Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Combustion and Flame 246 (2022) 112411

Contents lists available at ScienceDirect

Combustion and Flame


journal homepage: www.elsevier.com/locate/combustflame

Laminar burning velocities and nitric oxide formation in premixed


dimethyl ether/air flames: Experiments and kinetic modeling
Marco Lubrano Lavadera∗, Christian Brackmann, Alexander A. Konnov
Division of Combustion Physics, Lund University, P.O. Box 118, Lund SE-22100, Sweden

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

Article history: Adiabatic laminar burning velocities and post-flame NO mole fractions for flat, non-stretched, premixed
Received 27 April 2022 dimethyl ether/air flames were experimentally determined with a heat flux burner combined with laser-
Revised 21 September 2022
induced fluorescence diagnostics, over equivalence ratios ranging from 0.7 to 1.6, at atmospheric pressure
Accepted 22 September 2022
and initial temperatures from 298 to 338 K. The present burning velocity measurements were then com-
Available online 8 October 2022
pared with selected data available in the literature obtained with different techniques. The comparison
Keywords: showed reasonably good agreement with recent datasets obtained at different temperatures, as well as
Dimethyl ether possible outliers not suitable for the validation of kinetic models. The detailed kinetic mechanism of
Laminar burning velocity the authors was extended by the reactions of dimethyl ether. A comparison of experimental and com-
Nitric oxide putational results using two contemporary detailed chemical kinetic mechanisms, along with the one
Heat flux method from the authors presented in this work, was also conducted and discussed. Discrepancies between ex-
LIF
periments and model predictions and among models themselves were observed, especially under rich
Detailed kinetic mechanism
conditions. Further numerical analyzes were performed to identify the main causes of the observed dif-
ferences. Notwithstanding, the present model showed overall good performances in reproducing both
laminar burning velocities and nitric oxide mole fractions. Kinetic modeling was also performed to en-
hance the understanding of the intrinsic NOx emission characteristics of dimethyl ether by comparing the
present measurements with those obtained for ethanol/air flames under identical conditions.
© 2022 The Author(s). Published by Elsevier Inc. on behalf of The Combustion Institute.
This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/)

1. Introduction Any kinetic combustion mechanism, to be successful, must be


able to predict, among other fundamental combustion character-
With the recent advancements toward the development of istics, measured laminar burning velocities (SL ) with reasonable
cleaner and more efficient combustion processes, dimethyl ether agreement. Hence, SL for DME has been measured in previous
(DME) has received considerable attention as a promising alterna- studies at various equivalence ratios (), inlet temperatures (Tg ),
tive fuel or fuel additive for conventional Diesel [1–18], homoge- and pressures. Table S1 in the Supplementary Material (SM) lists
neous [19,20] and premixed charge compression ignition engines all the studies available in the literature together with the ex-
[21], as well as land-based industrial gas turbines [22,23]. With the perimental conditions investigated. Measurements have been car-
rising interest in DME as a prospective fuel for various practical ap- ried out using a variety of methods such as Bunsen burners
plications, the availability of accurate and reliable kinetic models [25], constant-volume cylindrical/spherical vessels [26–40], single-
that can predict its combustion characteristics and pollutant emis- jet wall-stagnation flames [41], counterflow twin flames [42], or
sions has become essential for the optimal design of combustion externally heated diverging channels [43–46]. Unfortunately, al-
devices. For these reasons, the chemical kinetics of DME has been though an extensive number of experimental data for SL of DME
widely investigated in the last two decades, and numerous experi- mixtures exists in the literature, available datasets are inconsistent,
mental and theoretical studies on its oxidation and pyrolysis under as pointed out in the review by Konnov et al. [47]. This hampers
various conditions have been performed, which led to the develop- the validation of kinetic models. Therefore, due to the significance
ment of many detailed and skeletal/reduced chemical kinetic mod- of SL information, additional, independent measurements of high
els, as recently summarized by Stagni et al. [24]. fidelity are necessary.
Although there is no general consensus on the best approach to
measure SL , the flat-flame-based heat flux method is currently con-

Corresponding author. sidered highly accurate since it produces a flame approaching the
E-mail address: marco.lubrano_lavadera@forbrf.lth.se (M. Lubrano Lavadera). stationary, adiabatic, one-dimensional, and stretchless state. In fact,

https://doi.org/10.1016/j.combustflame.2022.112411
0010-2180/© 2022 The Author(s). Published by Elsevier Inc. on behalf of The Combustion Institute. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/)
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

as shown in several investigations, SL measured using this method bonds [66]. Similar findings and explanations were later reported
under the same conditions in different laboratories often overlap by Lee et al. [67], who experimentally investigated the NOx emis-
within the experimental uncertainties, which are usually on the sion characteristics of DME, LPG, and their blend using a laminar
order of ±1 cm/s. Nevertheless, heat flux method experiments for counterflow diffusion flame. This mechanism of NO reduction was
DME/air mixtures were not available when the review of Konnov also subsequently confirmed by Yamamoto and Kajimura [63], who
et al. [47] was published. Recently, Wang et al. [48] and Shrestha studied DME reburning under flow-reactor conditions. However,
et al. [40] measured SL of DME/air mixtures using the heat flux Marrodán et al. [61], when studying acetylene/DME/NO mixtures in
method. Hence, it is expected that the availability of such recent a flow reactor, found that the highest NO consumption under fuel-
measurements can provide additional scrutiny of the data consis- rich conditions was achieved in the absence of DME because the
tency. availability of O-atoms in DME promotes HCCO oxidation, which
Apart from SL , also the need to curtail pollutants in combus- competes with the reburning channel. Such an explanation is in
tion systems has stimulated the study of DME detailed chemistry. contrast with the previous findings [63,66,67].
Due to the promising utilization of DME in compression ignition To clarify these issues, one of the most effective approaches
engines, a considerable number of studies on pollutant emissions is to experimentally and numerically investigate NO formation in
have been focused on engine applications. All the experiments premixed flames, where different NO formation mechanisms can
have reported nearly zero soot emissions from engines fueled with be separated by changing . In this regard, Frye et al. [68] com-
DME. This was attributed to the high oxygen content in the fuel pared the relative NO emissions from DME, propane, and n-butane
molecule, the low carbon-to-hydrogen ratio, and the absence of laminar premixed conical flames over a broad range of stoichiome-
carbon-carbon bonds. On the other hand, it was found that DME tries at atmospheric pressure. NO production was analyzed by
may produce lower [1,2,8,11,14], comparable [3,9], or even higher Fourier-transform infrared spectroscopy through probe sampling
[3,4,9,10,12,15,17] nitrogen oxides (NOx ) emissions compared to above the flame. Consistent with the abovementioned studies, NO
diesel fuel operation and it appears that such emissions in diesel concentrations from DME flames were found to be lower, or at
engines are not necessarily related to the chemical properties of worst equal, to those from propane and n-butane flames over the
DME but rather to the engine operating conditions, fuel supply sys- same range of stoichiometries. However, the authors interpreted
tem, and engine specifications. In fact, in practical engine systems, the observed behavior only in terms of differences in the SL , thus
it is not possible to maintain identical operating conditions for dif- different residence times. Moreover, the results obtained in the
ferent fuels because the change in physical and chemical fuel prop- study of Frye et al. [68] cannot be used to validate kinetic mod-
erties needs to be compensated by adjusting engine parameters. els for several reasons. The main reason is that the probe height
This can significantly modify the combustion environment affect- was not held constant with respect to the burner, and no quanti-
ing NOx formation, such as combustion mode, local flame temper- tative information about the height was provided. Furthermore, as
ature, residence time in the high-temperature zone, and local  stated by the authors themselves, the experimental results were
[4–9,11–15,17,18,21]. Therefore, the intrinsic NOx formation char- affected by a number of perturbations, such as dilution of the
acteristics of DME compared to those of hydrocarbon fuels are not combustion products due to the probe sampling above the flame
as clear as in the case of soot. For example, Teng et al. [6] sug- zone, uncertainties in the measured flow rates, along with intrinsic
gested that the influence of the oxygen content in DME on the non-idealities associated with conical flames. Recently, Eckart et al.
NOx formation behavior is insignificant, in contrast to what was re- [69] investigated premixed polyoxymethylene flames at different
ported by other authors [16,17]. Hence, in order to reconcile these temperatures on a heat flux burner configuration and measured
results, fundamental experimental studies under well-defined con- NOx profiles for DME as a function of height above the burner
ditions, which can be used for comparison with model predictions, (HAB) using probe sampling. The experimental results were com-
are required. To date, most fundamental research on the pollutant pared with simulations using a detailed kinetic mechanism devel-
emissions from DME combustion has been focused on soot pre- oped by the authors, which was found to overestimate the mea-
cursors [49–56] or on fuel-NOx interactions at low temperature to sured NOx concentrations. The authors attributed the observed dis-
mimic exhaust gas recirculation [57–65], while very few studies crepancy to a combination of experimental and model uncertain-
have explored NOx formation at high temperature. In detail, Hwang ties. As a matter of fact, the analysis by Eckart et al. [69] was
et al. [66] experimentally investigated the NOx emission behavior limited to  = 0.8, 1.0, and 1.2, which in the case of premixed
of DME/air laminar co-axial jet diffusion flames and compared it to methane flames was shown [70] to be the  range where distur-
that of ethane to investigate the effect of the oxygen atom in the bances to the flow and temperature field in the flame by the in-
DME molecular structure. The nitric oxide (NO) emissions of DME vasive probe can be relevant. In this regard, laser-induced fluores-
were found to be lower in comparison with those of ethane for cence (LIF) coupled with premixed flat flames represents a viable
the same conditions. This was mainly attributed to (1) the shorter arrangement for testing the kinetics of nitrogen since spatially re-
flame length of DME, which decelerates the rate of thermal-NO for- solved, quantitative, sensitive measurements of NO can be carried
mation due to a reduction in the residence time/increase of SL, and out non-intrusively, and it is easy to model the 1-D configuration.
(2) the presence of the oxygen atom in the ether structure, which In view of the above considerations, the goal of the current
decreases the level of prompt-NO in the primary zone due to the study is threefold:
scavenging of hydrocarbon radicals. An auxiliary computational in-
vestigation adopting the counterflow configuration with detailed
chemistry and transport models was conducted to interpret the re- (1) To extend the existing experimental database on the SL of
sults but not to predict them since, as the authors stated, the ex- DME/air flames by new measurements using the heat flux
periments were carried out in a multi-dimensional flame. The nu- method and to compare obtained data with those available
merical results revealed that another cause of comparatively lower to evaluate data consistency.
NOx emissions from DME is the activation of a distinct NO reburn- (2) To provide, for the first time, accurate experimental data on
ing mechanism NO+HCCO=HCNO+CO in the fuel-rich region, not NO mole fractions for these flames using LIF.
apparent in the ethane flame. The authors stated that this behav- (3) To compare the present and the available experimental data
ior is due to the fast pyrolysis and oxidation reactions of DME in with those computed with mechanisms selected from the
the fuel-rich region, owing to the existence of the C–O bond in literature, as well as the one from the authors extended
the chemical structure of DME, which breaks more easily than C–H to DME in this work, to assess their performances. Kinetic

