Next Article in Journal
Evaluation and Improvement of the Sound Quality of a Diesel Engine Based on Tests and Simulations
Previous Article in Journal
Multi-Disciplinary Optimization Design of Axial-Flow Pump Impellers Based on the Approximation Model
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Dual Fuel Reaction Mechanism 2.0 including NOx Formation and Laminar Flame Speed Calculations Using Methane/Propane/n-Heptane Fuel Blends

Institute of Chemical, Environmental and Bioscience Engineering, TU Wien, Getreidemarkt 9/166, 1060 Vienna, Austria
*
Author to whom correspondence should be addressed.
Energies 2020, 13(4), 778; https://doi.org/10.3390/en13040778
Submission received: 23 December 2019 / Revised: 27 January 2020 / Accepted: 30 January 2020 / Published: 11 February 2020
(This article belongs to the Section B: Energy and Environment)

Abstract

:
This study presents the further development of the TU Wien dual fuel mechanism, which was optimized for simulating ignition and combustion in a rapid compression expansion machine (RCEM) in dual fuel mode using diesel and natural gas at pressures higher than 60 bar at the start of injection. The mechanism is based on the Complete San Diego mechanism with n-heptane extension and was attuned to the RCEM measurements to achieve high agreement between experiments and simulation. This resulted in a specific application area. To obtain a mechanism for a wider parameter range, the Arrhenius parameter changes performed were analyzed and updated. Furthermore, the San Diego nitrogen sub-mechanism was added to consider NOx formation. The ignition delay time-reducing effect of propane addition to methane was closely examined and improved. To investigate the propagation of the flame front, the laminar flame speed of methane–air mixtures was simulated and compared with measured values from literature. Deviations at stoichiometric and fuel-rich conditions were found and by further mechanism optimization reduced significantly. To be able to justify the parameter changes performed, the resulting reaction rate coefficients were compared with data from the National Institute of Standards and Technology chemical kinetics database.

1. Introduction

In the Paris Climate Agreement, it was negotiated that measures have to be taken to keep global warming below 2 °C [1]. To achieve this, a significant reduction of greenhouse gases is required. Electrification in the mobile sector plays a major role in achieving this goal, but also in the field of conventional combustion technology, optimization of existing and development of highly efficient technologies and advanced fuels [2,3] can make an important contribution. For example, in the transport sector, the use of dual fuel engines instead of pure diesel engines offers the potential for emission savings [4]. In the dual fuel combustion process, a mixture of air and natural gas is compressed by the piston and ignited by injection of a diesel pilot. In this operation mode, low-calorie fuel gases such as biogas can be used [5]. For emission reasons, the gas content should be as large as possible in the combustion process. Due to the high global warming potential of methane [6], which is a major component of natural gas, it is essential in dual fuel combustion that methane is completely converted and not emitted unburned. The dual fuel combustion process is very complex and not yet understood in detail.
In [7], the dual fuel combustion process was investigated experimentally and by simulations. Since diesel and natural gas consist of a huge number of different hydrocarbons, the consideration of all species in the simulation would lead to an unmanageable simulation time. Therefore, surrogate fuels were defined for the simulation. n-heptane is a common diesel substitute in dual fuel investigations [8,9,10] and was used as a surrogate. It is therefore well suited as diesel substitute, since with a cetane number (which describes the flammability of the fuel) of 56 [11], n-heptane is in the range of the diesel cetane number, which has to be 51 or higher according to the DIN EN590 standard of October 2009. In analogy to diesel fuel, it is necessary to define a substitute fuel for natural gas. A frequently used substitute fuel is methane [8,10]. To better reflect the methane number (which is a measure of the knock resistance) of natural gas, a mixture of methane and propane was defined as a surrogate. In [7], in the first step, the ignition delay times (IDTs) of homogeneous methane–propane–n-heptane mixtures were investigated with a rapid compression machine (RCM) and a shock tube (ST) at PCFC (Physico Chemical Fundamentals of Combustion), RWTH Aachen [12,13], followed by the comparison of the measured values with simulation results using various n-heptane mechanisms available in the literature. For the simulations in [7], as well as the actual study, the program LOGEresearch (Version LSv1.09, LOGE AB, Lund, Sweden) [14] was used. The simulation of the RCM measurements was performed with the module “rapid compression machine” in LOGEresearch at a specific fuel composition, fuel–air equivalence ratio, starting temperature, and starting pressure of the mixture before compression. The RCM facility effects describe the non-ideal behavior of the test facility and the machine-specific compression of the mixture. The effects are considered in the simulation by reading in the effective volume profile derived from the pressure profile of the associated non-reactive RCM measurement in which oxygen was replaced by an inert gas. Details on determining the effective volume curve are available in [12,15].
As shown on the left side of Figure 1, the IDT is defined as the time between the point when the piston of the RCM reached its end position and the maximal pressure increase due to ignition of the mixture. The simulation of ST measurements was performed with the module “rapid compression machine” too, but with the difference that the temperature and pressure values reached in the ST after the shockwave reflection were used as starting values. The IDT is thus defined as the time difference between the simulation start and reaching the maximal pressure increase due to ignition as shown on the right side of Figure 1. A device-specific pressure increase of 8%/ms was considered in the ST-simulation. The constant volume (CV) module of LOGEresearch was used to perform CV simulations at specified initial conditions, like pressure, temperature, and simulation time. The IDT definition is identical to that of the ST simulations.
When comparing the simulation results using different n-heptane mechanisms in [7] over the entire parameter range considered, the most promising results were achieved with the Complete San Diego mechanism with n-heptane extension [16]. The second step in [7] comprised the study of inhomogeneous mixtures. The investigation of the ignition of a diesel spray injected into air and a natural gas–air mixture experimentally as well as theoretically by computational fluid dynamics (CFD) simulations showed in part large deviations of the IDTs, which motivated the adaptation of the Complete San Diego reaction mechanism with n-heptane extension. By means of sensitivity and flow analyzes, reactions for the mechanism adaption were identified. The adapted mechanism is able to describe ignition phenomena in the context of natural gas–diesel dual fuel combustion. However, when considering the ignition properties of methane–propane mixtures, it could be established that the IDT-reducing effect of the propane addition to methane was overestimated. This was pointed out in [7], but a detailed treatment had to be postponed to follow-up investigations. Some changes made to the Arrhenius parameters are quite large, resulting in a specific application area of the mechanism. Due to the identified optimization potential in order to obtain a larger parameter range for the application, the mechanism was further developed. The present study deals with the detailed investigation of the kinetic-controlled ignition process focusing on homogenous fuel mixtures in order to avoid a superposition of the actual ignition properties of the mixture with the jet preparation and the non-uniform distribution of the fuel influencing the ignition process in the case of inhomogeneous mixtures. A total of 10 different homogeneous mixtures with various proportions of methane, propane and n-heptane in the pressure range 60 to 100 bar were investigated. An overview of the mixtures is given in Table 1. Details on the selection of the homogeneous mixtures used are available in [7].

2. Materials and Methods

The mechanism presented in [7] is based on the Complete San Diego mechanism with n-heptane extension [16]. As this base mechanism has been expanded since the work presented in [7] and now consists of 303 reactions instead of 301, the most recent sub-mechanism files were downloaded from the San Diego mechanism database (Complete San Diego mechanism CK_2016-12-14, Heptane Chemistry CK_2015-03-01). In addition, the nitrogen sub-mechanism (Nitrogen Chemistry CK_2018-07-23) has been added in order to be able to calculate the formation of NOx during combustion. The compound mechanism (MechA) consists of 344 reactions and considers 75 species.
The further development process is explained in the following and is shown graphically in Figure 2. First, the Arrhenius parameter changes made in [7] were transferred to the mechanism MechA. Second, a re-evaluation of the Arrhenius parameter changes was made. In the third step, the overestimation of the IDT-reducing effect of the propane addition to methane was investigated. In the fourth step, the laminar flame speed of methane–air mixtures was simulated and compared with experimental values taken from the literature. Derived from this comparison, the final adaptation of the mechanism was performed. It should be noted that during the whole optimization process the thermodynamic and transport data remained unchanged and therefore correspond to the original data files of the Complete San Diego mechanism.

