Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Fuel 140 (2015) 626–632

Contents lists available at ScienceDirect

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

Experimental and kinetic study on ignition delay times of dimethyl


carbonate at high temperature
Erjiang Hu ⇑, Yizhen Chen, Zihang Zhang, Lun Pan, Qianqian Li, Yu Cheng, Zuohua Huang ⇑
State Key Laboratory of Multiphase Flow in Power Engineering, Xi’an Jiaotong University, Xi’an 710049, People’s Republic of China

h i g h l i g h t s

 Ignition delay times of dimethyl carbonate were measured in a shock tube for the first time at different conditions.
 A modified DMC kinetic model was proposed and can well predict ignition delay times and activation energies.
 DMC is primarily consumed through the H-abstractions and not fuel unimolecular decompositions for high temperature ignition.

a r t i c l e i n f o a b s t r a c t

Article history: Ignition delay times of dimethyl carbonate were measured in a shock tube for the first time at T = 1100–
Received 17 July 2014 1600 K, p = 0.12–1.0 MPa, fuel concentration = 0.5–2.0%, and / = 0.5–2.0. A modified chemical kinetic
Received in revised form 30 September model was developed and can well predict ignition delay times and activation energies. Further
2014
validation of the proposed kinetic model was made on the basis of the opposed flow diffusion flame data.
Accepted 3 October 2014
Available online 18 October 2014
Reasonable agreements were also achieved under the literature data. Reaction pathway analysis shows
that at high temperature the DMC molecule is primarily consumed through the H-abstractions but not
fuel unimolecular decompositions. Sensitivity analysis provides some key fuel-species reactions for
Keywords:
Dimethyl carbonate
DMC high temperature ignition.
Shock tube Ó 2014 Elsevier Ltd. All rights reserved.
Ignition delay time
Chemical kinetic model

1. Introduction Little work has been published regarding fundamental combus-


tion researches of DMC in recent years. Sinha and Thomson [9]
It has been proven that oxygenated fuels are effective in reduc- measured the species and temperatures in an opposed flow diffu-
ing soot emissions from diesel engines. Researches concerning sion flames of iso-propanol, dimethoxy methane (DMM), and
engine performance and emissions have been conducted widely dimethyl carbonate (DMC). For DMM and DMC, the absence of
on diesel engines fueled with various oxygenates, such as alcohols, C–C bonds effectively decreases the formation of soot precursors
ethers, esters, carbonates and acetates [1–4]. As a promising oxy- such as ethylene, acetylene, and propylene. Compared to the pro-
genated fuel, Dimethyl carbonate (DMC) possesses a very high per- pane flame, the ethylene levels of iso-propanol, DMM, and DMC
centage of oxygen. Engine studies indicated that the DMC addition flames were reduced by 41%, 77%, and 93%, respectively. However,
to the diesel could efficaciously decrease the smoke emissions their work did not cover the development of chemical kinetic
[5–7]. Previous studies revealed that the addition of ethers was model. Sinha and Thomson [9] studied the experimental species
more efficient in reducing the smoke emissions compared to the profiles by experimenting in an opposed flow diffusion flame and
addition of alcohols [8], and that dimethyl carbonate was even it gave the exclusive validation of the only available DMC kinetic
more efficient than other ethers [7]. However, there are numerous model [10], which is developed by Glaude and can well simulate
simultaneous processes in the engine which make it difficult to the DMM and DMC diffusion flame results. This model was also
determine the mechanism for such decrease. compared with species profiles obtained in premixed low-pressure
(30 Torr) flames of heptane with DMC addition, and an overall
satisfactory agreement was attained by Chen et al. [11]. Badin
measured the laminar flame speed using the heat flux method,
⇑ Corresponding authors. Tel.: +86 29 82665075; fax: +86 29 82668789. but Glaude DMC model significantly overpredicted the experimen-
E-mail addresses: hujiang@mail.xjtu.edu.cn (E. Hu), zhhuang@mail.xjtu.edu.cn tal values [12].
(Z. Huang).