2
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

modeling was also performed to enhance the understanding ular, heat losses were calculated with an optically-thin radiation
of the intrinsic NOx emission characteristics of DME. model, which included Planck absorption coefficients of H2 O, CO,
CO2 , CH4 , NO, and N2 O. Simulations have pointed out that radia-
2. Experimental details tive heat losses have an important impact on the thermal NO pre-
dictions (up to 7%), but their effect on SL is negligible (from 0.01 to
Planar, adiabatic, non-stretched, premixed laminar DME/air 0.4%, depending on ). All the numerical solutions were obtained
flames at atmospheric pressure were stabilized on a perforated by setting the GRAD and CURV parameters to 0.03, resulting in
plate heat flux burner of improved design. The experimental setup 50 0–70 0 grid points over the domain of 3 cm. To shift the sim-
has been described elsewhere [71], thus only the relevant infor- ulated distances of the free flame to the corresponding HAB, the
mation is provided here. The improved flat-flame burner used in initial distance of HAB = 0 mm was fixed to the distance at which
the present work has been described in detail in [72]. It consists the surface plate temperature of 368 K was achieved.
of a brass plate integrated into the burner head, perforated with The main objective of the kinetic modeling was to evaluate the
0.4 mm-diameter holes at a pitch of 0.47 mm. The plate thickness, ability of current kinetic schemes in predicting SL and NO pro-
nominal diameter, and effective perforation area are respectively duction for DME/air flames. It should be noted that, as written in
2 mm, 3 cm, and 6.9004 cm2 . The burner plate edge is maintained the introduction, many kinetic mechanisms for DME combustion
at 368 K by means of a heating jacket with circulating water, while are available, but only a few of them incorporate the NOx sub-
Tg was fixed at 298, 308, 323, and 338 K (±1 K) by a separate set. Thus, simulations were carried out with two comprehensive
water jacket surrounding the plenum chamber. SL was determined mechanisms that incorporate NOx chemistry, namely the San Diego
by changing the velocity of the unburned gas until a uniform ra- [73] and CRECK [74] models.
dial temperature distribution over the burner plate, recorded by 15
type E thermocouples (0.076 mm bare wire diameter) soldered in 4. Detailed kinetic mechanism
the burner plate, was achieved. Pure DME (99.9%) and synthetic
air (21% O2 , 1% relative uncertainty) were supplied by Linde, and Another goal of this work was to present an updated detailed
their required flow rates were controlled by two Bronkhorst High- chemical kinetic reaction mechanism denoted here as the Kon-
Tech digital thermal mass flow controllers. Both flow meters were nov model. Hierarchically structured and built according to the
calibrated using a Ritter TG10 drum-type gas meter. The total un- first-principles approach, it is based on the C0 –C3 core mechanism
certainty of the flow rate is a sum of the 0.5% stated accuracy of from [75] together with the NH3 /NOx subsets recently updated in
the calibrator and the stated flow repeatability of the mass flow [76,77]. Additionally, as a result of a continuing effort to model
controllers, which corresponds to 0.2% of the set operating condi- the combustion of hydrocarbons up to C3 , the kinetic subset de-
tions. scribing pyrolysis and oxidation of DME, including thermodynamic
Detailed analysis and quantification of the experimental uncer- and transport properties, has been developed in the present work
tainties were reported earlier [71,72], and the overall accuracy of and added to the base model with no adjustments to the core
SL in the present measurements was estimated to be better than mechanism. Thermodynamic and transport properties of relevant
±0.4 cm/s.  was varied from 0.7 to 1.6 with increments of 0.1 and species were adopted from the AramcoMech 3.0 model [78], except
a maximum uncertainty of 0.016. Then, at each  and Tg = 338 K, for CH3 OCH3 (DME) and CH3 OCH2 for which the thermodynamic
the adiabatic flame was produced to measure NO. All experimental data were taken from the Burcat database [79], and for CH3 OCH2 O
data with associated uncertainties are tabulated in the SM. that were taken from the Goldsmith database [80]. In total, 75
The setup described in detail in [70] for LIF measurements reactions related to the DME subset were evaluated one by one,
of NO mole fraction (XNO ) was employed in this study. Excita- and the most reliable rate constants were implemented, avoiding
tion of NO was made in the A2  + ←X2  (0–0) vibronic band at adjustment or tuning of the parameters to accommodate experi-
225.5 nm in air using a combined Nd:YAG and dye laser system. mental results. The rate constants for DME decomposition and H-
The measurement volume, defined by a laser beam focused across abstraction reactions by H, O, OH, HO2 , O2 , CH3 , and CH3 O have
the burner center at HAB = 10 mm (±0.5), was imaged onto the been adopted from available experimental and theoretical studies
slit of a spectrometer. A photomultiplier tube at the spectrome- [81–88]. The rate constants of H-abstraction reactions by CH3 O2 ,
ter exit, where an additional slit was mounted, was used to de- C2 H5 O2 , CH3 OCH2 O2 , and HCOO have never been experimentally
tect the fluorescence signal in the (0–1) γ -band of the A2  + →X or theoretically determined. Following Hashemi et al. [89], 20% and
2  (0–0) NO transition at 236 nm. The pulse energy used (2–
16% of the rate constant for H-abstraction by HO2 is attributed to
2.3 mJ/pulse) together with the beam focusing allowed to conduct the abstractions by CH3 O2 and C2 H5 O2 , respectively. The rate con-
measurements with a laser irradiance under saturated conditions. stant of the H-abstraction reaction by CH3 OCH2 O2 was estimated
The photomultiplier signal was registered by a digital oscilloscope by Fischer et al. [90] and has since then been implemented in
and averaged over 128 laser pulses. In addition, data recorded with many DME models. Hashemi et al. [89] suggested this rate to be
the laser tuned to 225.38 nm, off any NO resonance, were system- double the rate of reaction CH3 O2 +CH3 OH. In the present work,
atically subtracted from the corresponding measurements made it is assumed equal to that of the abstraction reaction by C2 H5 O2 .
on the Q2 (26.5) NO resonance. The resulting peak value was con- For reaction with HCOO, Fischer et al. [90] suggested a rate con-
verted into XNO using the calibration methodology described in stant twice that of the reaction with CH3 OCH2 O2 . In the present
[70], where the signal was measured in a fuel-lean syngas flame model, it is assumed equal to that of reaction with CH3 O2 . The re-
seeded with different levels (below 100 ppm) of NO. According to maining rate constants related to DME and methyl formate subsets
[70], the experimental uncertainty of NO quantification is within are mainly adopted from the Glarborg model [89], but rate con-
8.7%. stants for some reactions have been selected from other sources
[90–95] if considered more reliable. Besides methyl formate, formic
3. Modeling details acid is also an important intermediate in DME oxidation at low
temperatures. The sub-mechanism describing its oxidation was al-
The freely-propagating Premix laminar flame-speed code of the ready implemented in the model based on the experimental and
ANSYS Chemkin 17.0® software was used to simulate the flames. modeling study presented earlier [96], and no modifications have
The Soret effect, downstream radiative heat losses, and multicom- been applied to it in the present work. The choice of the rate con-
ponent diffusion were all included in the computations. In partic- stants not mentioned here is presented in Section 2 of the SM. In

3
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

addition, Table S2 in the SM lists the 75 reactions added to the


mechanism in this work, along with selected rate coefficients and
their sources. When pressure-dependent rate constants are pro-
vided in PLOG format, only the term at 1 atm is shown. Extensive
testing of the present model is reported in the SM (Figs. S1–S21)
and includes speciation and oscillation data in Jet Stirred [93,97,98]
and Plug Flow [90,99] Reactors, ignition delay times in shock tubes
[91,100–107], rapid compression machines [91,94,108], and flow
reactors [109], as well as SL at high pressure [27,31,40]. The full ki-
netic model consists of 3289 elementary reactions and 268 chem-
ical species, and is provided in Chemkin format in the SM.