2.1. Update of Arrhenius Parameters Adaption

At first, the IDTs were calculated using the compound mechanism MechA (Complete San Diego mechanism, n-heptane, and nitrogen sub-mechanism). In the next step, the parameter changes made in [7] were transferred to the compound mechanism resulting in MechB. A comparison of the simulated IDTs using MechA and MechB is shown in Figure 3.
In the mechanism presented in [7], the Arrhenius parameters of eight reactions were adjusted, including three methane reactions and five reactions of the n-heptane sub-mechanism. The main target of the adaption was to reduce the difference between the pressure curve (and derived therefrom the point of ignition) measured in a RCEM, operated in dual fuel mode with diesel and natural gas, and the corresponding CFD simulation results. The current study concentrates on the kinetically controlled ignition processes, especially in homogeneous mixtures, in order to exclude influences on the ignition due to inhomogeneities in the fuel–air mixture caused by diesel spray injection. As can be seen in Figure 3, the IDT calculated with MechB at an n-heptane content of about 9 mol% is underestimated for a temperature lower than 1000 K in which the reactions of the n-heptane sub-mechanism have a major impact on overall reactivity. Therefore, the changes made in the n-heptane sub-mechanism were re-evaluated. As mentioned before, some of the changes made to the Arrhenius parameters are large and partly outside the range that can be argued with data from the literature, which is an additional reason to re-evaluate the adjustments of all reactions considered.

2.1.1. Update of Methane Reaction Adaption

The methane reactions used for the adjustment in MechB are listed in Table 2. The largest change of the Arrhenius parameters was made in reaction 42. Reaction 44 and 45 were adapted only slightly to fine-tune the mechanism. To compare the reaction rate coefficients before and after adaption with an as large as possible number of reference data from literature, the NIST (National Institute of Standards and Technology) chemical kinetics database [17] was accessed. For reaction 42, 84 database entries [18,19,20,21,22,23,24,25,26,27,28,29,30,31,32,33,34,35,36,37,38,39,40,41,42,43,44,45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83,84,85,86,87,88,89,90] over the temperature range 178 to 3000 K were used. In the case of reaction 44 and 45, six [68,72,89,91,92,93] and three [68,72,94] entries, respectively, in the temperature range 300 to 2500 K were available. Using the reference data, the reaction rate coefficients for each data set in the specified temperature interval were calculated with a temperature resolution of 1 K. With these data points, a nonlinear fit based on the extended Arrhenius approach (1):
k = A · T β · e E A R · T
with the reaction rate coefficient k , the pre-exponential factor A , the temperature exponent β , the activation energy E A , the temperature T , and the universal gas constant R , was performed with the program Origin (OriginPro 2017, OriginLab Corporation, Northampton, MA, USA) [95], where Equation (1) has been used in the rewritten form (2):
k * = log ( k ) = A * + log [ ( 1000 x ) β ] E A · x 1000 · l n ( 10 ) · R
with x = 1000 T and A * = l o g ( A ) for the fitting procedure.
The effect of adjusting the Arrhenius parameters on the reaction rate coefficients is shown in Figure 4, Figure 5 and Figure 6. In addition to the reaction rate coefficients, the graphs also show the fit function as well as the 95% prediction band in green and the 95% confidence band in red of the fit function. The confidence band is mostly so narrow that it is usually covered by the fit curve. The data from the NIST database are represented as black open circles. Due to the plot density with a temperature resolution of 1 K, the NIST data points usually appear as black lines.
As can be seen in Figure 4, at temperatures in excess of approximately 400 K, the reaction rate coefficient of reaction 42 in MechB lies within the 95% prediction band of the fit function, but below all reference data. This was due to the fact that the target in [7] was to optimized the mechanism for a specific parameter range. However, since the focus of the current study is on developing a mechanism for a larger parameter range, the parameter adjustments carried out in [7] were revised so that for temperatures higher than 300 K the resulting temperature-dependent reaction rate coefficient of reaction 42 in the resulting mechanism MechC is within the range of reference data. In the case of reaction 44 and 45, only minor adjustments were made for fine-tuning, with the pre-exponential factors A remaining unchanged, the temperature exponent β being increased and decreased from 0 to 0.02 and –0.05, respectively, and the activation energy E A being reduced by 7.7% and 6.7%, respectively, compared to the original parameters of the San Diego mechanism.

2.1.2. Update of n-heptane Mechanism Adaption

Due to the difference between the measured and calculated IDT for mixtures containing about 9 mol% n-heptane (mix 9 and 10 in Table 1), a sensitivity analysis was performed with LOGEresearch using mix 9 to determine the most influential reactions to total reactivity. The starting point of the sensitivity analysis is found in Equation (3):
Y t = P ( Y , u , T ) + ω ( Y , T )
where Y is the n-dimensional vector of the mass fractions of the n species considered by the mechanism, T is the temperature, u the velocity field, P represents the spatial differential operator, and ω the chemical source term. In the course of the analysis, the influence of a small perturbation of the reaction rate coefficient on the concentration of a selected species is investigated. A detailed mathematical description can be found in [96,97]. All species considered by the mechanism, as well as the temperature, can be selected as an analysis target in LOGEresearch. Since the IDT- and temperature-sensitivity are correlated as shown by Ji et al. [98], the temperature was chosen as the analysis target in this study. For the sensitivity analysis, MechA was used to avoid any influence from the Arrhenius parameter adjustments made. As an example, Figure 7 shows the time-dependent sensitivity curve of reaction 274 (C7H16 + OH ↔ H2O + n-C7H15). In addition, the pressure curve and the first derivative of the pressure curve is shown. When searching for the maximum positive or negative sensitivity value, only the range of values between the simulation start and ignition (maximum of first derivative of pressure curve) of the mixture was considered. In the diagram, this range is highlighted in blue. The temperature-dependent course of the maximal sensitivity values can be seen in Figure 8 for the 16 most sensitive reactions.
Based on the differences between the measured and calculated IDTs and the course of sensitivities, the changes in Arrhenius parameters made to the n-heptane sub-mechanism were re-evaluated. As a result, instead of five n-heptane reactions, the number of reactions used for adaption could be reduced to one lumped reaction (reaction 302 in Figure 8), which considers the branching decomposition n-C7-OQOOH ↔ OH + CH2O + CO + C2H4 + n-C3H7. The change made in this reaction is a reduction of the activation energy by about 1%. A comparison of the reaction rate coefficients of reaction 302 in its original state (MechA), with the parameter changes transferred from [7] (MechB) and after revision of the parameter changes (MechC) is shown in Figure 9. Because this reaction is not included in the NIST database, no comparison with literature data is apparent.