http://dx.doi.org/10.1016/j.fuel.2014.10.013
0016-2361/Ó 2014 Elsevier Ltd. All rights reserved.
E. Hu et al. / Fuel 140 (2015) 626–632 627

In fact, many reaction rate constants in this DMC model were CH3O-CH3 (351.5 kJ/mol). Therefore, in Glaude DMC model [10],
determined by estimation. Further development and validation of it is improper to adopt the rate constant of CH3O-CH3 decomposi-
the DMC kinetic model are lagging due to the lack of more accurate tion reaction as that of R1565 and R1566.
fundamental experimental data. For instance, there are not Bond energies of CH3OC⁄OO-CH3 (374.5 kJ/mol) and CH3OC⁄O-
reported ignition delay time data yet. OCH3 (427.2 kJ/mol) in DMC are almost identical to those of
Therefore, the objectives of the present work include the mea- C3H7-C⁄OO-CH3 (364.1 kJ/mol) and C3H7C⁄O-OCH3 (423.9 kJ/mol)
surement the ignition delay time of DMC under different condi- in methyl butanoate. On the basis of analogical method, the
tions, the validation of the DMC chemical kinetic model, and the reaction rates of these two bond fission should have the same
interpretation of the ignition chemistry through reaction pathway values at the same temperature. As a result, the rate constants of
and sensitivity analysis. CH3OC⁄OO-CH3 (R1565) and CH3OC⁄O-OCH3 (R1566) decomposi-
tion, recommended by Dooley et al. [19] for the decompositions
of methyl butanoate, were adopted in the present model.
2. Experimental and calculated approach
It is reported by Glaude [10] that R1567 is a new molecular
elimination path for the DMC and the activation energy of this
Ignition delay times were measured in a shock tube and the
elementary reaction was determined by CBS-Q plus corrections
detailed description of the experimental apparatus can be referred
with isodesmic reactions. The present model employed the rate
to the previous publications [13,14]. The time interval between the
constant of R1567 of Glaude’s calculation. R1568 is another DMC
arrival of incident shock wave at the endwall and the intercept of
unimolecular decomposition reaction, and its rate constant was
the maximum slope of the CH⁄ trajectory with the zero line is
determined by Glaude’s estimation [10,20].
defined as the measured ignition delay time (s) in the current
study. Ignition temperatures (T) are calculated with Gaseq [15].
3.2. Hydrogen abstraction
The tested ranges of the equivalence ratio and the pressure of
DMC/oxygen/argon mixtures were 0.5–2.0 and 0.12–1.0 MPa,
It was proved that H-abstraction is the main pathway for the con-
respectively. The respective purities of DMC, oxygen and argon
sumption of typical fuels at high temperature ignition [14,25,26]. In
are 99.9%, 99.999% and 99.999%. Table 1 shows the compositions
Glaude DMC model, the rate constants of H-abstraction by radicals
of test mixtures in detail.
(R1573, R1575 and R1577) are assumed to that of the secondary
Calculations were carried out using Chemkin [16] and Senkin
C–H of iso-octane. Analogism can be made between the other
[17] codes and applying the constant volume adiabatic model.
H-abstraction reactions and the primary and secondary C–H of
The maximum increase rate of temperature profile (max dT/dt) is
n-heptane. Fig. 2 gives the bond dissociation energies in Glaude
used to define the calculated ignition delay time in this study.