5. Results and discussion

5.1. Laminar burning velocity

As mentioned in the introduction, a large number of experi-


mental data on the SL of DME/air flames is available in the liter-
ature. However, Konnov et al. [47] found a notable scatter among
literature datasets, especially at the extreme sides of stoichiometry. Fig. 1. Experimental (symbols) and numerical (lines) SL versus  at Tg = 298 K and
p = 1 atm for DME/air flames. Black symbols: heat flux method (squares present
In particular, even considering only stretch-corrected data, the dis-
work, circles [48], triangles [40]). Blue squares: outwardly propagating spherical
agreement in the experimental SL was found to be within 20% at flames with non-linear extrapolation [40]. Red symbols: outwardly propagating
 = 1 and reached about 60% and 40% under very rich or very lean spherical flames with linear extrapolation (squares [31], circles [35], up triangles
conditions, respectively. Furthermore, differences in the  corre- [36], down triangles [27]). Green squares: counterflow flames with non-linear ex-
trapolation [42]. Solid black line: Konnov. Dashed red line: San Diego [73]. Dotted
sponding to peak SL of up to 0.23 were also observed. This scat-
blue line: CRECK [74].
ter is far beyond the evaluated uncertainty (if reported) from ex-
isting studies, and it is much higher than the observed discrep-
ancies among numerical predictions using recent models, which
is, in general, below 10% and reaches about 20% only under very data provides a valid opportunity to test the ability of the devel-
rich conditions. Considering the criticism raised by Konnov et al. oped model to predict the SL of DME flames over a wide range
[47] on some of the reported data, in the following, the comparison of . Thus, the computed SL based on the aforementioned Kon-
at ambient temperature will focus only on the data that were con- nov, San Diego, and CRECK detailed kinetic models are also shown
sidered more reliable by Konnov et al., namely [27,31,35,36,42] and with lines in Fig. 1. Qualitatively, the models reproduce the gen-
on the data that were published after their review [40,48]. Hence, eral trends of the experiments, with the peak of SL occurring un-
a comparison of the experimental SL versus  obtained in this der slightly rich conditions. The Konnov and CRECK models predict
study at room temperature and atmospheric pressure with those the peak value at  = 1.2, while the experimental one, as well
by Qin and Ju [27], de Vries et al. [31], Song et al. [35], Yu et al. as that computed with the San Diego model, occurs at  = 1.1.
[36], Wang et al. [42], Wang et al. [48] and Shrestha et al. [40] is Quantitatively, model predictions with the three mechanisms are
shown in Fig. 1. Wang et al. [48] measured SL using the heat flux very close to each other for lean mixtures and lie within the ex-
method, while the experiments of Shrestha et al. [40] were con- perimental scatter. For near-stoichiometric flames, the experimen-
ducted at TU-Freiberg utilizing a heat flux burner and at KAUST tal results are very well predicted by the San Diego and Konnov
using a constant volume spherical vessel with non-linear extrap- mechanisms, whereas the CRECK mechanism shows slightly higher
olation for extracting unstretched SL . Non-linear stretch corrected values. If only consistent datasets are considered under rich condi-
data were also reported by Qin and Ju [27] using the counterflow tions, the discrepancy between experimental data and model pre-
twin flame technique, whereas all the other datasets [31,35,36,42] dictions, as well as among predictions themselves, increases with
were obtained using the spherical flame method with linear ex- increasing , reaching a maximum relative difference of about 24%
trapolation to zero stretch. at  = 1.6, where the San Diego mechanism shows the best agree-
Considering all the datasets, the disagreement in the experi- ment. Overall, it can be concluded that if consistent datasets are
mental SL is found to be within 6% at 1.0 <  < 1.2 and reaches used for the comparison, all the mechanisms provide a satisfactory
about 32% and 23% under very lean or very rich conditions, respec- prediction of SL data over the full range of  since the level of
tively. discrepancy is considered moderate compared to what was shown
Referring only to experimental data obtained using the heat in the previous DME studies. In this regard, the current measure-
flux method, present measurements and those obtained by Wang ments provide additional benchmark data of high fidelity for test-
et al. [48] overlap within experimental uncertainties over the ing the accuracy of DME combustion models.
whole stoichiometric range explored. Although Shrestha et al. As mentioned before, the SL measurements have been per-
[40] reported comparatively lower values, the overall experimen- formed at several Tg in the present work. Fig. 2 compares the new
tal scatter observed in Fig. 1 is reduced to only about 5–7% for SL data at Tg = 298, 308, 323, and 338 K with predictions of the
0.9 <  < 1.6, which is well within the observed discrepancies Konnov model, which accurately reproduces the SL increase with
among model predictions, except for  = 0.7, where the experi- temperature. Calculations with the San Diego and CRECK mecha-
mental scatter reaches 15%, which however is still well below the nisms are not shown in Fig. 2 for clarity, but both mechanisms
32% observed among earlier studies. In particular, at  < 1.0, the predict a similar temperature dependence, as will be shown below.
experimental SL obtained by Qin and Ju [27] and by Yu et al. The overall dependence of SL on temperature can be described
[36] lie outside the scatter constrained by the recent measure- by a simple empirical correlation SL /SL 0 = (Tg /Tg 0 )α , where SL 0
ments. The same applies to the SL measured by de Vries et al. [31], is the burning velocity at a reference temperature Tg 0 [47]. Re-
Song et al. [35], Wang et al. [42] at  > 1.4. At  between 1.0 and lying on this correlation, one could expect a linear variation of
1.4, all the datasets are within the scatter. The consistency of these the experimental SL when plotted on a log(SL ) vs. log(Tg ) scale,