2.2. Investigating the Influence of Propane Addition

In Figure 3, amongst others, the IDT of a methane–air mixture without propane addition and a propane admixture of 5, 10, and 30 mol% is shown. It can be observed that the deviation between measured and calculated IDTs increases with increasing propane concentration, suggesting that the IDT-reducing effect of the propane addition is overestimated by the mechanism. In order to determine which reactions have a significant effect on IDT in a methane–propane–air mixture, a sensitivity analysis towards temperature was performed with MechA. For the analysis, mix 4 in Table 1 with a propane content of 30 mol% was used. The left side of Figure 10 shows the temperature-dependent progressions of the 10 most sensitive reactions. For the sake of completeness, the measured IDTs and the values calculated with MechA for methane–propane mixtures with 0, 5, 10, and 30 mol% propane content (mix 1 to 4 in Table 1) are shown on the right-hand side of Figure 10.
Based on the course of the reaction sensitivities and the temperature-dependent deviation between measured and calculated IDTs for the various methane–propane mixtures, five reactions were selected for the mechanism adaptation. Figure 11 shows an extract of the propane oxidation scheme derived from MechA. The five reactions selected for adaptation, including the OH- and HO2-radical attack on propane, the oxygen addition on the propyl and hydroperoxyalkyl radical as well as the decomposition of carbonylhydroperoxide, are marked in red. In addition to the corresponding reaction numbers, the species with which the educts react are indicated. All reactions used for the adaption process are summarized in Table 3.
As described in the previous chapter, reference data from the NIST chemical kinetics database were used, if available, to evaluate the effect of changes in the Arrhenius parameters on the reaction rate coefficients. In the case of reaction 235, six entries [39,79,99,100,101,102] were available in the temperature range 250 to 1220 K. For reaction 238, two entries [103,104] could be found in the database for temperature values between 300 and 2500 K. However, for adjusting the Arrhenius parameters of reaction 238, the parameters used in the AramcoMech 3.0 [105] mechanism served as template. In total, nine sources [18,73,103,106,107,108,109,110,111] were available for comparison purposes for reaction 244 between 296 and 2000 K. No reference data were included in the NIST database for reactions 247 and 248, which is why a graphical comparison with reference data is missing. Therefore, care was taken not to make too large changes to the parameters. The pre-exponential factors and the temperature exponents of the reaction 247 and 248 remained unchanged, the absolute values of the activation energy were increased by 7.1% and 9.3%, respectively. Figure 12, Figure 13, Figure 14, Figure 15 and Figure 16 shows the temperature-dependent course of the reaction rate coefficients before (MechA) and after parameter adjustment (MechD).

2.3. Laminar Flame Speed Calculation

After ignition by injection of a diesel pilot in dual fuel combustion, the flame front spreads into the surrounding natural gas–air mixture in the combustion chamber. To correctly reproduce the propagation velocity of this front, the mechanism must be able to correctly calculate the laminar flame speed. In order to assess the extent to which the mechanism can reproduce the laminar flame speed, a comparison was made between simulated laminar flame speed values of methane–air mixtures (since methane is the main component of natural gas) and measurements taken from the literature. The LOGEresearch module “Freely Propagating” was used to simulate the flame speed. In addition to the gas composition, temperature, pressure, and fuel–air equivalence ratio, the gridding settings must be defined for the flame speed calculation. The grid used consisted of 1000 points with a minimum discretization size of 10−9 m and a maximum discretization size of 5 × 10−2 m. In the simulation, the effect of thermal diffusion was taken into account; the influence of radiation effects was neglected. The experimental determined flame speed values were taken from the supplementary materials of the publication of Ranzi et al. [112]. The gas mixtures investigated include methane–air mixtures with a start temperature of 300 to 400 K in the pressure range of 10 to about 20 bar and a fuel–air equivalence ratio between 0.7 and 1.4. The comparison between measured and calculated laminar flame speeds is shown in Figure 17.
For a fuel–air equivalence ratio of less than 0.9, the experimental values can be reproduced quite well by the simulation. When increasing the equivalence ratio, a deviation between the calculated and measured values is determinable, whereby the simulation underestimates the laminar flame speed. To improve the reproduction of the laminar flame speed, the “Freely Propagating” module from LOGEresearch was used to perform sensitivity analyses using MechD for detecting influential reactions to the flame speed. For each of the five parameter sets shown in Figure 17, a sensitivity analysis was performed and the 10 most sensitive reactions were filtered out. The reactions, which were among the 10 most sensitive reactions for all five investigated parameter sets, are shown in Figure 18 and Figure 19.
Based on the sensitivity progressions and the detected deviations between measured and calculated laminar flame speed, reactions 1, 25, and 56 (listed in Table 4) were selected for the mechanism adjustment.
In the case of reaction 56, not individual Arrhenius parameters were adapted, but the entire data set was changed to the dataset used in the mechanism AramcoMech 3.0, also including the third body collision coefficients and the LOW and TROE parameters. When adjusting the Arrhenius parameters of reaction 1 and 25, a comparison with reference data from the NIST database was made, with 68 entries [68,69,72,73,113,114,115,116,117,118,119,120,121,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168] for reaction 1 and 83 entries [25,26,36,37,43,51,68,73,74,75,134,137,140,164,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219] for reaction 25 available for this purpose in the temperature range 250 to 5500 K and 80 to 3150 K, respectively. Figure 20 and Figure 21 show the reaction rate coefficients before (MechD) and after adaption (TU Wien dual fuel mechanism 2.0) and the reference data from the NIST database. For the adjustment of the Arrhenius parameters of reaction 25, the reaction rate coefficients used in the mechanism of Zsely et al. [220] and Varga et al. [221] served as guideline, therefore the corresponding rate coefficients are presented in Figure 21.

3. Results and Discussion

In the course of the mechanism optimization process, in total the Arrhenius parameters of 12 reactions were adjusted. Table 5 provides an overview and a brief explanation why the respective reactions were used for the adaption process and which mechanism resulted from the individual adaptation steps.

3.1. IDT Calculation with TU Wien Dual Fuel Mechanism 2.0

The comparison of the IDT simulated with the Complete San Diego mechanism with heptane and nitrogen extension (MechA) with the mechanism MechB, in which the Arrhenius parameter changes carried out in [7] were transferred, was already shown in Figure 3. In the following Figure 22, the comparison between the mechanism MechB and the final version TU Wien dual fuel mechanism 2.0 is presented, in which the parameter changes carried out in MechB were re-evaluated, the propane influence on the IDT was corrected and the flame speed calculation was adapted.
By adapting the Arrhenius parameter of the reactions 235, 238, 244, 247, and 248, the TU Wien dual fuel mechanism 2.0 can reproduce the IDT-reducing effect of propane addition to a methane–air mixture well as shown in the diagram in the upper left corner of Figure 22. With the maximum investigated addition of 30 mol% propane, at which a significant deviation between the simulated and measured values was detected before the mechanism adaptation, the calculated values now lie within the uncertainty of the measured values. For the most part, an approximation of the calculated values to the experimental data can also be seen for the other investigated methane–propane mixtures. In the case of mixtures with an n-heptane content of less than 5 mol%, an almost continuous approximation to the experimental values can be observed. Since the n-heptane sub-mechanism changes transferred from [7] were adjusted and approximated to the original parameters of the San Diego mechanism, the simulated IDTs of the mixtures with the highest investigated n-heptane content of approximately 9 mol% show an increase in the temperature range from 850 to 1100 K, whereby the measured values at 950 K are therefore reproduced well, but between 1010 and 1080 K the IDT is overestimated.

3.2. Laminar Flame Speed Calculation with TU Wien Dual Fuel Mechanism 2.0

To adapt the laminar flame speed calculation, the Arrhenius parameters of the reactions 1, 25, and 56 were modified. The changes made had only a marginal effect on the calculated IDTs in the investigated parameter range, therefore the representation of the IDTs calculated with MechD is skipped. By the reaction parameter adjustment, the deviation between the measured and calculated laminar flame speed values could be reduced significantly. A comparison of the simulation results before and after the adjustment of the Arrhenius parameters of reaction 1, 25, and 56 is shown in Figure 23.