DMC model of n-heptane and iso-octane at 298 K [24]. The bond
dissociation energy of C–H in DMC molecule (422.6 kJ/mol) is
3. Chemical kinetic model different to that of n-heptane (367.8 kJ/mol and 409.6 kJ/mol)
and iso-octane (399.6 kJ/mol), but it is close to that of primary
The unique DMC chemical model available at present was C–H (414.2 kJ/mol) in methyl butanoate. Therefore, the present
developed by Glaude in 2005 [10]. It consists of the C1–C3 model model assumed the rate constants of H-abstraction reactions of
developed by Curran and the DMC sub-model developed by Gla- DMC to be identical to that of methyl butanoate.
ude. Although this model was validated with DMC opposed-flow
diffusion flames [10], it cannot give a satisfactory prediction to 3.3. Ether-acid conversion
the experimentally measured ignition delay times of this study.
The modified chemical kinetic model in this study was R1579 is another consumption pathway of DMC molecule
enhanced by adding the DMC sub-model to Aramco Mech 1.3 producing the COC⁄OOH. Since this reaction has small branching
model [18] and it consists of 275 species and 1586 reactions as a ratio and consequently has little effect on the ignition delay time,
whole. DMC sub-model consists of 24 elementary reactions the rate constant of this elementary reaction was also adopted
absorbing the Glaude DMC sub-model [10] and Dooley MB sub- by Glaude’s estimation [10].
model [19]. Detailed mechanism including thermochemical data
is given in Supporting Information. It can be observed that the 3.4. Hydrogen abstraction of COC⁄OOH
modification was mainly conducted on the DMC sub-model, while
the original C4 chemistry remained the same. Table 2 [10,19–23] COC⁄OOH is mainly consumed by hydrogen abstraction with
provides further details of the modified DMC sub-model. small radicals (R1580–R1583). These reactions are also taken from
the Glaude’s estimation [10].
3.1. Unimolecular decomposition
3.5. Radical decomposition
Fig. 1 gives the bond dissociation energies of dimethyl carbon-
ate, methyl butanoate and dimethyl ether [24]. The bond energies CJOC⁄OOH radical is the product of hydrogen abstraction of
of CH3OC⁄OO-CH3 (374.5 kJ/mol) and CH3OC⁄O-OCH3 (427.2 kJ/ COC⁄OOH, and it decomposes to CH2O, CO and OH through
mol) in DMC are different and they are also unequal to that of R1584. COC⁄OOJ radical, the product of R1565, produces CH3O
and CO2. COC⁄COOCJ radical is the main product of DMC molecule
Table 1 consumption, decomposing CH3OCO and CO2. These three rate
Composition of DMC–O2–Ar mixtures. constants (R1584–R1586) were assumed by Glaude [10].
Mixtures / XDMC (%) XO2 (%) XAr (%) p (MPa) Decomposition of the CH3OCO radical has been the subject of
6 0.5 1.0 6.0 93.0 0.12, 0.5, 1.0
recent attention [10,27]. In this model, we employed Glaude’s
7 1.0 1.0 3.0 96.0 0.12, 0.5, 1.0 expression of CH3OCO decomposition calculated by CBS-Q method,
8 2.0 1.0 1.5 97.5 0.12, 0.5, 1.0 which has been validated with speciation data of an opposed flow
9 1.0 0.5 1.5 98.0 0.5 diffusion flame in his modeling study. The model prediction indi-
10 1.0 1.5 4.5 94.0 0.5
cates the importance of the decomposition of CH3OCO, because
11 1.0 2.0 6.0 92.0 0.5
its productions – CO2 and CH3 radicals – dominate free radical
628 E. Hu et al. / Fuel 140 (2015) 626–632