4
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

carbon fuels [47]. The variation of α derived from the present mea-
surements is in good agreement with that predicted by the three
kinetic mechanisms, except for  > 1.4 where model predictions
start to diverge with the San Diego model being the best, in agree-
ment with what is observed in Fig. 1. Consistent with the trends
shown in Fig. 3, α derived from the experimental data of Varghese
et al. are higher than those derived at lower temperatures in the
present work, particularly for lean mixtures. In this regard, Alek-
seev et al. [110] found that for hydrogen flames, α tends to rise
when shifting the correlation interval to Tg > 500 K, especially in
lean mixtures. To verify if this also applies to DME flames, α de-
rived at 400 < Tg < 650 K is compared with predictions obtained
in the same Tg range using the San Diego mechanism, shown with
the dashed red line. The computations are in excellent agreement
with the experimental data of Varghese et al. [43], which indirectly
proves their consistency with the present datasets; thus, both can
be used for the validation of kinetic mechanisms. Moreover, this
analysis also points out that the variation of α with the fitted
temperature range is a general behavior and does not depend on
Fig. 2. Present experimental (symbols) and numerical (lines) SL versus  at differ- the fuel. Therefore, α derived in the present work should not be
ent initial temperatures and p = 1 atm for DME/air flames.
used to extrapolate SL at Tg > 500 K, as well as α derived from
[43] should not be used to extrapolate SL down to ambient tem-
perature. This is the reason why, contrary to the review of Konnov
which helps in the examination of the consistency of experimen- et al. [47], the data from [43] have been excluded from the com-
tal datasets obtained at different Tg and identification of possible parison at ambient temperature (Fig. 1).
outliers not suitable for validation of kinetic models. There are few
experimental data available for comparison in the higher tempera- 5.2. NO profiles
ture range for DME/air mixtures. In particular, Shrestha et al. [40],
in addition to the results presented in Fig. 1, conducted experi- Results of post-flame XNO measurements, carried out 10 mm
ments at 373 K using the heat flux method over the 0.6–1.9  above the burner surface, are shown in Fig. 5 as a function of .
range. Yu et al. [36] and Wu et al. [37] measured the SL of DME/air As  increases, XNO rises until a peak of ∼84 ppm is reached at
mixtures at  = 1.0, respectively, in the 320–395 K and 333–  = 1.0, then smoothly decreases until a plateau is reached from
453 K ranges using a constant volume cylindrical chamber with  = 1.3 to 1.6. Fig. 5 also presents with lines the computed results
linear extrapolation for extracting unstretched SL , whereas Vargh- for XNO using the three models. Generally, the Konnov model offers
ese et al. [43] presented SL results from 400 to 650 K over the an improvement over previous mechanisms in the whole  range.
0.7–1.4  range using the externally heated mesoscale diverging Specifically, all models reproduce experimental trends for  < 1.3;
channel technique. however, the San Diego and CRECK overestimate the peak XNO by
Figure 3 shows the variation of SL with Tg at different . The about 18 and 15 ppm, respectively, while the Konnov model pre-
present and literature experimental values are shown with differ- dicts it well. At  > 1.2 the Konnov model slightly overpredicts
ent symbols, while the solid lines show a linear fit based on the XNO with a maximum discrepancy of about 7 ppm. Nevertheless,
Tg range explored in the current measurements. First, it can be ob- while the computations of the Konnov mechanism are in excellent
served that the SL obtained in the current work (solid squares) fol- agreement with the experimental trend in rich mixtures, evident
low linear trends at each , which proves the consistency of the discrepancies are found in the qualitative behavior of the other two
measurements obtained at different Tg . Then, the extrapolated val- schemes. Possible explanations for the discrepancies observed be-
ues at higher Tg are in fair agreement with the data reported in the tween predictions and experimental data are discussed below, to-
literature, while the measurements at  = 1 and 320 < Tg < 395 K gether with the effect of  on XNO .
by Wu et al. [37] (green circles) are obvious outliers. It should be Sensitivity analyzes of NO mole fractions for stoichiometric and
noted that the results reported by Wu et al. [37] were obtained rich ( = 1.3) mixtures under the conditions shown in Fig. 5 have
in the same laboratory as those by Yu et al. [36] (green triangles) been carried out as illustrated in Figs. 6 and 7, where a posi-
but published in separate papers. In this regard, the authors men- tive/negative normalized A-factor sensitivity coefficient means that
tioned that the error in , though not being quantified, was the XNO is increased/decreased with the increase of the specific rate
major source of experimental uncertainty due to the partial pres- constant. From Fig. 6 it can be observed that the most sen-
sure method used for mixture preparation. This is a reasonable ex- sitive reactions are the same for the three models, except for
planation of the inconsistency observed in Fig. 3, as the data by NNH+O=NH+NO, which is not implemented in the San Diego
Wu et al. [37] correlate with the present trend obtained at  = 0.9. mechanism. Indeed, nitrous oxide- and NNH-routes, which are im-
In addition, it can be seen that the results obtained by Varghese portant in lean flames, are still operational at  = 1.0. However,
et al. [43] (open squares) start to diverge from the present trend the peak at  = 1.0 shown in Fig. 5 is mainly due to the Zel’dovich
at Tg > 500 K, particularly under lean conditions, resulting in dif- thermal-NO mechanism, which contributes the most due to the
ferent temperature dependencies. A comparison of the power ex- very high temperature (2257 K) and sufficient residence time given
ponent α derived from the empirical power-law correlation from by the selected HAB = 10 mm. The mechanism of thermal-NO for-
both the present data at 298 < Tg < 338 K and those by Vargh- mation is well established: N2 reacts with oxygen atoms in the re-
ese et al. at 400 < Tg < 650 K, as well as from modeling results action N2 +O=NO+N (presented in the reverse direction in Fig. 6),
could be very instructive to elucidate the reasons for the observed followed by oxidation of N by OH or O2 . Thus, the formation of
inconsistency across different temperatures and . NO through the thermal mechanism is also dependent on the O/H
As shown in Fig. 4, α for DME/air flames varies non-linearly radical pool, as shown in Fig. 6, which is usually controlled by the
with , with a shape very similar to that found for many hydro- fuel oxidation chemistry. Considering that all models predict SL at

5
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

Fig. 3. Experimental (symbols) SL and linear fit (lines) versus Tg at different  and p = 1 atm for DME/air flames. Solid squares: present work (heat flux method). Crosses:
[40] (heat flux method). Up triangles: [36] (outwardly propagating spherical flames). Circles: [37] (outwardly propagating spherical flames). Open squares: [43] (externally
heated diverging channels).

Fig. 4. Experimental (symbols) and numerical (lines) α versus  at p = 1 atm Fig. 5. Experimental (symbols) and numerical (lines) XNO versus  at
for DME/air flames. Black squares: present work (298 < Tg < 338 K). Red circles: HAB = 10 mm, Tg = 338 K and p = 1 atm for DME/air flames. Black solid
[43] (400 < Tg < 650 K). Black line: Konnov (298 < Tg < 338 K). Solid red line: San line: Konnov. Red dashed line: San Diego [73]. Blue dotted line: CRECK [74].
Diego [73] (298 < Tg < 338 K). Dashed red line: San Diego [73] (400 < Tg < 650 K).
Blue line: CRECK [74] (298 < Tg < 338 K).

near-stoichiometric conditions in a satisfactory way (see Fig. 1), an


equally accurate description of the O/H radical pool would be ex-
pected (see Fig. S22 in the SM). This means that the overestimation
of the peak XNO using the San Diego and CRECK models is mainly
due to an overestimation of the rate constant for the rate-limiting
step N2 +O=NO+N. As a matter of fact, in the Konnov model, the
rate constant for this key reaction was derived in a recent study
on methane/enriched air flames [76], and at 20 0 0 K, it is about
5% lower compared to that presented in the San Diego and CRECK
mechanisms.
On the other hand, Fig. 5 reveals that the description of the
NO formation in rich mixtures has the highest uncertainty since
the XNO profiles predicted by three kinetic mechanisms are in
Fig. 6. NO sensitivity coefficients calculated at  = 1.0, Tg = 338 K, p = 1 atm, and
significant disagreement among themselves, and only the Konnov
HAB = 10 mm using the Konnov (black bars), San Diego [73] (red bars) and CRECK
mechanism provides a fairly good description of the experimen- [74] (blue bars) mechanisms.
tal data. The XNO sensitivity analysis in Fig. 7 performed at the

6
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

Fig. 7. NO sensitivity coefficients calculated at  = 1.3, Tg = 338 K, p = 1 atm, and


HAB = 10 mm using the Konnov (black bars), San Diego [73] (red bars) and CRECK
[74] (blue bars) mechanisms.
Fig. 8. Experimental (symbols) and numerical (lines) XNO versus  at
HAB = 10 mm, Tg = 338 K and p = 1 atm for DME/air (black symbols and
line) and ethanol/air (red symbols and line) flames. Experimental results for
conditions of experiments at  = 1.3 shows that the nitrous ox- ethanol are from [111].
ide and thermal-NO routes are suppressed due to lack of atomic
oxygen, while the prompt-NO route gains importance and is re-
sponsible for the plateau observed in Fig. 5. According to the
Konnov model, the amount of NO formed almost exclusively de-
pends on the rate of the prompt mechanism initiation reaction
CH+N2 =H+NCN, which is the most sensitive, and subsequent con-
version of NCN into HCN through the reaction NCN+H=HCN+N.
Reactions that promote/inhibit the formation of CH are also im-
portant because of the direct involvement of such a radical in the
interplay between fuel oxidation chemistries and nitrogen through
the prompt-NO route. From Fig. 7 it can be seen that the NCN
pathway is not implemented in the San Diego mechanism, where
the prompt-NO formation is described through the spin-forbidden
reaction CH+N2 =HCN+N, which is the most sensitive. This can ex-
plain the disagreement between experiments and the San Diego
model under rich conditions. Correct prediction of the prompt-NO
formation requires, in addition to an accurate rate constant for the
prompt mechanism initiation reaction, also the ability to predict
the CH concentration, as shown in Fig. 7. Updates of rate constants
for sensitive reactions that lead to an improved description of the Fig. 9. Simulated CH mole fraction (solid lines, left axis) and flame temperature
CH concentration in the Konnov model, together with the update (dashed lines, right axis) in rich flames of DME (black lines) and ethanol (red lines)
at  = 1.3.
of the prompt mechanism initiation reaction, were recently pre-
sented and discussed by Han et al. [76] and can be the reason for
better agreement compared to the CRECK model.
As mentioned in the introduction, another objective of this sponding DME/air flames at  ≤ 1.2, which is well predicted by
work was to enhance the understanding of the intrinsic NOx emis- the Konnov mechanism. This is due to the fact that DME exhibits
sion characteristics of DME due to the conflicting explanations higher flame temperatures compared to ethanol due to its higher
found in previous studies. For this purpose, Fig. 8 shows a com- enthalpy of formation, which confirms the thermal-NO as the pre-
parison between experimental XNO results from this study with dominant mechanism in this  range. In rich mixtures, the XNO
those previously measured in [111] for pure ethanol flames on a measured for both fuels overlap. However, as in the case of DME,
similar installation operated under identical conditions so that a the model tends to slightly overpredict XNO formed in rich ethanol
direct comparison is feasible. This analysis offers the opportunity flames with a maximum discrepancy of 9 ppm at  = 1.4. De-
to elucidate the mechanisms by which NO is produced/consumed spite this, the kinetic model satisfactorily reproduces the experi-
in the case of oxygenated compounds with the same elemen- mental trend as a function of  and the similar production of NO
tal formula (C2 H6 O) but different molecular structures and func- in ethanol and DME flames observed experimentally under rich
tional groups. Kinetic modeling is performed using only the Kon- conditions. Thus, the simulations can be used to elucidate simi-
nov model because it was shown in Fig. 5 to be the best in predict- larities and differences in the kinetics of the two isomers. As dis-
ing the present XNO data. Note that, while the experimental XNO cussed above (Fig. 7), the coupling of the fuel chemistry and the
for ethanol was already published, modeling results for these con- overall process of the prompt-NO formation passes via CH radi-
ditions are reported here for the first time. This extends the range cals. Therefore, the CH radical and temperature profiles calculated
of validation of the developed model. for both fuels at  = 1.3 are compared in Fig. 9. This condition is
Interestingly, the shape of the experimental dependence of XNO chosen as a representative case because XNO is the same for DME
versus  is the same for both isomeric fuels. However, Fig. 8 shows and ethanol flames in both experiments and simulations, and it is
that ethanol/air flames produce lower levels of NO than the corre- analyzed in full detail.