3.3. Influence of Nitrogen Sub-Mechanism on IDT and Laminar Flame Speed Calculation

To determine the influence of the nitrogen sub-mechanism on the calculation of the IDT and the laminar flame speed, the nitrogen sub-mechanism was removed from the TU Wien dual fuel mechanism 2.0 for test purposes, followed by a re-run of the simulations. In Figure 24 and Figure 25, the simulation results with the TU Wien dual fuel mechanism 2.0 are shown as lines, the simulation results with the mechanism without the nitrogen sub-mechanism are indicated with symbols.
The comparison shows that the nitrogen sub-mechanism has a negligible influence on the calculation of the IDT in the parameter range under consideration. The same applies to the calculation of the laminar flame speed, where no differences can be seen when the nitrogen sub-mechanism is taken into account or neglected.
It can be stated that by the modification of the TU Wien mechanism for dual fuel combustion presented in [7], the mechanism TU Wien dual fuel mechanism 2.0 was developed, which can reproduce the IDT of the investigated homogeneous methane, methane–propane, methane–n-heptane, and methane–propane–n-heptane mixtures in the pressure range 60 to 100 bar well and provides improved results in the calculation of the laminar flame speed of methane–air mixtures. The mechanism and associated transport and thermodynamic files are available online and can be downloaded from supplementary materials.

4. Conclusions

To describe the dual fuel combustion process and the ignition by injection of a diesel pilot into a natural gas–air mixture, the TU Wien mechanism for dual fuel combustion based on the Complete San Diego mechanism with heptane extension was developed and presented in a previous study. With this mechanism, it is possible to reproduce pressure curves measured with a rapid compression expansion machine operated in dual fuel mode using diesel and natural gas, satisfactorily. The mechanism was attuned to the rapid compression expansion machine measurements to achieve a high agreement between experiments and simulations, which results in a specific field of application for the mechanism. In order to simulate the ignition and combustion process within a manageable period of time, diesel and natural gas had to be replaced by substitutes due to the large number of species contained in the fuels. n-heptane was used as a one-component substitute of diesel, a mixture of methane and propane served as natural gas surrogate.
To obtain a mechanism applicable for a wider parameter range as well as to be able to calculate the formation of NOx during combustion, the TU Wien dual fuel mechanism has been extended and further developed, leading to the TU Wien dual fuel mechanism 2.0 presented in this study. The focus in the optimization process was on the kinetically controlled ignition process. Therefore, the ignition properties of homogeneous mixtures with different proportions of methane, propane, and n-heptane were investigated as shown in Table 1, whereby, depending on the composition, the fuel–air equivalence ratio varied between 0.526 and 0.816, the pressure value between 60 and 100 bar, and the temperature ranged from 671 to 1284 K. The simulated values were compared with rapid compression machine and shock tube measurement results. As a first step, the Arrhenius parameter changes made in the previous study were revised to ensure that the reaction rate coefficient changes are supported by literature data taken from the National Institute of Standards and Technology chemical kinetics database. Since the ignition delay time-reducing effect of propane addition to a methane–air mixture was overestimated, by means of a sensitivity analysis using a mixture consisting of 70 mol% methane and 30 mol% propane, reactions with a significant impact on the ignition delay time were determined. By adapting the Arrhenius parameters of five reactions from the propane oxidation scheme, the overestimated ignition delay time-reducing effect of propane addition to methane–air mixtures was corrected.
After ignition by injecting a diesel pilot into the natural gas–air mixture in the dual fuel combustion process, the flame front spreads into the entire combustion chamber. In order to reproduce the flame propagation precisely, the mechanism must be able to correctly calculate the laminar flame speed. Since the background gas in the combustion chamber (mixture of natural gas and air) consists mainly of methane, the laminar flame speed of methane–air mixtures at pressures between 10 and about 20 bar, temperatures of 300 to 400 K and fuel–air equivalence ratios in the range of 0.7 to 1.4 was simulated and compared with experimental values from literature. The comparison revealed an underestimation of the laminar flame speed by the simulation at a fuel–air equivalence ratio higher than 0.9. By means of sensitivity analyses, those reactions were determined, which were suitable for adapting the mechanism in the corresponding fuel–air equivalence ratio range. A total of three reactions were used for the laminar flame speed adaptation process.
In both the ignition delay time calculation and the calculation of the laminar flame speed, a partly significant reduction of the difference between measurement and simulation results could be achieved by adapting the original mechanism. Especially, the improved consideration of the ignition delay time-reducing effect of propane addition to methane–air mixtures should be emphasized. Despite its compactness, the mechanism, which consists of 344 reactions and takes into account 75 species, is able to approximate the ignition delay time well over a large part of the investigated parameter range. With measured values between 0.06 ms and 92 ms, ignition delay times over about three orders of magnitude are covered.
In order not to restrict the possible field of applications of the mechanism by the optimization process, a careful adjustment of the Arrhenius parameters was performed to ensure that the resulting changes in the reaction rate coefficients could be justified by means of literature data. Therefore, more than 200 entries from the National Institute of Standards and Technology database were used to compare the adjusted reaction rate coefficients with data from the literature, where available. By specifying the temperature dependent courses of all modified reaction rate coefficients in this article in comparison to the available reference data, the user of the mechanism is able to determine and assess to what extent the mechanism is suitable for use under the envisaged condition.

Supplementary Materials

The following are available online at https://www.mdpi.com/1996-1073/13/4/778/s1, Adapted reaction mechanism and associated thermodynamic- and transport data (ZIP).

Author Contributions

Conceptualization, S.S. and F.W.; methodology, S.S.; software, S.S.; validation, S.S.; formal analysis, S.S.; investigation, S.S.; resources, S.S.; data curation, S.S. and F.W.; writing—original draft preparation, S.S.; writing—review and editing, S.S. and F.W.; visualization, S.S.; supervision, F.W.; project administration, F.W.; funding acquisition, F.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Austrian Research Promotion Agency (FFG), grant number 850690.