Table 2
DMC sub-model.

Num Reaction Rate constant References


Unimolecular decomposition
R1565 COC⁄OOC(+M) , COC⁄OOJ + CH3(+M) 2.552E+23–1.99 8.810E+04 [19]
LOW/1.744E+73–1.596E+01 8.532E+04/
TROE/2.18E-01 1.00E+00 6.37E+03 8.21E+09/
R1566 COC⁄OOC(+M) , CH3OCO + CH3O(+M) 7.426E+21–1.38 9.890E+04 [19]
LOW/1.807E+58–1.170E+01 9.127E+04/
TROE/2.51E-01 3.72E+02 9.48E+09 5.00E+09 /
R1567 COC⁄OOC , CH3OCH3 + CO2 5.0E+11 0.19 69800.0 [10]
R1568 COC⁄OOCJ + H , COC⁄OOC 5.0E+13 0.0 0.0 [10,20]
Hydrogen abstraction of DMC
R1569 COC⁄OOC + C2H3 , C2H4 + COC⁄OOCJ 5.010E+11 0.00 1.800E+04 [19]
R1570 COC⁄OOC + C2H5 , C2H6 + COC⁄OOCJ 5.010E+10 0.00 1.340E+04 [19]
R1571 COC⁄OOC + CH3 , CH4 + COC⁄OOCJ 2.265E+00 3.46 5.481E+03 [19]
R1572 COC⁄OOC + CH3O , CH3OH + COC⁄OOCJ 2.175E+11 0.00 4.571E+03 [19]
R1573 COC⁄OOC + CH3O2 , CH3O2H + COC⁄OOCJ 1.229E+04 2.60 1.391E+04 [19]
R1574 COC⁄OOC + H , H2 + COC⁄OOCJ 0.975E+06 2.40 4.471E+03 [19]
R1575 COC⁄OOC + HO2 , H2O2 + COC⁄OOCJ 1.229E+04 2.60 1.391E+04 [19]
R1576 COC⁄OOC + O , OH + COC⁄OOCJ 8.280E+05 2.45 2.830E+03 [19]
R1577 COC⁄OOC + O2 , HO2 + COC⁄OOCJ 3.000E+13 0.00 4.964E+04 [19]
R1578 COC⁄OOC + OH , H2O + COC⁄OOCJ 7.020E+07 1.61–3.500E+01 [19]
Ether-acid conversion
R1579 COC⁄OOC + H ) COC⁄OOH + CH3 3.79E+16–1.39 5402.0 [10,21]
Hydrogen abstraction of COC⁄OOH
R1580 COC⁄OOH + OH , CJOC⁄OOH + H2O 5.25E+09 0.97 1590.0 [10,22]
R1581 COC⁄OOH + H , CJOC⁄OOH + H2 9.40E+04 2.75 6280.0 [10,22]
R1582 COC⁄OOH + CH3 , CJOC⁄OOH + CH4 4.52E-01 3.65 7154.0 [10,22]
R1583 COC⁄OOH + O , CJOC⁄OOH + OH 9.65E+04 2.68 3716.0 [10,22]
Radical decomposition
R1584 CJOC⁄OOH ) CH2O + CO + OH 6.10E+21–2.40 3.252E4 [10,22]
R1585 CH3O + CO2 , COC⁄OOJ 1.000E+11 0.00 9.20E+03 [10,23]
R1586 CH3OCO + CH2O , COC⁄OOCJ 1.06E+11 0.0 7350.0 [10,20]
R488 CH3 + CO2 , CH3OCO 4.760E+007 1.540 34700.0 [10]
R489 CH3O + CO , CH3OCO 1.550E+006 2.016 5730.0 [10]

H CH3 H H H
H H
399.6

H C C C C C H
H C O 351.5 C 401.7 H

H CH3 H CH3 H
H H

H H
H H H H H H H
O
367.8 409.6
H C O C 427.2 O 374.5 C 422.6 H H C C C C C H
C C

H H H H H H H H H

H H H H Fig. 2. Bond dissociation energies (BDE298, kJ/mol) of n-heptane and i-octane [24].
O

H C C C C 423.9 O 364.1 C 414.2 H esters model developed by Pascal [28], were also taken from the
Glaude’s calculation.

H H H H
4. Results and discussion
Fig. 1. Bond dissociation energies (BDE298, kJ/mol) of dimethyl carbonate, methyl
butanoate and dimethyl ether [24].
4.1. Ignition delay time measurement

Fig. 3 gives the measured DMC ignition data. The logarithmic


chain branching and termination in most of sensitive reactions in ignition delay times under all conditions exhibit good linear
the system. Moreover, the rate constants of R488 and R489 in Ara- dependence upon 1000/T. Therefore, through regression, a fitting
mco Mech 1.3 model developed by Metcalfe et al. [18], the methyl correlation of DMC ignition delay times as a function of p, XDMC,
butanoate model developed by Dooley et al. [19] and the methyl / and T for the DMC–O2–Ar mixtures is correlated as follows:
E. Hu et al. / Fuel 140 (2015) 626–632 629

Fig. 3. Measured and fitted ignition delay times of DMC. (Symbols: measured values; lines: fitted values using Eq. (1)).