7
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

glet CH2 , efficiently converting NO back to HCN and HNO. On


the other hand, in the case of ethanol flames, about 40% of the
total rate of NO consumption is due to HCCO through the re-
actions HCCO+NO=HCNO+CO and HCCO+NO=HCN+CO2 , which
do not play an important role for DME (∼20%). This result re-
inforces the findings of Marrodán et al. [61], and it is in con-
trast with other studies [63,66,67] discussed in the introduction.
The reason is given by the fact that HCCO is mainly formed
by C2 H2 +O=HCCO+H and, as explained above, DME converts
only a small amount of its carbon to C2 species compared to
ethanol as a consequence of the absence of C–C bonds. In the
case of DME, C2 hydrocarbons are formed entirely from methyl
recombination, while ethanol can decompose to yield ethylene
(C2 H5 OH=C2 H4 +H2 O), which in turn produces acetylene through
the sequence C2 H4 →C2 H3 →C2 H2 . In addition, in ethanol flames,
Fig. 10. Most relevant rates of NO production calculated for DME (black) and a minor path that leads to HCCO is represented by the reaction
ethanol (red) flames at the point of maximum concentration of CH and  = 1.3, CH2 CO+H=HCCO+H2 , which is absent in the case of DME.
normalized with respect to the total rate of NO production.

6. Conclusion

As seen in Fig. 9, the simulated peak mole fractions of the CH In this work, post-flame NO mole fractions and adiabatic lam-
radical are shifted with respect to HAB, but they occur at the same inar burning velocities for laminar premixed DME/air flames were
flame temperature of ∼1840 K. The CH peak is higher in the case measured with the flat flame-based heat flux method coupled with
of DME, but the layer width is bigger in the ethanol flame due the LIF technique over equivalence ratios ranging from 0.7 to 1.6
to a lower temperature gradient, which partially offsets the higher and initial temperatures from 298 to 338 K, at atmospheric pres-
amount of CH in the DME flame and explains the similarity of sure. The present study yields three main findings:
XNO in rich mixtures between DME and ethanol flames observed New SL measurements were compared with selected data re-
in Fig. 8. A rate of production and reaction path analysis has been ported in the literature. Reasonably good agreement was found
performed at the point of maximum concentration of CH radicals at ambient temperature with recent datasets measured with the
for both fuels. Analysis of the reaction pathways indicates, as al- heat flux method [40,48] and spherical expanding flames [40]. The
ready known from the literature, that the formation of CH radi- temperature dependence of SL was used to examine the consis-
cals is strongly linked to the fuel-breakdown pathways. In partic- tency of experimental datasets obtained at higher temperatures,
ular, the formation of CH radicals depends on the competition be- and the comparison indicated that the measurements obtained by
tween CH3 pyrolysis and oxidation routes, where the former leads Wu et al. [37] are inconsistent with present and other SL exper-
to CH via singlet and triplet CH2 , while the latter mainly leads iments [36,40]. The power exponent α expressing the tempera-
to formaldehyde through the reaction CH3 +O=CH2 O+H. The fuel ture dependence of SL was derived as a function of  based on
structure appears to be critical for the formation and consumption the present results and compared to that obtained in a different
of CH3 . As a matter of fact, since DME does not have C–C bonds, H- temperature range using externally heated diverging channels [43],
abstraction reactions form the methoxymethyl radical (CH3 OCH2 ), which was found to be much higher. A closer look into this dis-
which promptly decomposes to form formaldehyde and methyl. crepancy revealed that the two datasets are consistent and the rea-
The high amount of formaldehyde inhibits CH3 oxidation via the son for the deviation is due to the rise of α when shifting the cor-
reverse reaction and enhances CH3 pyrolysis. On the other hand, relation interval to temperatures higher than 500 K, especially in
acetaldehyde is the main intermediate product of ethanol, which lean mixtures.
has a C–C bond, and this stresses the CH3 decomposition/oxidation A detailed kinetic model for the combustion of DME was pre-
competition, leading to a lower maximum of the simulated CH sented and validated against the new data and selected data from
profile. The rate of production analysis shown in Fig. 10 has been the literature. Overall good agreement between the experimen-
used to analyze the influence of individual reaction rates on pre- tal and modeling results was observed. A comparison between
dicted NO under fuel-rich conditions. In order to facilitate such the current consistent experimental data and numerical results
analysis, the rates of reactions have been normalized with respect obtained using two detailed kinetic mechanisms, San Diego and
to the total rate of NO production for both fuels. CRECK, was also conducted and discussed. The San Diego model
As already shown in Fig. 7, the prompt mechanism is the main was found to better predict the present SL experimental data,
source of NO at  = 1.3. The NCN formed via the prompt mecha- while the authors’ model was the best in the prediction of XNO .
nism initiation reaction is quickly converted to NO inside the flame Discrepancies between XNO experiments and model predictions
zone through a complex sequence of major reactions that involve and among models themselves were observed, particularly under
the interconversion between several fixed nitrogen intermediates rich conditions. Kinetic analyzes were used to identify the main
such as HCN, CN, HCNO, NCO, HNO, NH, and N [112]. Although the shortcomings. At near-stoichiometric conditions, the uncertainty in
important reactions N+O2 =NO+O and N+OH=NO+H also belong the rate constant of the thermal-NO initiation reaction was found
to the thermal mechanism, the analysis reveals that at  = 1.3, N to be responsible for the observed discrepancies. Under rich condi-
radicals are mainly formed through the reactions NCN+H=HCN+N tions, a notable difference in the prompt-NO mechanism descrip-
and NH+H=N+H2 . tion was found for the three mechanisms.
In general, the effects of molecular structure on the prompt- A quantitative comparison of experiments and simulation of
NO production and consumption through reburning reactions are XNO in ethanol and DME flames was systematically performed to
interwoven and cannot be easily decoupled. However, a clear dif- enhance the understanding of the intrinsic NOx emission charac-
ference can be observed in the reburning mechanisms of the two teristics of DME and to extend the validation range of the present
isomers. In detail, for DME flames, NO is predominantly consumed model, which could satisfactorily simulate XNO for both fuels. XNO
via its reactions with small radicals such as CH, HCO, and sin- in DME flames was found to be higher than those in ethanol