Acknowledgments

We would like to show our gratitude to LOGE AB for providing the simulation software LOGEresearch. Open Access Funding by TU Wien.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Paris Agreement. Available online: https://ec.europa.eu/clima/policies/international/negotiations/paris_en (accessed on 20 November 2019).
  2. Ge, J.C.; Kim, H.Y.; Yoon, S.K.; Choi, N.J. Reducing volatile organic compound emissions from diesel engines using canola oil biodiesel fuel and blends. Fuel 2018, 218, 266–274. [Google Scholar] [CrossRef]
  3. Anto, S.; Mukherjee, S.S.; Muthappa, R.; Mathimani, T.; Deviram, G.; Kumar, S.S.; Verma, T.N.; Pugazhendhi, A. Algae as green energy reserve: Technological outlook on biofuel production. Chemosphere 2019, 242, 125079. [Google Scholar] [CrossRef] [PubMed]
  4. Schlick, H. Potentials and challenges of gas and dual-fuel engines for marine application. In Proceedings of the 5th CIMAC CASCADES, Busan, Korea, 23 October 2014. [Google Scholar]
  5. Yoshimoto, Y.; Kinoshita, E.; Luge, S.; Ohmura, T. Combustion Characteristics of a Dual Fuel Diesel Engine with Natural Gas (Lower limit of Cetane Number for Ignition of the Fuel). SAE Int. J. Fuels Lubr. 2012, 5, 1165–1173. [Google Scholar] [CrossRef]
  6. Understanding Global Warming Potentials. Available online: https://www.epa.gov/ghgemissions/understanding-global-warming-potentials (accessed on 9 December 2019).
  7. Schuh, S.; Frühhaber, J.; Lauer, T.; Winter, F. A Novel Dual Fuel Reaction Mechanism for Ignition in Natural Gas–Diesel Combustion. Energies 2019, 12, 4396. [Google Scholar] [CrossRef] [Green Version]
  8. Eder, L.; Ban, M.; Pirker, G.; Vujanovic, M.; Priesching, P.; Wimmer, A. Development and Validation of 3D-CFD Injection and Combustion Models for Dual Fuel Combustion in Diesel Ignited Large Gas Engines. Energies 2018, 11, 643. [Google Scholar] [CrossRef] [Green Version]
  9. Ritzke, J.; Andree, S.; Theile, M.; Henke, B.; Schleef, K.; Nocke, J.; Hassel, E. Simulation of a Dual-Fuel Large Marine Engines using combined 01-D and 3-D Approaches. In Proceedings of the CIMAC Congress, Helsinki, Finland, 6–10 June 2016. [Google Scholar]
  10. Li, Y.; Li, H.; Guo, H.; Li, Y.; Yao, M. A numerical investigation on methane combustion and emissions from a natural gas-diesel dual fuel engine using CFD model. Appl. Energy 2017, 205, 153–162. [Google Scholar] [CrossRef]
  11. Tschöke, H. Diesel- und Benzindirekteinspritzung V: Spraybildung, Simulation, Applikation, Messtechnik: Mit 29 Tabellen; Expert-Verlag: Renningen, Germany, 2009. [Google Scholar]
  12. Schuh, S.; Ramalingam, A.K.; Minwegen, H.; Heufer, K.A.; Winter, F. Experimental Investigation and Benchmark Study of Oxidation of Methane–Propane–n-Heptane Mixtures at Pressures up to 100 bar. Energies 2019, 12, 3410. [Google Scholar] [CrossRef] [Green Version]
  13. Pachler, R.F.; Ramalingam, A.K.; Heufer, K.A.; Winter, F. Reduction and validation of a chemical kinetic mechanism including necessity analysis and investigation of CH4/C3H8 oxidation at pressures up to 120 bar using a rapid compression machine. Fuel 2016, 172, 139–145. [Google Scholar] [CrossRef]
  14. LOGEsoft. Available online: https://logesoft.com/ (accessed on 16 May 2019).
  15. Sung, C.-J.; Curran, H.J. Using rapid compression machines for chemical kinetics studies. Prog. Energy Combust. Sci. 2014, 44, 1–18. [Google Scholar] [CrossRef]
  16. Chemical-Kinetic Mechanisms for Combustion Applications, San Diego Mechanism Web Page, Mechanical and Aerospace Engineering (Combustion Research), University of California at San Diego. Available online: http://combustion.ucsd.edu (accessed on 4 October 2019).
  17. Manion, J.A.; Huie, R.E.; Levin, R.D.; Burgess, D.R., Jr.; Orkin, V.L.; Tsang, W.; McGivern, W.S.; Hudgens, J.W.; Knyazev, V.D.; Atkinson, D.B.; et al. NIST Chemical Kinetics Database, NIST Standard Reference Database 17; Version 7.0 (Web Version), Data Version 2015.09; National Institute of Standards and Technology: Gaithersburg, MD, USA, 2015.
  18. Baker, R.R.; Baldwin, R.R.; Walker, R.W. The use of the H2+O2 reaction in determining the velocity constants of elementary reaction in hydrocarbon oxidation. Symp. (Int.) Combust. 1971, 13, 291–299. [Google Scholar] [CrossRef]
  19. Baulch, D.L.; Craven, R.J.B.; Din, M.; Drysdale, D.D.; Grant, S.; Richardson, D.J.; Walker, A.; Watling, G. Rates of hydroxy radical reactions with methane, ethane and propane over the temperature range 403–696 K. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1983, 79, 689–698. [Google Scholar] [CrossRef]
  20. Bonard, A.; Daële, V.; Delfau, J.-L.; Vovelle, C. Kinetics of OH Radical Reactions with Methane in the Temperature Range 295−660 K and with Dimethyl Ether and Methyl-tert-butyl Ether in the Temperature Range 295−618 K. J. Phys. Chem. A 2002, 106, 4384–4389. [Google Scholar] [CrossRef]
  21. Bott, J.F.; Cohen, N. A shock tube study of the reaction of the hydroxyl radical with H2, CH4, c-C5H10, and i-C4H10. Int. J. Chem. Kinet. 1989, 21, 485–498. [Google Scholar] [CrossRef]
  22. Bryukov, M.G.; Knyazev, V.D.; Lomnicki, S.M.; McFerrin, C.A.; Dellinger, B. Temperature-Dependent Kinetics of the Gas-Phase Reactions of OH with Cl2, CH4, and C3H8. J. Phys. Chem. A 2004, 108, 10464–10472. [Google Scholar] [CrossRef]
  23. Calpini, B.; Jeanneret, F.; Bourqui, M.; Clappier, A.; Vajtai, R.; van den Bergh, H. Direct measurement of the total reaction rate of OH in the atmosphere. Analusis 1999, 27, 328–336. [Google Scholar] [CrossRef] [Green Version]
  24. Cox, R.A.; Derwent, R.G.; Eggleton, A.E.J.; Lovelock, J.E. Photochemical oxidation of halocarbons in the troposphere. Atmos. Environ. (1967) 1976, 10, 305–308. [Google Scholar] [CrossRef]
  25. Cox, R.A.; Derwent, R.G.; Holt, P.M. Relative rate constants for the reactions of OH radicals with H2, CH4, CO, NO and HONO at atmospheric pressure and 296 K. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1976, 72, 2031–2043. [Google Scholar] [CrossRef]
  26. Davis, D.D.; Fischer, S.; Schiff, R. Flash photolysis-resonance fluorescence kinetics study: Temperature dependence of the reactions OH + CO → CO2 + H and OH + CH4 → H2O + CH3. J. Chem. Phys. 1974, 61, 2213–2219. [Google Scholar] [CrossRef]
  27. Dixon-Lewis, G.; Williams, A. Some observations on the combustion of methane in premixed flames. Symp. (Int.) Combust. 1967, 11, 951–958. [Google Scholar] [CrossRef]