!
1
4 0:38 162:59  1:68 kJ mol activation energy. However, the overprediction of Glaude model
s ¼ 1:11  10 p X 0:71
DMC /
0:65
exp ; in ignition delay times of DMC makes it incompetent for quantita-
RT
tive simulation, while the modified model can well predict the
R2 ¼ 0:987 ð1Þ ignition delay times and activation energies of DMC across the
whole temperature range.
where s, p and / are ignition delay time in ls, pressure in MPa and
As shown in Fig. 4b, the ignition delay times are dependent on
equivalence ratio, respectively. R is the universal gas constant and
pressure as well besides the fuel mole fraction. Ignition delay times
its value is 8.314 J/mol K, and T is the temperature in K. R2 = 0.987
decrease with the increase of pressure. The reactant concentrations
indicate that Eq. (1) has good regressions to the ignition data. This
increase significantly at high pressures, which promotes the igni-
correlation is only applicable to the temperature range of 1100–
tion process. The negative pressure exponent in the ignition delay
1550 K, pressure range of 0.12–1.0 MPa, equivalence ratio range of
correlation shown in Eq. (1) also suggests this phenomenon.
0.5–2.0 and fuel mole fraction range of 0.5–2.0% because of the lim-
Similarly, the modified model can well predict the pressure depen-
ited range of the study.
dence and gives good quantitative agreement with the experimen-
tal data at different pressures, while the Glaude model again gives
4.2. Validation of the kinetic model over-predictions on the ignition delay times of DMC.
In order to extend the application of the modified model, the
4.2.1. Ignition delay time effect of equivalence ratio on ignition delay times at low and high
In this section, we used the Glaude DMC model and modified pressures is given in Fig. 4c and d. The same trend is presented at
model to calculate the ignition delay times in order to compare 0.12 MPa and 1.0 MPa, showing that the ignition delay times
and validate their respective performances under different fuel decrease with decreasing equivalence ratio. The reason for this is
mole fractions, pressures, and equivalence ratios. that, at high temperatures, the chain branching reaction (H + O2 =
Fig. 4 shows the measured ignition delay times and the simu- OH + O) plays a dominant role in the ignition process. Higher oxy-
lated ones for DMC with modified model and Glaude model. The gen concentration (lower equivalence ratio) favors to the increase
kinetic models were validated at different fuel mole fractions. As of chain branching, resulting increased reactivity and decreased
a matter of fact, the fuel mole fraction in our study has an equiva- ignition delay times. Compared to the Glaude DMC model, the
lent effect as the dilution ratio used by other researchers [29]. modified model can capture the dependence of equivalence ratio
Fig. 4a gives the measured and calculated ignition delay times at and gives good quantitative prediction to the experimental values.
different DMC mole fractions. The ignition delay times decrease
with the increase of fuel fraction. The reason is that high fuel and
oxygen concentrations enhance the overall reaction rate. Accord- 4.2.2. Opposed-flow diffusion flame data
ing to the figure, both Glaude model and modified model are capa- Since Glaude model was developed using mole fractions in an
ble of capturing the trend of the curves and predicting the opposed flow diffusion flame of DMC, a comparison between
630 E. Hu et al. / Fuel 140 (2015) 626–632

Fig. 4. Measured and calculated ignition delay times with modified model and Glaude model for DMC. (Symbols: measured values; lines: calculated values.)