8
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

flames at  < 1.2 as a result of the higher flame temperature of [17] J. Jeon, S.I. Kwon, Y.H. Park, Y. Oh, S. Park, Visualizations of combustion and
DME compared to the case of ethanol, while in rich flames, no dif- fuel/air mixture formation processes in a single cylinder engine fueled with
DME, Appl. Energy 113 (2014) 294–301.
ferences were observed. Rate of production analyzes revealed that [18] G. Thomas, B. Feng, A. Veeraragavan, M.J. Cleary, N. Drinnan, Emissions from
the key pathways producing and consuming NO in rich flames are DME combustion in diesel engines and their implications on meeting future
driven by C1 species in the case of DME and by C2 species in the emissions norms: a review, Fuel Process. Technol. 119 (2014) 286–304.
[19] M. Yao, Z. Chen, Z. Zheng, B. Zhang, Y. Xing, Study on the controlling strate-
case of ethanol as a consequence of the absence of C–C bonds in gies of homogeneous charge compression ignition with fuel of dimethyl ether
the DME molecule. and methanol, Fuel 85 (2006) 2046–2056.
[20] Y. Putrasari, N. Jamsran, O. Lim, An investigation on the DME HCCI autoigni-
tion under EGR and boosted operation, Fuel 200 (2017) 447–457.
Declaration of Competing Interest [21] Y. Zhao, Y. Wang, D. Li, X. Lei, S. Liu, Combustion and emission characteristics
of a DME (dimethyl ether)-diesel dual fuel premixed charge compression ig-
The authors declare that they have no known competing finan- nition engine with EGR (exhaust gas recirculation), Energy 72 (2014) 608–617.
[22] M.C. Lee, S.B. Seo, J.H. Chung, Y.J. Joo, D.H. Ahn, Combustion performance test
cial interests or personal relationships that could have appeared to of a new fuel DME to adapt to a gas turbine for power generation, Fuel 87
influence the work reported in this paper. (2008) 2162–2167.
[23] M.C. Lee, S.B. Seo, J.H. Chung, Y.J. Joo, D.H. Ahn, Industrial gas turbine com-
bustion performance test of DME to use as an alternative fuel for power gen-
Acknowledgments eration, Fuel 88 (2009) 657–662.
[24] A. Stagni, S. Schmitt, M. Pelucchi, A. Frassoldati, K. Kohse-Höinghaus, T. Far-
This work was supported by the Swedish Energy Agency avelli, Dimethyl ether oxidation analyzed in a given flow reactor: experimen-
tal and modeling uncertainties, Combust. Flame 240 (2022) 111998.
through the Center for Combustion Science and Technology
[25] G.J. Gibbs, H.F. Calcote, Effect of molecular structure on burning velocity, J.
(Project KC-CECOST 22538-4) and by the Knut and Alice Wallen- Chem. Eng. Data 4 (1959) 226–237.
berg Foundation through grant KAW2019.0084 COCALD. The help [26] C.A. Daly, J.M. Simmie, J. Würmel, N. Djebaïli, C. Paillard, Burning velocities
of dimethyl ether and air, Combust. Flame 125 (2001) 1329–1340.
of Yijing Mao during the initial stage of experiments is gratefully
[27] X. Qin, Y. Ju, Measurements of burning velocities of dimethyl ether and
acknowledged. air premixed flames at elevated pressures, Proc. Combust. Inst. 30 (2005)
233–240.
[28] Z. Huang, Q. Wang, J. Yu, Y. Zhang, K. Zeng, H. Miao, D. Jiang, Measurement
Supplementary materials
of laminar burning velocity of dimethyl ether-air premixed mixtures, Fuel 86
(2007) 2360–2366.
Supplementary material associated with this article can be [29] Z. Chen, X. Qin, Y. Ju, Z. Zhao, M. Chaos, F.L. Dryer, High temperature igni-
found, in the online version, at doi:10.1016/j.combustflame.2022. tion and combustion enhancement by dimethyl ether addition to methane-air
mixtures, Proc. Combust. Inst. 31 (2007) 1215–1222.
112411. [30] Z. Chen, L. Wei, Z. Huang, H. Miao, X. Wang, D. Jiang, Measurement of laminar
burning velocities of dimethyl ether-air premixed mixtures with N2 and CO2
References dilution, Energy Fuels 23 (2009) 735–739.
[31] J. de Vries, W.B. Lowry, Z. Serinyel, H.J. Curran, E.L. Petersen, Laminar flame
[1] S.C. Sorenson, S. Mikkelsen, Performance and emissions of a 0.273 liter di- speed measurements of dimethyl ether in air at pressures up to 10 atm, Fuel
rect injection Diesel engine fuelled with neat dimethyl ether, SAE Trans. 104 90 (2011) 331–338.
(1995) 80–90. [32] W.B. Lowry, Z. Serinyel, M.C. Krejci, H.J. Curran, G. Bourque, E.L. Petersen, Ef-
[2] R. Christensen, S.C. Sorenson, M.G. Jensen, K.F. Hansen, Engine operation on fect of methane-dimethyl ether fuel blends on flame stability, laminar flame
dimethyl ether in a naturally aspirated, DI Diesel engine, SAE Trans. 106 speed, and Markstein length, Proc. Combust. Inst. 33 (2011) 929–937.
(1997) 1863–1872. [33] Z. Chen, C. Tang, J. Fu, X. Jiang, Q. Li, L. Wei, Z. Huang, Experimental and
[3] S. Kajitani, Z.L. Chen, M. Konno, K.T. Rhee, Engine performance and ex- numerical investigation on diluted DME flames: thermal and chemical kinetic
haust characteristics of direct-injection diesel engine operated with DME, SAE effects on laminar flame speeds, Fuel 102 (2012) 567–573.
Trans. 106 (1997) 1568–1577. [34] D. Liu, J. Santner, C. Togbé, D. Felsmann, J. Koppmann, A. Lackner, X. Yang,
[4] M. Konno, S. Kajitani, M. Oguma, T. Iwase, K. Shima, NO emission character- X. Shen, Y. Ju, K. Kohse-Höinghaus, Flame structure and kinetic studies of car-
istics of a CI engine fueled with neat dimethyl ether, SAE Trans. 108 (1999) bon dioxide-diluted dimethyl ether flames at reduced and elevated pressures,
460–467. Combust. Flame 160 (2013) 2654–2668.
[5] S.C. Sorenson, Dimethyl ether in diesel engines: progress and perspectives, J. [35] W.S. Song, S.W. Jung, J. Park, O.B. Kwon, Y.J. Kim, T.H. Kim, J.H. Yun, S.I. Keel,
Eng. Gas Turbines Power 123 (2001) 652–658. Effects of syngas addition on flame propagation and stability in outwardly
[6] H. Teng, J.C. McCandless, J.B. Schneyer, Thermochemical characteristics of propagating spherical dimethyl ether-air premixed flames, Int. J. Hydrog. En-
dimethyl ether – an alternative fuel for compression-ignition engines, SAE ergy 38 (2013) 14102–14114.
Trans. 110 (2001) 96–106. [36] H. Yu, E. Hu, Y. Cheng, X. Zhang, Z. Huang, Experimental and numerical
[7] R. Egnell, Comparison of heat release and NOx formation in a DI diesel engine study of laminar premixed dimethyl ether/methane-air flame, Fuel 136 (2014)
running on DME and diesel fuel, SAE Trans. 110 (2001) 492–506. 37–45.
[8] J. Song, Z. Huang, X. Qiao, W. Wang, Performance of a controllable premixed [37] H. Wu, E. Hu, H. Yu, Q. Li, Z. Zhang, Y. Chen, Z. Huang, Experimental and
combustion engine fueled with dimethyl ether, Energy Convers. Manag. 45 numerical study on the laminar flame speed of n-butane/dimethyl ether-air
(2004) 2223–2232. mixtures, Energy Fuels 28 (2014) 3412–3419.
[9] D. Cipolat, Analysis of energy release and NOx emissions of a CI engine fu- [38] B. Zhang, H.D. Ng, An experimental investigation of the explosion character-
elled on diesel and DME, Appl. Therm. Eng. 27 (2007) 2095–2103. istics of dimethyl ether-air mixtures, Energy 107 (2016) 1–8.
[10] M.Y. Kim, S.H. Bang, C.S. Lee, Experimental investigation of spray and com- [39] Y. Yamamoto, T. Tachibana, Burning velocities of dimethyl ether
bustion characteristics of dimethyl ether in a common-rail diesel engine, En- (DME)-nitrous oxide (N2 O) mixtures, Fuel 217 (2018) 160–165.
ergy Fuels 21 (2007) 793–800. [40] K.P. Shrestha, S. Eckart, A.M. Elbaz, B.R. Giri, C. Fritsche, L. Seidel, W.L. Roberts,
[11] W. Ying, L. Genbao, Z. Wei, Z. Longbao, Study on the application of H. Krause, F. Mauss, A comprehensive kinetic model for dimethyl ether and
DME/diesel blends in a diesel engine, Fuel Process. Technol. 89 (2008) dimethoxymethane oxidation and NOx interaction utilizing experimental lam-
1272–1280. inar flame speed measurements at elevated pressure and temperature, Com-
[12] M.Y. Kim, S.H. Yoon, B.W. Ryu, C.S. Lee, Combustion and emission character- bust. Flame 218 (2020) 57–74.
istics of DME as an alternative fuel for compression ignition engines with a [41] Z. Zhao, A. Kazakov, F.L. Dryer, Measurements of dimethyl ether/air mixture
high pressure injection system, Fuel 87 (2008) 2779–2786. burning velocities by using particle image velocimetry, Combust. Flame 139
[13] C. Arcoumanis, C. Bae, R. Crookes, E. Kinoshita, The potential of di-methyl (2004) 52–60.
ether (DME) as an alternative fuel for compression-ignition engines: a review, [42] Y.L. Wang, A.T. Holley, C. Ji, F.N. Egolfopoulos, T.T. Tsotsis, H.J. Curran, Propa-
Fuel 87 (2008) 1014–1030. gation and extinction of premixed dimethyl-ether/air flames, Proc. Combust.
[14] J. Liu, S. Liu, Y. Li, Y. Wei, G. Li, Z. Zhu, Regulated and nonregulated emissions Inst 32 (2009) 1035–1042.
from a dimethyl ether powered compression Ignition engine, Energy Fuels 24 [43] R.J. Varghese, V. Ratna Kishore, A. Mohammad, Y. Soon, S. Kumar, Burning
(2010) 2465–2469. velocities of DME (dimethyl ether)-air premixed flames at elevated tempera-
[15] I.M. Youn, S.H. Park, H.G. Roh, C.S. Lee, Investigation on the fuel spray and tures, Energy 126 (2017) 34–41.
emission reduction characteristics for dimethyl ether (DME) fueled multi– [44] A. Naseer Mohammed, K.A. Juhany, S. Kumar, V. Ratna Kishore, A. Moham-
cylinder diesel engine with common-rail injection system, Fuel Process. Tech- mad, Effects of CO2 /N2 dilution on laminar burning velocity of stoichiomet-
nol. 92 (2011) 1280–1287. ric DME-air mixture at elevated temperatures, J. Hazard. Mater. 333 (2017)
[16] S.H. Park, C.S. Lee, Applicability of dimethyl ether (DME) in a compression 215–221.
ignition engine as an alternative fuel, Energy Convers. Manag. 86 (2014)
848–863.