  28. Dunlop, J.R.; Tully, F.P. A kinetic study of hydroxyl radical reactions with methane and perdeuterated methane. J. Phys. Chem. 1993, 97, 11148–11150. [Google Scholar] [CrossRef]
  29. Ernst, J.; Wagner, H.G.; Zellner, R. A Combined Flash Photolysis/Shock-Tube Study of the Absolute Rate Constants for Reactions of the Hydroxyl Radical with CH4 and CF3H around 1300 K. Ber. Bunsenges. Phys. Chem. 1978, 82, 409–414. [Google Scholar] [CrossRef]
  30. Fairchild, P.W.; Smith, G.P.; Crosley, D.R. A laser pyrolysis/laser fluorescence technique for combustion chemical kinetics. Symp. (Int.) Combust. 1982, 19, 107–115. [Google Scholar] [CrossRef]
  31. Felder, W.; Madronich, S. High Temperature Photochemistry (HTP): Kinetics and Mechanism Studies of Elementary Combustion Reactions over 300-1700 K. Combust. Sci. Technol. 1986, 50, 135–150. [Google Scholar] [CrossRef]
  32. Fenimore, C.P.; Jones, G.W. Rate of reaction of methane with H atoms and OH radicals in flames. J. Phys. Chem. 1961, 65, 2200–2203. [Google Scholar] [CrossRef]
  33. Finlayson-Pitts, B.J.; Ezell, M.J.; Jayaweera, T.M.; Berko, H.N.; Lai, C.C. Kinetics of the reactions of OH with methyl chloroform and methane: Implications for global tropospheric OH and the methane budget. Geophys. Res. Lett. 1992, 19, 1371–1374. [Google Scholar] [CrossRef] [Green Version]
  34. Fristrom, R.M. Radical concentrations and reactions in a methane-oxygen flame. Symp. (Int.) Combust. 1963, 9, 560–575. [Google Scholar] [CrossRef]
  35. Gierczak, T.; Talukdar, R.K.; Herndon, S.C.; Vaghjiani, G.L.; Ravishankara, A.R. Rate Coefficients for the Reactions of Hydroxyl Radicals with Methane and Deuterated Methanes. J. Phys. Chem. A 1997, 101, 3125–3134. [Google Scholar] [CrossRef]
  36. Gordon, S.; Mulac, W.A. Reaction of the OH (X2π) Radical Produced by the Pulse Radiolysis of Water Vapor. Proc. Symp. Chem. Kinet. Data Upper Lower Atmos. 1975, 7, 289–299. [Google Scholar]
  37. Greiner, N.R. Hydroxyl-Radical Kinetics by Kinetic Spectroscopy. I. Reactions with H2, CO, and CH4 at 300 °K. J. Chem. Phys. 1967, 46, 2795–2799. [Google Scholar] [CrossRef]
  38. Greiner, N.R. Hydroxyl Radical Kinetics by Kinetic Spectroscopy. IV. Some Deuterium Isotope Effects. J. Chem. Phys. 1968, 48, 1413. [Google Scholar] [CrossRef]
  39. Greiner, N.R. Hydroxyl Radical Kinetics by Kinetic Spectroscopy. VI. Reactions with Alkanes in the Range 300–500 °K. J. Chem. Phys. 1970, 53, 1070–1076. [Google Scholar] [CrossRef]
  40. Hong, Z.; Davidson, D.F.; Lam, K.-Y.; Hanson, R.K. A shock tube study of the rate constants of HO2 and CH3 reactions. Combust. Flame 2012, 159, 3007–3013. [Google Scholar] [CrossRef]
  41. Horne, D.G.; Norrish, R.G.W. Rate of H-abstraction by OH from Hydrocarbons. Nature 1967, 215, 1373–1374. [Google Scholar] [CrossRef]
  42. Howard, C.J.; Evenson, K.M. Rate constants for the reactions of OH with CH4 and fluorine, chlorine, and bromine substituted methanes at 296 K. J. Chem. Phys. 1976, 64, 197–202. [Google Scholar] [CrossRef]
  43. Husain, D.; Plane, J.M.; Slater, N.K.H. Kinetic investigation of the reactions of OH (X2π) with the hydrogen halides, HCl, DCl, HBr and DBr by time-resolved resonance fluorescence (A2Σ+−X2π). J. Chem. Soc. Faraday Trans. 2 1981, 77, 1949–1962. [Google Scholar] [CrossRef]
  44. Jeong, K.M.; Kaufman, F. Kinetics of the reaction of hydroxyl radical with methane and with nine chlorine- and fluorine-substituted methanes. 1. Experimental results, comparisons, and applications. J. Phys. Chem. 1982, 86, 1808–1815. [Google Scholar] [CrossRef]
  45. Jonah, C.D.; Mulac, W.A.; Zeglinski, P. Rate Constants for the Reaction of OH + CO, OD + CO, and OH + Methane as a Function of Temperature. J. Phys. Chem. 1984, 88, 4100–4104. [Google Scholar] [CrossRef]
  46. Lancar, I.T.; LeBras, G.; Poulet, G. Redetermination de la constante de vitesse de la reaction CH4 + OH et son implication atmospherique. C.R. Acad. Sc. Paris 1992, 315, 1487–1492. [Google Scholar]
  47. Madronich, S.; Felder, W. Direct measurements of the rate coefficient for the reaction OH+CH4 → CH3+H2O over 300-1500 K. Symp. (Int.) Combust. 1985, 20, 703–713. [Google Scholar] [CrossRef]
  48. Margitan, J.J.; Kaufman, F.; Anderson, J.G. The reaction of OH with CH4. Geophys. Res. Lett. 1974, 1, 80–81. [Google Scholar] [CrossRef]
  49. Mellouki, A.; Teton, S.; Laverdet, G.; Quilgars, A.; Le Bras, G. Kinetic studies of OH reactions with H2O, C3H8 and CH4 using the pulsed laser photolysis—laser induced fluorescence method. J. Chim. Phys. 1994, 91, 473–487. [Google Scholar] [CrossRef]
  50. Overend, R.P.; Paraskevopoulos, G.; Cvetanović, R.J. Rates of OH Radical Reactions. I. Reactions with H2, CH4, C2H6, and C3H8 at 295 K. Can. J. Chem. 1975, 53, 3374–3382. [Google Scholar] [CrossRef]
  51. Peeters, J.; Mahnen, G. Reaction mechanisms and rate constants ofelementary steps in methane-oxygen flames. Symp. (Int.) Combust. 1973, 14, 133–146. [Google Scholar] [CrossRef]
  52. Sharkey, P.; Smith, I.W.M. Kinetics of elementary reactions at low temperatures: Rate constants for the reactions of OH with HCl (298 ⩾T/K ⩾ 138), CH4(298 ⩾T/K ⩾ 178) and C2H6(298 ⩾T/K ⩾ 138). J. Chem. Soc. Faraday Trans. 1993, 89, 631–637. [Google Scholar] [CrossRef]
  53. Smith, G.P.; Fairchild, P.W.; Jeffries, J.B.; Crosley, D.R. Laser pyrolysis/laser fluorescence studies of high-temperature reaction rates: Description of the method and results for OH + CH4, C3H8, and C3H6. J. Phys. Chem. 1985, 89, 1269–1278. [Google Scholar] [CrossRef]
  54. Srinivasan, N.K.; Su, M.C.; Sutherland, J.W.; Michael, J.V. Reflected Shock Tube Studies of High-Temperature Rate Constants for OH + CH4 → CH3 + H2O and CH3 + NO2 → CH3O + NO. J. Phys. Chem. A 2005, 109, 1857–1863. [Google Scholar] [CrossRef] [PubMed]
  55. Steinert, W.; Zellner, R. Rates of Reaction of OH with CO and CH4 over an Extended Temperature Range. Deuxiene Symp. Eur. Combust. 1975, 2, 31. [Google Scholar]
  56. Sworski, T.J.; Hochanadel, C.J.; Ogren, P.J. Flash photolysis of water vapor in methane. Hydrogen and hydroxyl yields and rate constants for methyl reactions with hydrogen and hydroxyl. J. Phys. Chem. 1980, 84, 129–134. [Google Scholar] [CrossRef]
