology, Supervision. Wenyu
Li: Supervision, Validation, Writing – review & editing. Qianjin Lin:
Methodology, Data curation. Lixin Lu: Investigation, Data curation.
Wenxiang Xia: Methodology, Data curation.

Declaration of Competing Interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence
the work reported in this paper.

Fig. 23. The mole fraction of HO2 at the temperature of 1300 K, the pressures Acknowledgements
of 2 and 10 atm, and the equivalence ratio of 0.5 for DME50.
This work was supported by the general program (No. 51776081) of
referred to as the DME-C3H8 model was constructed based on the the National Natural Science Foundation of China.
OXYMECH 2.0 model proposed by Xia et al. and the DME sub-
mechanism of Aramco 3.0. The DME-C3H8 model was validated by the Appendix A. Supplementary data
IDTs measured in the present study and those in O2/Ar and O2/N2 at­
mosphere obtained from the literature. Supplementary data to this article can be found online at https://doi.
The DME-C3H8 model showed the overall best agreement between org/10.1016/j.fuel.2022.123301.
model predictions and experimental data of IDTs of DME/C3H8 mixture
in O2/CO2 atmosphere compared to the Aramco 3.0, NUIG 1.0, and References
Dames models. The underestimation in rates of R569 (C3H5-A + HO2
<=>C3H5O + OH), R2572 (H2O2 (+M) <=> 2OH (+M)) and R2632 [1] Hong D, Guo X, Wang C. A reactive molecular dynamics study of HCN oxidation
(CH2O + HO2 <=> H2O2 + HCO) is the main reason for the over­ during pressurized oxy-fuel combustion. Fuel Process Technol 2021;224.
[2] Kim D, Yang W, Huh KY, Lee Y. Demonstration of 0.1 MWth pilot-scale pressurized
prediction made by the Aramco 3.0 model in O2 /CO2 atmosphere. oxy-fuel combustion for unpurified natural gas without CO2 dilution. Energy 2021;
The IDTs of DME/C3H8 mixtures in O2/CO2 atmosphere at pressures 223.
of 2 and 10 atm decrease linearly with increased DME blending ratios. A [3] Yadav S, Mondal SS. A review on the progress and prospects of oxy-fuel carbon
capture and sequestration (CCS) technology. Fuel 2022;308.
synergy effect between DME and propane was observed in O2/N2 at­
[4] Kim D, Ahn H, Yang W, Huh KY, Lee Y. Experimental analysis of CO/H2 syngas
mosphere at the pressure of 2 atm, which is caused by the reciprocal with NOx and SOx reactions in pressurized oxy-fuel combustion. Energy 2021;219.
inhibition between CH3OCH3 (+M) <=> CH3 + CH3O (+M) and [5] Jafari H, Yang W, Ryu C. Evaluation of a distributed combustion concept using 1-D
modeling for pressurized oxy-combustion system with low flue gas recirculation.
C3H8(+M) <=> CH3 + C2H5(+M). The synergy effect is weakened at the
Fuel 2020;263.
pressure of 10 atm. [6] Chen L, Yong SZ, Ghoniem AF. Oxy-fuel combustion of pulverized coal:
The physical effects of CO2 significantly inhibit the ignition and its Characterization, fundamentals, stabilization and CFD modeling. Prog Energy
caused percentage variation is larger than that due to the chemical ef­ Combust Sci 2012;38(2):156–214.
[7] World Energy Outlook 2018 by International Energy Agency.
fects. The promotion impacts on the IDTs by the chemical effect of CO2 [8] Long Y, Li G, Zhang Z, Liang J. Application of reformed exhaust gas recirculation
significantly increase with increased blending ratios. The chaperon ef­ on marine LNG engines for NO emission control. Fuel 2021;291.
fects sourced from CH3OCH3 (+M) <=> CH3 + CH3O (+M) weaken the [9] Qureshi Y, Ali U, Sher F. Part load operation of natural gas fired power plant with
CO2 capture system for selective exhaust gas recirculation. Appl Therm Eng 2021;
synergy effect at the pressure of 2 atm in O2/CO2 atmosphere. 190.
Reaction CH3 + HO2 <=> OH + CH3O (R2655) is enhanced at [10] Aliyu AAA, Akram M, Hughes KJ, Ma L, Ingham DB, Pourkashanian M.
Investigation into simulating Selective Exhaust Gas Recirculation and varying