9
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

[45] A. Mohammad, K.A. Juhany, Inert gas dilution and temperature effects on [72] M. Lubrano Lavadera, A.A. Konnov, Data consistency of the burning velocity
laminar burning velocity of DME + air mixtures, Energy Fuels 32 (2018) measurements using the heat flux method: syngas flames, Energy Fuels 34
6347–6354. (2020) 3725–3742.
[46] A. Mohammad, K.A. Juhany, Laminar burning velocity and flame structure [73] Chemical-Kinetic Mechanisms for Combustion ApplicationsMechanical
of DME/methane + air mixtures at elevated temperatures, Fuel 245 (2019) and Aerospace Engineering (Combustion Research), University of Califor-
105–114. nia at San Diego, 2022. https://web.eng.ucsd.edu/mae/groups/combustion/
[47] A.A. Konnov, A. Mohammad, V. Ratna Kishore, N. Il Kim, C. Prathap, S. Ku- mechanism.html; https://combustion.ucsd.edu.
mar, A comprehensive review of measurements and data analysis of laminar [74] https://creckmodeling.chem.polimi.it/menu- kinetics/menu- kinetics- detailed-
burning velocities for various fuel+air mixtures, Prog. Energy Combust. Sci. mechanisms (2022).
68 (2018) 197–267. [75] M. Lubrano Lavadera, A.A. Konnov, Laminar burning velocities of propionic
[48] Z. Wang, S. Wang, R. Whiddon, X. Han, Y. He, K. Cen, Effect of hydrogen addi- acid + air flames: experimental, modeling and data consistency study, Com-
tion on laminar burning velocity of CH4 /DME mixtures by heat flux method bust. Flame 230 (2021) 111431.
and kinetic modeling, Fuel 232 (2018) 729–742. [76] X. Han, M. Lubrano Lavadera, C. Brackmann, Z. Wang, Y. He, A.A. Kon-
[49] K. Hoon Song, P. Nag, T.A. Litzinger, D.C. Haworth, Effects of oxygenated addi- nov, Experimental and kinetic modeling study of NO formation in premixed
tives on aromatic species in fuel-rich, premixed ethane combustion: a mod- CH4 +O2 +N2 flames, Combust. Flame 223 (2021) 349–360.
eling study, Combust. Flame 135 (2003) 341–349. [77] X. Han, M. Lubrano Lavadera, A.A. Konnov, An experimental and kinetic mod-
[50] J. Wu, K. Hoon Song, T. Litzinger, S. Lee, R. Santoro, M. Linevsky, Reduction of eling study on the laminar burning velocity of NH3 +N2 O+air flames, Com-
PAH and soot in premixed ethylene-air flames by addition of dimethyl ether, bust. Flame 228 (2021) 13–28.
Combust. Sci. Technol. 178 (2006) 837–863. [78] C. Zhou, Y. Li, U. Burke, C. Banyon, K.P. Somers, S. Ding, S. Khan, J.W. Hargis,
[51] C.S. McEnally, L.D. Pfefferle, The effects of dimethyl ether and ethanol on ben- T. Sikes, O. Mathieu, E.L. Petersen, M. AlAbbad, A. Farooq, Y. Pan, Y. Zhang,
zene and soot formation in ethylene nonpremixed flames, Proc. Combust. Inst. Z. Huang, J. Lopez, Z. Loparo, S.S. Vasu, H.J. Curran, An experimental and
31 (2007) 603–610. chemical kinetic modeling study of 1,3-butadiene combustion: ignition de-
[52] J. Wang, Y. Struckmeier, B. Yang, T.A. Cool, P. Osswald, K. Kohse-Höing- lay time and laminar flame speed measurements, Combust. Flame 197 (2018)
haus, T. Kasper, N. Hansen, P.R. Westmoreland, Isomer-specific influences on 423–438.
the composition of reaction intermediates in dimethyl ether/propene and [79] E. Goos, A. Burcat, B. Ruscic, Extended Third Millennium Ideal Gas Thermo-
ethanol/propene flame, J. Phys. Chem. A 112 (2008) 9255–9265. chemical Database with updates from Active Thermochemical Tables. http:
[53] S.S. Yoon, D.H. Anh, S.H. Chung, Synergistic effect of mixing dimethyl ether //burcat.technion.ac.il/dir (2010).
with methane, ethane, propane, and ethylene fuels on polycyclic aromatic hy- [80] C.F. Goldsmith, G.R. Magoon, W.H. Green, Database of small molecule ther-
drocarbon and soot formation, Combust. Flame 154 (2008) 368–377. mochemistry for combustion, J. Phys. Chem. A 116 (2012) 9033–9057.
[54] B.A.V. Bennett, C.S. McEnally, L.D. Pfefferle, M.D. Smooke, M.B. Colket, Com- [81] R. Sivaramakrishnan, J.V. Michael, A.F. Wagner, R. Dawes, A.W. Jasper,
putational and experimental study of the effects of adding dimethyl ether L.B. Harding, Y. Georgievskii, S.J. Klippenstein, Roaming radicals in the thermal
and ethanol to nonpremixed ethylene/air flames, Combust. Flame 156 (2009) decomposition of dimethyl ether: experiment and theory, Combust. Flame
1289–1302. 158 (2011) 618–632.
[55] C. Esarte, Á. Millera, R. Bilbao, M.U. Alzueta, Effect of ethanol, dimethylether, [82] V. da Rocha, R.R.V. Castro, G. Machado, G. Bauerfeldt, Theoretical prediction
and oxygen, when mixed with acetylene, on the formation of soot and gas of rate coefficients and reassessment of the dimethyl ether combustion mech-
products, Ind. Eng. Chem. Res. 49 (2010) 6772–6779. anism, Proceedings of the 8th European Combustion Meeting (2017), paper
[56] A. Frassoldati, T. Faravelli, E. Ranzi, K. Kohse-Höinghaus, P.R. Westmore- 0450.
land, Kinetic modeling study of ethanol and dimethyl ether addition to [83] K. Takahashi, O. Yamamoto, T. Inomata, M. Kogoma, Shock-tube studies on the
premixed low-pressure propene-oxygen-argon flames, Combust. Flame 158 reactions of dimethyl ether with oxygen and hydrogen atoms, Int. J. Chem.
(2011) 1264–1276. Kinet. 39 (2007) 97–108.
[57] T. Amano, F.L. Dryer, Effect of dimethyl ether, NOx , and ethane on CH4 oxi- [84] C. Bänsch, J. Kiecherer, M. Szöri, M. Olzmann, Reaction of dimethyl ether with
dation: high pressure, intermediate-temperature experiments and modeling, hydroxyl radicals: kinetic isotope effect and prereactive complex formation, J.
Symp. (Int.) Combust. 27 (1998) 397–404. Phys. Chem. A 117 (2013) 8343–8351.
[58] M.U. Alzueta, J. Muro, R. Bilbao, P. Glarborg, Oxidation of dimethyl ether and [85] J. Mendes, C. Zhou, H.J. Curran, Rate constant calculations of H-atom ab-
its interaction with nitrogen oxides, Isr. J. Chem. 39 (1999) 73–86. straction reactions from ethers by HȮ2 radicals, J. Phys. Chem. A 118 (2014)
[59] P. Dagaut, J. Luche, M. Cathonnet, Mutual sensitization of the oxidation of 1300–1308.
DME and nitric oxide: experimental and detailed kinetic modeling, Combust. [86] Y. Guan, J. Gao, Y. Song, Y. Li, H. Ma, J. Song, Variational effect and anhar-
Sci. Technol. 165 (2001) 61–84. monic torsion on kinetic modeling for initiation reaction of dimethyl ether
[60] I. Liu, N.W. Cant, J.H. Bromly, F.J. Barnes, P.F. Nelson, B.S. Haynes, Formate combustion, J. Phys. Chem. A 121 (2017) 1121–1132.
species in the low-temperature oxidation of dimethyl ether, Chemosphere 42 [87] R.S. Tranter, P.T. Lynch, C.J. Annesley, Shock tube investigation of
(2001) 583–589. CH3 + CH3 OCH3 , J. Phys. Chem. A 116 (2012) 7287–7292.
[61] L. Marrodán, L. Berdusán, V. Aranda, Á. Millera, R. Bilbao, M.U. Alzueta, Influ- [88] Y. Guan, Y. Li, L. Zhao, Y. Song, H. Ma, J. Song, Hydrogen transfer between
ence of dimethyl ether addition on the oxidation of acetylene in the absence dimethyl ether and the methoxy radical: understanding and kinetic modeling
and presence of NO, Fuel 183 (2016) 1–8. with anharmonic torsions, Comput. Theor. Chem. 1089 (2016) 45–53.
[62] W. Ye, J.C. Shi, R.T. Zhang, X.J. Wu, X. Zhang, M.L. Qi, S.N. Luo, Experimental [89] H. Hashemi, J.M. Christensen, P. Glarborg, High-pressure pyrolysis and oxida-
and kinetic modeling study of CH3 OCH3 ignition sensitized by NO2 , Energy tion of DME and DME/CH4 , Combust. Flame 205 (2019) 80–92.
Fuels 30 (2016) 10900–10908. [90] S.L. Fischer, F.L. Dryer, H.J. Curran, The reaction kinetics of dimethyl ether. I:
[63] T. Yamamoto, S. Kajimura, Kinetic study on NO reduction using dimethyl high-temperature pyrolysis and oxidation in flow reactors, Int. J. Chem. Kinet.
ether as a reburning fuel, Energy Fuels 31 (2017) 12500–12507. 32 (20 0 0) 713–740.
[64] L. Marrodán, Á.J. Arnal, Á. Millera, R. Bilbao, M.U. Alzueta, The inhibiting [91] U. Burke, K.P. Somers, P. O’Toole, C.M. Zinner, N. Marquet, G. Bourque, E.L. Pe-
effect of NO addition on dimethyl ether high-pressure oxidation, Combust. tersen, W.K. Metcalfe, Z. Serinyel, H.J. Curran, An ignition delay and kinetic
Flame 197 (2018) 1–10. modeling study of methane, dimethyl ether, and their mixtures at high pres-
[65] H. Zhang, S. Schmitt, L. Ruwe, K. Kohse-Höinghaus, Inhibiting and promot- sures, Combust. Flame 162 (2015) 315–330.
ing effects of NO on dimethyl ether and dimethoxymethane oxidation in a [92] A.J. Eskola, S.A. Carr, R.J. Shannon, B. Wang, M.A. Blitz, M.J. Pilling,
plug-flow reactor, Combust. Flame 224 (2021) 94–107. P.W. Seakins, S.H. Robertson, Analysis of the kinetics and yields of OH rad-
[66] C. Hwang, C. Lee, K. Lee, Fundamental studies of NOx emission characteris- ical production from the CH3 OCH2 + O2 reaction in the temperature range
tics in dimethyl ether (DME)/air non-premixed flames, Energy Fuels 23 (2009) 195–650 K: an experimental and computational study, J. Phys. Chem. A 118
754–761. (2014) 6773–6788.
[67] D. Lee, J. Seong Lee, H. Young Kim, C. Kyun Chun, S.C. James, S.S. Yoon, [93] A. Rodriguez, O. Frottier, O. Herbinet, R. Fournet, R. Bounaceur, C. Fittschen,
Experimental study on the combustion and NOx emission characteristics of F. Battin-Leclerc, Experimental and modeling investigation of the low-temper-
DME/LPG blended fuel using counterflow burner, Combust. Sci. Technol. 184 ature oxidation of dimethyl ether, J. Phys. Chem. A 119 (2015) 7905–7923.
(2012) 97–113. [94] E.E. Dames, A.S. Rosen, B.W. Weber, C.W. Gao, C. Sung, W.H. Green, A detailed
[68] C.A. Frye, A.L. Boehman, P.J.A. Tijm, Comparison of CO and NO emissions combined experimental and theoretical study on dimethyl ether/propane
from propane, n-butane, and dimethyl ether premixed flames, Energy Fuels blended oxidation, Combust. Flame 168 (2016) 310–330.
13 (1999) 650–654. [95] Z. Zhao, M. Chaos, A. Kazakov, F.L. Dryer, Thermal decomposition reaction
[69] S. Eckart, L. Cai, C. Fritsche, F. Vom Lehn, H. Pitsch, H. Krause, Laminar and a comprehensive kinetic model of dimethyl ether, Int. J. Chem. Kinet. 40
burning velocities, CO, and NOx emissions of premixed polyoxymethylene (2008) 1–18.
dimethyl ether flames, Fuel 293 (2021) 120321. [96] M. Lubrano Lavadera, A.A. Konnov, Laminar burning velocities of
[70] C. Brackmann, T. Methling, M. Lubrano Lavadera, G. Capriolo, A.A. Konnov, methane + formic acid + air flames: experimental and modeling study,
Experimental and modeling study of nitric oxide formation in premixed Combust. Flame 225 (2021) 65–73.
methanol + air flames, Combust. Flame 213 (2020) 322–330. [97] P. Dagaut, C. Daly, J.M. Simmie, M. Cathonnet, The oxidation and ignition of
[71] V.A. Alekseev, J.D. Naucler, M. Christensen, E.J.K. Nilsson, E.N. Volkov, L.P.H. de dimethylether from low to high temperature (50 0–160 0 K): experiments and
Goey, A.A. Konnov, Experimental uncertainties of the heat flux method for kinetic modeling, Symp. (Int.) Combust. 27 (1998) 361–369.
measuring burning velocities, Combust. Sci. Technol. 188 (2016) 853–894. [98] K. Stoehr, N. Peters, J. Beeckmann, Low temperature oscillations of DME com-
bustion in a jet-stirred reactor, Proc. Combust. Inst. 35 (2015) 3601–3607.