  57. Tully, F.P.; Ravishankara, A.R. Flash photolysis-resonance fluorescence kinetic study of the reactions hydroxyl + molecular hydrogen → water + atomic hydrogen and hydroxyl + methane → water + methyl from 298 to 1020 K. J. Phys. Chem. 1980, 84, 3126–3130. [Google Scholar] [CrossRef]
  58. Vaghjiani, G.L.; Ravishankara, A.R. New measurement of the rate coefficient for the reaction of OH with methane. Nature 1991, 350, 406–409. [Google Scholar] [CrossRef]
  59. Wilson, W.E.; Westenberg, A.A. Study of the reaction of hydroxyl radical with methane by quantitative ESR. Symp. (Int.) Combust. 1967, 11, 1143–1150. [Google Scholar] [CrossRef]
  60. Yetter, R.A.; Dryer, F.L. Inhibition of moist carbon monoxide oxidation by trace amounts of hydrocarbons. Symp. (Int.) Combust. 1992, 24, 757–767. [Google Scholar] [CrossRef] [Green Version]
  61. Zellner, R. Rate Measurements of Some Bimolecular Reactions of the Hydroxyl Radical over an Extended Temperature Range. Mol. Rate Process. Pap. Symp. 1975, 7, 1. [Google Scholar]
  62. Zellner, R.; Steinert, W. A flash photolysis study of the rate of the reaction OH + CH4 → CH3 + H2O over an extended temperature range. Int. J. Chem. Kinet. 1976, 8, 397–409. [Google Scholar] [CrossRef]
  63. Atkinson, R. Kinetics and mechanisms of the gas-phase reactions of the hydroxyl radical with organic compounds under atmospheric conditions. Chem. Rev. 1986, 86, 69–201. [Google Scholar] [CrossRef]
  64. Atkinson, R.; Baulch, D.L.; Cox, R.A.; Hampson, R.F.; Kerr, J.A.; Rossi, M.J.; Troe, J. Evaluated Kinetic, Photochemical and Heterogeneous Data for Atmospheric Chemistry: Supplement V.  IUPAC Subcommittee on Gas Kinetic Data Evaluation for Atmospheric Chemistry. J. Phys. Chem. Ref. Data 1997, 26, 521–1011. [Google Scholar] [CrossRef]
  65. Atkinson, R.; Baulch, D.L.; Cox, R.A.; Hampson, R.F.; Kerr, J.A.; Troe, J. Evaluated Kinetic and Photochemical Data for Atmospheric Chemistry: Supplement III. IUPAC Subcommittee on Gas Kinetic Data Evaluation for Atmospheric Chemistry. J. Phys. Chem. Ref. Data 1989, 18, 881–1097. [Google Scholar] [CrossRef]
  66. Atkinson, R.; Baulch, D.L.; Cox, R.A.; Hampson, R.F.; Kerr, J.A.; Troe, J. Evaluated Kinetic and Photochemical Data for Atmospheric Chemistry: Supplement IV. IUPAC Subcommittee on Gas Kinetic Data Evaluation for Atmospheric Chemistry. J. Phys. Chem. Ref. Data 1992, 21, 1125–1568. [Google Scholar] [CrossRef]
  67. Baulch, D.L.; Bowers, M.; Malcolm, D.G.; Tuckerman, R.T. Evaluated Kinetic Data for High-Temperature Reactions. Volume 5. Part 1. Homogeneous Gas Phase Reactions of the Hydroxyl Radical with Alkanes. J. Phys. Chem. Ref. Data 1986, 15, 465–592. [Google Scholar] [CrossRef]
  68. Baulch, D.L.; Cobos, C.J.; Cox, R.A.; Esser, C.; Frank, P.; Just, T.; Kerr, J.A.; Pilling, M.J.; Troe, J.; Walker, R.W.; et al. Evaluated Kinetic Data for Combustion Modelling. J. Phys. Chem. Ref. Data 1992, 21, 411–734. [Google Scholar] [CrossRef] [Green Version]
  69. Cohen, N.; Westberg, K.R. Chemical Kinetic Data Sheets for High-Temperature Chemical Reactions. J. Phys. Chem. Ref. Data 1983, 12, 531–590. [Google Scholar] [CrossRef] [Green Version]
  70. DeMore, W.B.; Sander, S.P.; Golden, D.M.; Hampson, R.F.; Kurylo, M.J.; Howard, C.J.; Ravishankara, A.R.; Kolb, C.E.; Molina, M.J. Chemical Kinetics and Photochemical Data for Use in Stratospheric Modeling: Evaluation Number 11; Jet Propulsion Lab., California Inst. of Tech.: Pasadena, CA, USA, 1994; p. 18. [Google Scholar]
  71. DeMore, W.B.; Sander, S.P.; Golden, D.M.; Hampson, R.F.; Kurylo, M.J.; Howard, C.J.; Ravishankara, A.R.; Kolb, C.E.; Molina, M.J. Chemical Kinetics and Photochemical Data for Use in Stratospheric Modeling. Evaluation No. 12; Jet Propulsion Lab., California Inst. of Tech.: Pasadena, CA, USA, 1997; p. 20. [Google Scholar]
  72. Tsang, W.; Hampson, R.F. Chemical Kinetic Data Base for Combustion Chemistry. Part I. Methane and Related Compounds. J. Phys. Chem. Ref. Data 1986, 15, 1087–1279. [Google Scholar] [CrossRef]
  73. Warnatz, J. Rate Coefficients in the C/H/O System. In Combustion Chemistry; Gardiner, W.C., Ed.; Springer: New York, NY, USA, 1984; pp. 197–360. [Google Scholar] [CrossRef]
  74. Westenberg, A.A.; Fristrom, R.M. Methane-oxygen flame structure. iv. chemical kinetic considerations. J. Phys. Chem. 1961, 65, 591–601. [Google Scholar] [CrossRef]
  75. Wilson, W.E. A Critical Review of the Gas-Phase Reaction Kinetics of the Hydroxyl Radical. J. Phys. Chem. Ref. Data 1972, 1, 535–573. [Google Scholar] [CrossRef] [Green Version]
  76. Zellner, R. Non-Arrhenius behavior in bimolecular reactions of the hydroxyl radical. J. Phys. Chem. 1979, 83, 18–23. [Google Scholar] [CrossRef]
  77. Bravo-Pérez, G.; Alvarez-Idaboy, J.R.; Jiménez, A.G.; Cruz-Torres, A. Quantum chemical and conventional TST calculations of rate constants for the OH+alkane reaction. Chem. Phys. 2005, 310, 213–223. [Google Scholar] [CrossRef]
  78. Cohen, N. The use of transition-state theory to extrapolate rate coefficients for reactions of oh with alkanes. Int. J. Chem. Kinet. 1982, 14, 1339–1362. [Google Scholar] [CrossRef]
  79. Cohen, N. Are reaction rate coefficients additive? Revised transition state theory calculations for OH + alkane reactions. Int. J. Chem. Kinet. 1991, 23, 397–417. [Google Scholar] [CrossRef]
  80. Cohen, N.; Benson, S.W. Transition-state-theory calculations for reactions of hydroxyl radicals with haloalkanes. J. Phys. Chem. 1987, 91, 162–170. [Google Scholar] [CrossRef]
  81. Dobbs, K.D.; Dixon, D.A.; Komornicki, A. Ab initio prediction of the barrier height for abstraction of H from CH4 by OH. J. Chem. Phys. 1993, 98, 8852–8858. [Google Scholar] [CrossRef]
  82. Gaillard-Cusin, F.; James, H.; Rouan, J.-P. No. 105.—Le phenomene d’inhibition-promotion par le methane de la combustion de hautes pressions de l’oxyde de carbone. J. Chim. Phys. 1969, 66, 751–756. [Google Scholar] [CrossRef]
  83. Gonzalez, C.; McDouall, J.J.W.; Schlegel, H.B. Ab initio study of the reactions between methane and hydroxyl, hydrogen atom, and triplet oxygen atom. J. Phys. Chem. 1990, 94, 7467–7471. [Google Scholar] [CrossRef]
  84. Leroy, G.; Sana, M.; Tinant, A. Etude theorique des reactions d’abstraction d’hydrogene RH + X = R + HX, avec R, X≡H, CH3, NH2, OH et F. Can. J. Chem. 1985, 63, 1447–1456. [Google Scholar] [CrossRef]