12
J. Yang et al. Fuel 316 (2022) 123301

Pressurized Hot Water temperature on the performance of the Pilot-scale Advanced [31] Liu Y, Zou C, Cheng J, Jia H, Zheng C. Experimental and numerical study of the
CO2 Capture Plant with 40 wt(%) MEA. Int J Greenhouse Gas Control 2021;107. effect of CO2 on the ignition delay times of methane under different pressures and
[11] Herraiz L, Fernández ES, Palfi E, Lucquiaud M. Selective exhaust gas recirculation temperatures. Energy Fuels 2018;32(10):10999–1009.
in combined cycle gas turbine power plants with post-combustion CO2 capture. Int [32] Liu Y, Cheng J, Zou C, Lu L, Jing H. Ignition delay times of ethane under O2/CO2
J Greenhouse Gas Control 2018;71:303–21. atmosphere at different pressures by shock tube and simulation methods. Combust
[12] De Santis A, Ingham DB, Ma L, Pourkashanian M. CFD analysis of exhaust gas Flame 2019;204:380–90.
recirculation in a micro gas turbine combustor for CO2 capture. Fuel 2016;173: [33] Zhou C-W, Li Y, Burke U, Banyon C, Somers KP, Ding S, et al. An experimental and
146–54. chemical kinetic modeling study of 1,3-butadiene combustion: Ignition delay time
[13] Mahesh Nayak G, Kolhe P, Balusamy S. Experimental study of buoyancy-induced and laminar flame speed measurements. Combust Flame 2018;197:423–38.
instability in the DME and LPG jet diffusion flame. Fuel 2021;291. [34] Li Z, Wang W, Huang Z, Oehlschlaeger MA. Dimethyl ether autoignition at engine-
[14] Park SH, Lee CS. Combustion performance and emission reduction characteristics relevant conditions. Energy Fuels 2013;27(5):2811–7.
of automotive DME engine system. Prog Energy Combust Sci 2013;39(1):147–68. [35] http://c3.nuigalway.ie/combustionchemistrycentre/mechanismdownloads/.
[15] Huang W, Zhao Q, Huang Z, Curran HJ, Zhang Y. A kinetics and dynamics study on [36] Morley C. Gaseq v0.76 http://www.gaseq.co.uk.
the auto-ignition of dimethyl ether at low temperatures and low pressures. Proc [37] Ninnemann E, Pryor O, Barak S, Neupane S, Loparo Z, Laich A, et al. Reflected
Combust Inst 2021;38(1):601–9. shock-initiated ignition probed via simultaneous lateral and endwall high-speed
[16] Burke U, Somers KP, O’Toole P, Zinner CM, Marquet N, Bourque G, et al. An imaging with a transparent, cylindrical test-section. Combust Flame 2021;224:
ignition delay and kinetic modeling study of methane, dimethyl ether, and their 43–53.
mixtures at high pressures. Combust Flame 2015;162(2):315–30. [38] Laich AR, Baker J, Ninnemann E, Sigler C, Naumann C, Braun-Unkhoff M, et al.
[17] Arcoumanis C, Bae C, Crookes R, Kinoshita E. The potential of di-methyl ether Effects of high fuel loading and CO2 dilution on oxy-methane ignition inside a
(DME) as an alternative fuel for compression-ignition engines: a review. Fuel 2008; shock tube at high pressure. J Energy Res Technol 2020;142(10).
87(7):1014–30. [39] Petersen EL, Hanson RK. Measurement of reflected-shock bifurcation over a wide
[18] Tang C, Wei L, Zhang J, Man X, Huang Z. Shock tube measurements and kinetic range of gas composition and pressure. Shock Waves 2006;15(5):333–40.
investigation on the ignition delay times of methane/dimethyl ether mixtures. [40] Gaydon A, Hurle I. Shock tube in high temperature chemical physics. NY: Reinhold
Energy Fuels 2012;26(11):6720–8. Pub. Corp; 1963.
[19] Hu E, Jiang X, Huang Z, Zhang J, Zhang Z, Man X. Experimental and kinetic studies [41] Zhang Z, Hu E, Pan L, Chen Y, Gong J, Huang Z. Shock-tube measurements and
on ignition delay times of dimethyl ether/n-butane/O2/Ar mixtures. Energy Fuels kinetic modeling study of methyl propanoate ignition. Energy Fuels 2014;28(11):
2012;27(1):530–6. 7194–202.
[20] Jiang X, Deng F, Yang F, Huang Z. Ignition delay characteristics and kinetic [42] Petersen EL, Rickard MJA, Crofton MW, Abbey ED, Traum MJ, Kalitan DM.
investigation of dimethyl Ether/n-pentane binary mixtures: interpreting the effect A facility for gas- and condensed-phase measurements behind shock waves. Meas