10
M. Lubrano Lavadera, C. Brackmann and A.A. Konnov Combustion and Flame 246 (2022) 112411

[99] Z. Wang, X. Zhang, L. Xing, L. Zhang, F. Herrmann, K. Moshammer, F. Qi, [107] L. Dai, L. Lu, C. Zou, Q. Lin, W. Xia, H. Shi, J. Luo, C. Peng, S. Wang, An
K. Kohse-Höinghaus, Experimental and kinetic modeling study of the low- improved low and high-temperature dimethyl ether kinetic model for the
and intermediate-temperature oxidation of dimethyl ether, Combust. Flame combustion atmospheres with high CO2 concentration, Combust. Flame 238
162 (2015) 1113–1125. (2022) 111922.
[100] U. Pfahl, F.G. Adomeit, Self-ignition of diesel-relevant hydrocarbon-air mix- [108] H. Wang, R. Fang, B.W. Weber, C. Sung, An experimental and modeling study
tures under engine conditions, Symp. (Int.) Combust. 26 (1996) 781–789. of dimethyl ether/methanol blends autoignition at low temperature, Combust.
[101] C. Tang, L. Wei, J. Zhang, X. Man, Z. Huang, Shock tube measurements and Flame 198 (2018) 89–99.
kinetic investigation on the ignition delay times of methane/dimethyl ether [109] A. Sudholt, L. Cai, H. Pitsch, Laminar flow reactor experiments for ignition
mixtures, Energy Fuels 26 (2012) 6720–6728. delay time and species measurements at low temperatures: linear alkanes
[102] Z. Li, W. Wang, Z. Huang, M. Oehlschlaeger, Dimethyl ether autoignition at and dimethyl ether, Combust. Flame 202 (2019) 347–361.
engine-relevant conditions, Energy Fuels 27 (2013) 2811–2817. [110] V.A. Alekseev, M. Christensen, A.A. Konnov, The effect of temperature on the
[103] E. Hu, Z. Zhang, L. Pan, J. Zhang, Z. Huang, Experimental and modeling study adiabatic burning velocities of diluted hydrogen flames: a kinetic study using
on ignition delay times of dimethyl ether/propane/oxygen/argon mixtures at an updated mechanism, Combust. Flame 162 (2015) 1884–1898.
20 bar, Energy Fuels 27 (2013) 4007–4013. [111] M. Lubrano Lavadera, C. Brackmann, G. Capriolo, T. Methling, A.A. Konnov,
[104] E. Hu, X. Jiang, Z. Huang, J. Zhang, Z. Zhang, X. Man, Experimental and kinetic Measurements of the laminar burning velocities and NO concentrations in
studies on ignition delay times of dimethyl ether/n-butane/O2 /Ar mixtures, neat and blended ethanol and n-heptane flames, Fuel 288 (2021) 119585.
Energy Fuels 27 (2013) 530–536. [112] N. Lamoureux, H. El Merhubi, L. Pillier, S. de Persis, P. Desgroux, Modeling of
[105] X. Jiang, Y. Zhang, X. Man, L. Pan, Z. Huang, Shock tube measurements and ki- NO formation in low pressure premixed flames, Combust. Flame 163 (2016)
netic study on ignition delay times of lean DME/n-butane blends at elevated 557–575.
pressures, Energy Fuels 27 (2013) 6238–6246.
[106] X. Jiang, Y. Zhang, X. Man, L. Pan, Z. Huang, Experimental and modeling study
on ignition delay times of dimethyl ether/n-butane blends at a pressure of 2.0
MPa, Energy Fuels 28 (2014) 2189–2198.

11

You might also like