calculated and measured diffusion flame data of DMC was made to 4.3. Reaction pathway and sensitivity analysis
further validate the present kinetic model. The DMC diffusion
flame experiments were conducted at p = 0.1 MPa [9]. The oxidizer 4.3.1. Reaction pathway analysis
stream containing 39% O2 and 61% N2 was sent through the top Reaction pathway analysis is an effective method in determin-
burner port; 8% fuel (DMC) and 92% N2 was sent through the bot- ing main reaction pathways for the concerned species. Fig. 6 shows
tom. Fuel flow line and burner port were heated to 45 °C to avoid the reaction pathways using the present model at T = 1350 K,
any condensation. Validations against experimental diffusion p = 0.12 MPa and / = 1.0 for DMC ignition. We chose the timing
flame data for the Glaude model and present model are given in of 20% fuel consumption for the analysis like other works
Fig. 5. The present model almost gives the same calculated values [30,31]. It can be observed that DMC molecule is primarily con-
on DMC diffusion flame data as the Glaude model, which indicates sumed through the H-abstractions (their branching ratios are
that the present model can well predict not only the ignition delay 79.2%), while the unimolecular decomposition gives small
times but also the diffusion flame data of DMC. contribution (20.8%). Obviously, the unimolecular decomposition

Fig. 5. Measured and calculated mole fraction profiles of major species and temperature in the DMC opposed-flow diffusion flame. (Symbols: measured values [9]; lines:
calculated values.)
E. Hu et al. / Fuel 140 (2015) 626–632 631

COC*OOC 5. Conclusions
R1565 R1571 R1579
R1574 77.5% 1.7% Experimental and kinetics study on ignition delay times of DMC
20.8% R1576 were conducted in a shock tube at different equivalence ratios
R1578
(0.5–2.0), pressures (0.12–1.0 MPa) and fuel mole fractions (0.5–
COC*OOJ COC*OOH
COC*OOCJ 2.0%) at high temperatures. Main conclusions are summarized as
R1585 R1580 follows:
R1586 100%
-1583
100% 100%
(1) Ignition delay times of DMC were measured for the first time
CH3O CJOC*OOH in a shock tube. Similar to typical hydrocarbon fuels, ignition
CH3OCO
delay times are decreased with the increase of pressure, the
R488 R489 R1584 100% decrease of equivalence ratio and the increase of fuel con-
59.4% 40.6%
centration. Correlation of ignition delay time is given by
multiple linear regression on experimental data.
CH3 CH3O CH2O (2) A modified chemical kinetic model for DMC is proposed. The
model yields fairly good agreement with the measured igni-
Fig. 6. Reaction pathway of present model for DMC in shock tube at p = 0.12 MPa, /
tion delay times. The present model was validated against
= 1.0, T = 1300 K, 20% fuel consumption.
the diffusion flame data.
(3) Reaction pathway analysis shows that DMC molecule is pri-
reaction R1565 dominates the fuel breakdown, while the branch- marily consumed through the H-abstractions. Sensitivity
ing ratio of R1566 is negligible. This is because the bond energies analysis reveals the importance of small radical reactions,
of CH3OC⁄OO-CH3 (374.5 kJ/mol) is much smaller than that of CH3- especially the main chain branching reaction H + O2 =
OC⁄O-OCH3 (427.2 kJ/mol) as shown in Fig. 1. O + OH at the tested temperature range. Some fuel-specific
reactions are also found to have relatively large sensitivity
coefficients.
4.3.2. Sensitivity analysis
To identify the above variation and clarify the elementary reac-
tions that dominate the ignition chemistry, sensitivity analysis
(Sensitivity coefficient, S ¼ sð2:0k1:5i Þsðksð0:5kiÞ
, here s is ignition delay Acknowledgments