of the equivalence ratio and dimethyl ether blending. Energy Fuels 2018;32(3): Sci Technol 2005;16(9):1716–29.
3814–23. [43] Holman JP. Experimental methods for engineers. 1994.
[21] Chen Z, Qin X, Ju Y, Zhao Z, Chaos M, Dryer FL. High temperature ignition and [44] CHEMKIN-PRO 15151; Reaction Design: San Diego, CA; 2015.
combustion enhancement by dimethyl ether addition to methane–air mixtures. [45] Hargis JW, Petersen EL. Methane ignition in a shock tube with high levels of CO2
Proc Combust Inst 2007;31(1):1215–22. dilution: consideration of the reflected-shock bifurcation. Energy Fuels 2015;29
[22] Jiang X, Tian Z, Zhang Y, Huang Z. Shock tube measurement and simulation of (11):7712–26.
DME/n-butane/air mixtures: effect of blending in the NTC region. Fuel 2017;203: [46] Nativel D, Cooper SP, Lipkowicz T, Fikri M, Petersen EL, Schulz C. Impact of shock-
316–29. tube facility-dependent effects on incident- and reflected-shock conditions over a
[23] Wu H, Shi Z, Lee C-f, Zhang H, Xu Y. Experimental and kinetic study on ignition of wide range of pressures and Mach numbers. Combust Flame 2020;217:200–11.
DME/n-butane mixtures under high pressures on a rapid compression machine. [47] Zander L, Vinkeloe J, Djordjevic N. Ignition delay and chemical–kinetic modeling
Fuel 2018;225:35–46. of undiluted mixtures in a high-pressure shock tube: Nonideal effects and
[24] Zhang J, Hu E, Pan L, Zhang Z, Huang Z. Shock-tube measurements of ignition comparative uncertainty analysis. Int J Chem Kinet 2021;53(5):611–37.
delay times for the ethane/dimethyl ether blends. Energy Fuels 2013;27(10): [48] Zhang Y, Huang Z, Wei L, Zhang J, Law CK. Experimental and modeling study on
6247–54. ignition delays of lean mixtures of methane, hydrogen, oxygen, and argon at
[25] Hu E, Zhang Z, Pan L, Zhang J, Huang Z. Experimental and modeling study on elevated pressures. Combust Flame 2012;159(3):918–31.
ignition delay times of dimethyl ether/propane/oxygen/argon mixtures at 20 bar. [49] Zhang D, Wang Y, Zhang C, Li P, Li X. Experimental and numerical investigation of
Energy Fuels 2013;27(7):4007–13. vitiation effects on the auto-ignition of n-heptane at high temperatures. Energy
[26] Dames EE, Rosen AS, Weber BW, Gao CW, Sung C-J, Green WH. A detailed 2019;174:922–31.
combined experimental and theoretical study on dimethyl ether/propane blended [50] Zhang J, Niu S, Zhang Y, Tang C, Jiang X, Hu E, et al. Experimental and modeling
oxidation. Combust Flame 2016;168:310–30. study of the auto-ignition of n-heptane/n-butanol mixtures. Combust Flame 2013;
[27] Lu L, Zou C, Xia W, Lin Q, Shi H. Experimental and numerical study on the ignition 160(1):31–9.
delay times of dimethyl ether/ethane/oxygen/carbon dioxide mixtures. Fuel 2020; [51] Sivaramakrishnan R, Srinivasan NK, Su MC, Michael JV. High temperature rate
280. constants for OH+ alkanes. Proc Combust Inst 2009;32(1):107–14.
[28] Djordjevic N, Rekus M, Vinkeloe J, Zander L. Shock Tube and Kinetic Study on the [52] Burke SM, Burke U, Mc Donagh R, Mathieu O, Osorio I, Keesee C, et al. An
Effects of CO2 on Dimethyl Ether Autoignition at High Pressures. Energy Fuels experimental and modeling study of propene oxidation. Part 2: Ignition delay time
2019;33(10):10197–208. and flame speed measurements. Combust Flame 2015;162(2):296–314.
[29] Liu D, Santner J, Togbé C, Felsmann D, Koppmann J, Lackner A, et al. Flame [53] G.P. Smith, Y. Tao, H. Wang, Foundational Fuel Chemistry Model Version 1.0
structure and kinetic studies of carbon dioxide-diluted dimethyl ether flames at (FFCM-1), 2016; http: //nanoenergy.stanford.edu/ffcm1/.
reduced and elevated pressures. Combust Flame 2013;160(12):2654–68. [54] Metcalfe WK, Burke SM, Ahmed SS, Curran HJ. A hierarchical and comparative
[30] Xia W, Peng C, Zou C, Liu Y, Lu L, Luo J, et al. Shock tube and modeling study of kinetic modeling study of C1–C2 hydrocarbon and oxygenated fuels. Int J Chem
ignition delay times of propane under O2/CO2/Ar atmosphere. Combust Flame Kinet 2013;45(10):638–75.
2020;220:34–48.

13

You might also like