time and ki is pre-exponential factor of ith reaction). was con-


This study is supported by the National Natural Science Founda-
ducted in this study. Positive sensitivity coefficients indicate an
tion of China (51306144), the National Basic Research Program
inhibiting influence on the ignition process, and vice versa. Fig. 7
(2013CB228406) and the State Key Laboratory of Engines at Tianjin
gives the key sensitive reactions for the DMC at T = 1350 K,
University (SKLE201302). The support from the Fundamental
p = 0.12 MPa and / = 1.0.
Research Funds for the Central Universities is also appreciated.
For DMC as shown in Fig. 7, chain branching reaction (R1) domi-
nants the ignition chemistry. Small radical reactions have a
remarkable influence on ignition delay times, and only two of Appendix A. Supplementary material
the twelve most important reactions are fuel-related reactions.
Fuel-related reaction R1574, as an H-abstraction reaction by H rad- Supplementary data associated with this article can be found, in
ical, has the largest positive value of S at high temperature. This the online version, at http://dx.doi.org/10.1016/j.fuel.2014.10.013.
suggests that the ignition process can be accelerated by the
increase of the reaction rate. The fuel-related reaction R1565, as
an important unimolecular decomposition reaction, gives the sec- References
ond largest negative sensitivity coefficient because of its low bond
[1] Ren Y, Huang ZH, Miao HY, Di YG, Jiang DM, Zeng K, et al. Combustion and
dissociation energy. Moreover, R488 and R489 are also found to be emissions of a DI diesel engine fuelled with diesel-oxygenate blends. Fuel
important for the DMC ignition. 2008;87(12):2691–7.
[2] Wang XG, Cheung CS, Di YG, Huang ZH. Diesel engine gaseous and particle
emissions fueled with diesel-oxygenate blends. Fuel 2012;94:317–23.
[3] Wen LB, Xin CY, Yang SC. The effect of adding dimethyl carbonate (DMC) and
ethanol to unleaded gasoline on exhaust emission. Appl Energy
2010;87(1):115–21.
[4] Zhu RJ, Miao HY, Wang XB, Huang ZH. Effects of fuel constituents and injection
timing on combustion and emission characteristics of a compression-ignition
engine fueled with diesel-DMM blends. Proc Combust Inst
2013;34(2):3013–20.
[5] Maricq MM, Chase RE, Podsiadlik DH, Seigl WO, Kaiser EW. The effect of
dimethoxy methane additive on diesel vehicle particulate emissions. SAE
technical paper 982572; 1998.
[6] Murayama T, Zeng M, Chikahisa T, Arima T. Simultaneous reductions of smoke
and NOx from a DI diesel engine with EGR and dimethyl carbonate. SAE
technical paper 952518; 1995.
[7] Miyamoto N, Ogawa H, Arima T. Improvement of diesel combustion and
emissions with addition of various oxygenated agents to diesel fuels. SAE
technical paper 962115; 1996.
[8] Liotta FJ, Montalvo DM. The effect of oxygenated fuels on emissions from a
modern heavy-duty diesel engine. SAE technical paper 932734; 1993.
[9] Sinha A, Thomson MJ. The chemical structures of opposed flow diffusion
flames of C3 oxygenated hydrocarbons (isopropanol, dimethoxy methane, and
dimethyl carbonate) and their mixtures. Combust Flame 2004;136(4):548–56.
[10] Glaude PA, Pitz WJ, Thomson MJ. Chemical kinetic modeling of dimethyl
Fig. 7. Sensitivity analysis for DMC in shock tube at T = 1350 K, p = 0.12 MPa carbonate in an opposed-flow diffusion flame. Proc Combust Inst
and / = 1.0. 2005;30(1):1111–8.
632 E. Hu et al. / Fuel 140 (2015) 626–632

[11] Chen G, Yu W, Fu J, Mo J, Huang ZH, Yang JC, et al. Experimental and modeling [22] Curran HJ, Gaffuri P, Pitz WJ, Westbrook CK. A comprehensive modeling study
study of the effects of adding oxygenated fuels to premixed n-heptane flames. of n-heptane oxidation. Combust Flame 1998;114(1–2):149–77.
Combust Flame 2012;159:2324–35. [23] Fisher EM, Pitz WJ, Curran HJ, Westbrook CK. Detailed chemical kinetic
[12] Bardin ME, Ivanov EV, Nilsson EJK, Vinokurov VA, Konnov AA. Laminar burning mechanisms for combustion of oxygenated fuels. Proc Combust Inst
velocities of dimethyl carbonate with air. Energy Fuels 2013;27:5513–7. 2000;28(2):1579–86.
[13] Zhang JX, Niu SD, Zhang YJ, Tang CL, Jiang X, Hu EJ, et al. Experimental and [24] Luo YR. Comprehensive handbook of chemical bond energies. CRC Press;
modeling study of the auto-ignition of n-heptane/n-butanol mixtures. 2007.
Combust Flame 2013;160(1):31–9. [25] Zhang JX, Wei LJ, Man XJ, Jiang X, Zhang YJ, Hu EJ, et al. Experimental and
[14] Hu EJ, Jiang X, Huang ZH, Zhang JX, Zhang ZH, Man XJ. Experimental and modeling study of n-butanol oxidation at high temperature. Energy Fuels
kinetic studies on ignition delay times of dimethyl ether/n-butane/O2/ar 2012;26(6):3368–80.
mixtures. Energy Fuels 2012;27(1):530–6. [26] Man XJ, Tang CL, Zhang JX, Zhang YJ, Pan L, Huang ZH, et al. An experimental
[15] Morley C. Gaseq v0.76. <http://www.gaseq.co.uk>. and kinetic modeling study of n-propanol and i-propanol ignition at high
[16] Kee RJ, Rupley FM, Miller JA. Chemkin-II: A Fortran chemical kinetics package temperatures. Combust Flame 2014;161(3):644–56.
for the analysis of gas-phase chemical kinetics. Report no. SAND89-8009. [27] McCunn LR, Lau KC, Krisch MJ, Butler LJ, Tsung JW, Lin JJ. Unimolecular
Sandia National Laboratories; 1989. dissociation of the CH3OCO radical: an intermediate in the CH3O + CO
[17] Lutz AE, Kee RJ, Miller JA. Senkin: A Fortran program for predicting reaction. J Phys Chem A 2005;110(4):1625–34.
homogeneous gas phase chemical kinetics with sensitivity analysis. Report [28] Diévart P, Won SH, Gong J, Dooley S, Ju YG. A comparative study of the
no. SAND87-8248. Sandia National Laboratories; 1988. chemical kinetic characteristics of small methyl esters in diffusion flame
[18] Metcalfe WK, Burke SM, Ahmed SS, Curran HJ. A hierarchical and comparative extinction. Proc Combust Inst 2013;34(1):821–9.
kinetic modeling study of C1–C2 hydrocarbon and oxygenated fuels. Int J [29] Akih-Kumgeh B, Bergthorson JM. Experimental and modeling study of trends
Chem Kinet 2013;45(10):638–75. in the high-temperature ignition of methyl and ethyl esters. Energy Fuels
[19] Dooley S, Curran HJ, Simmie JM. Autoignition measurements and a validated 2011;25(10):4345–56.
kinetic model for the biodiesel surrogate, methyl butanoate. Combust Flame [30] Black G, Curran HJ, Pichon S, Simmie JM, Zhukov V. Bio-butanol: combustion
2008;153(1–2):2–32. properties and detailed chemical kinetic model. Combust Flame
[20] Curran HJ, Gaffuri P, Pitz WJ, Westbrook CK. A comprehensive modeling study 2010;157(2):363–73.
of iso-octane oxidation. Combust Flame 2002;129(3):253–80. [31] Yasunaga K, Mikajiri T, Sarathy SM, Koike T, Gillespie F, Nagy T, et al. A shock
[21] Held TJ, Dryer FL. A comprehensive mechanism for methanol oxidation. Int J tube and chemical kinetic modeling study of the pyrolysis and oxidation of
Chem Kinet 1998;30(11):805–30. butanols. Combust Flame 2012;159(6):2009–27.

You might also like