  85. Masgrau, L.; González-Lafont, À.; Lluch, J.M. The reactions CHnD4−n+OH→P and CH4+OD→CH3+HOD as a test of current direct dynamics multicoefficient methods to determine variational transition state rate constants. II. J. Chem. Phys. 2001, 115, 4515–4526. [Google Scholar] [CrossRef]
  86. Masgrau, L.; González-Lafont, À.; Lluch, J.M. The reactions CHnD4−n+OH→P and CH4+OD→CH3+HOD as a test of current direct dynamics computational methods to determine variational transition-state rate constants. I. J. Chem. Phys. 2001, 114, 2154–2165. [Google Scholar] [CrossRef]
  87. Melissas, V.S.; Truhlar, D.G. Interpolated variational transition state theory and tunneling calculations of the rate constant of the reaction OH+CH4 at 223–2400 K. J. Chem. Phys. 1993, 99, 1013–1027. [Google Scholar] [CrossRef]
  88. Schwartz, M.; Marshall, P.; Berry, R.J.; Ehlers, C.J.; Petersson, G.A. Computational Study of the Kinetics of Hydrogen Abstraction from Fluoromethanes by the Hydroxyl Radical. J. Phys. Chem. A 1998, 102, 10074–10081. [Google Scholar] [CrossRef] [Green Version]
  89. Shaw, R. Semi-empirical extrapolation and estimation of rate constants for abstraction of H from methane by H, O, HO, and O2. J. Phys. Chem. Ref. Data 1978, 7, 1179–1190. [Google Scholar] [CrossRef] [Green Version]
  90. Truong, T.N.; Truhlar, D.G. Ab initio transition state theory calculations of the reaction rate for OH+CH4→H2O+CH3. Ber. Bunsenges. Phys. Chem. 1990, 93, 1761–1769. [Google Scholar] [CrossRef]
  91. Mayer, S.W.; Schieler, L. Activation energies and rate constants computed for reactions of oxygen with hydrocarbons. J. Phys. Chem. 1968, 72, 2628–2631. [Google Scholar] [CrossRef]
  92. Skinner, G.B.; Lifshitz, A.; Scheller, K.; Burcat, A. Kinetics of Methane Oxidation. J. Chem. Phys. 1972, 56, 3853–3861. [Google Scholar] [CrossRef]
  93. Yamaguchi, Y.; Teng, Y.; Shimomura, S.; Tabata, K.; Suzuki, E. Ab Initio Study for Selective Oxidation of Methane with NOx (x = 1, 2). J. Phys. Chem. A 1999, 103, 8272–8278. [Google Scholar] [CrossRef]
  94. Baldwin, R.R.; Jones, P.N.; Walker, R.W. Determination of the rate constant for HO2 + CH4 →H2O2 → H2O2 + CH3 at 443 °C. J. Chem. Soc. Faraday Trans. 2 1988, 84, 199–207. [Google Scholar] [CrossRef]
  95. OriginLab Corporation. OriginPro; Version 2017; OriginLab Corporation: Northampton, MA, USA, 2017. [Google Scholar]
  96. Soyhan, H.S.; Mauss, F.; Sorusbay, C. Chemical kinetic modeling of combustion in internal combustion engines using reduced chemistry. Combust. Sci. Technol. 2002, 174, 73–91. [Google Scholar] [CrossRef]
  97. Løvås, T.; Amnéus, P.; Mauss, F.; Mastorakos, E. Comparison of automatic reduction procedures for ignition chemistry. Proc. Combust. Inst. 2002, 29, 1387–1393. [Google Scholar] [CrossRef]
  98. Ji, W.; Ren, Z.; Law, C.K. Evolution of sensitivity directions during autoignition. Proc. Combust. Inst. 2019, 37, 807–815. [Google Scholar] [CrossRef]
  99. Darnall, K.R.; Atkinson, R.; Pitts, J.N. Rate constants for the reaction of the hydroxyl radical with selected alkanes at 300 K. J. Phys. Chem. 1978, 82, 1581–1584. [Google Scholar] [CrossRef]
  100. Atkinson, R.; Aschmann, S.M.; Carter, W.P.L.; Winer, A.M.; Pitts, J.N., Jr. Kinetics of the reactions of OH radicals with n-alkanes at 299 ± 2 K. Int. J. Chem. Kinet. 1982, 14, 781–788. [Google Scholar] [CrossRef]
  101. Atkinson, R. Estimations of OH radical rate constants from H-atom abstraction from C–H and O–H bonds over the temperature range 250–1000 K. Int. J. Chem. Kinet. 1986, 18, 555–568. [Google Scholar] [CrossRef]
  102. Hu, W.-P.; Rossi, I.; Corchado, J.C.; Truhlar, D.G. Molecular Modeling of Combustion Kinetics. The Abstraction of Primary and Secondary Hydrogens by Hydroxyl Radical. J. Phys. Chem. A 1997, 101, 6911–6921. [Google Scholar] [CrossRef]
  103. Tsang, W. Chemical Kinetic Data Base for Combustion Chemistry. Part 3: Propane. J. Phys. Chem. Ref. Data 1988, 17, 887–951. [Google Scholar] [CrossRef] [Green Version]
  104. Baldwin, R.R.; Fuller, A.R.; Longthorn, D.; Walker, R.W. Combustion Institute European Symposium; Academic Press: London, UK, 1973. [Google Scholar]
  105. Zhou, C.-W.; Li, Y.; Burke, U.; Banyon, C.; Somers, K.P.; Ding, S.; Khan, S.; Hargis, J.W.; Sikes, T.; Mathieu, O.; et al. An experimental and chemical kinetic modeling study of 1,3-butadiene combustion: Ignition delay time and laminar flame speed measurements. Combust. Flame 2018, 197, 423–438. [Google Scholar] [CrossRef]
  106. Baker, R.R.; Baldwin, R.R.; Walker, R.W. Velocity constants for the reactions of alkyl radicals with oxygen. J. Chem. Soc. D Chem. Commun. 1969, 1382–1383. [Google Scholar] [CrossRef]
  107. Baker, R.R.; Baldwin, R.R.; Walker, R.W. Addition of C3H8 to slowly reacting mixtures of hydrogen and oxygen at 480 °C. Trans. Faraday Soc. 1970, 66, 3016–3031. [Google Scholar] [CrossRef]
  108. Baldwin, R.R.; Langford, D.H.; Matchan, M.J.; Walker, R.W.; Yorke, D.A. The high-temperature oxidation of aldehydes. Symp. (Int.) Combust. 1971, 13, 251–259. [Google Scholar] [CrossRef]
  109. Baldwin, R.R.; Walker, R.W.; Yorke, D.A. Reaction of n-propyl radicals with oxygen, hydrogen and deuterium. J. Chem. Soc. Faraday Trans. 1 Phys. Chem. Condens. Phases 1973, 69, 826–832. [Google Scholar] [CrossRef]
  110. Slagle, I.R.; Park, J.-Y.; Gutman, D. Experimental investigation of the kinetics and mechanism of the reaction of n-propyl radicals with molecular oxygen from 297 to 635 K. Symp. (Int.) Combust. 1985, 20, 733–741. [Google Scholar] [CrossRef]
  111. DeSain, J.D.; Klippenstein, S.J.; Miller, J.A.; Taatjes, C.A. Measurements, Theory, and Modeling of OH Formation in Ethyl + O2 and Propyl + O2 Reactions. J. Phys. Chem. A 2003, 107, 4415–4427. [Google Scholar] [CrossRef]
  112. Ranzi, E.; Frassoldati, A.; Grana, R.; Cuoci, A.; Faravelli, T.; Kelley, A.P.; Law, C.K. Hierarchical and comparative kinetic modeling of laminar flame speeds of hydrocarbon and oxygenated fuels. Prog. Energy Combust. Sci. 2012, 38, 468–501. [Google Scholar] [CrossRef]
  113. Baldwin, R.R.; Melvin, A. 348. The reaction of hydrogen atoms with oxygen and with ethane. J. Chem. Soc. (Resumed) 1964, 1785–1791. [Google Scholar] [CrossRef]
  114. Baulch, D.L.; Cobos, C.J.; Cox, R.A.; Frank, P.; Hayman, G.; Just, T.; Kerr, J.A.; Murrells, T.; Pilling, M.J.; Troe, J.; et al. Evaluated Kinetic Data for Combustion Modeling. Supplement I. J. Phys. Chem. Ref. Data 1994, 23, 847–848. [Google Scholar] [CrossRef] [Green Version]
  115. Baulch, D.L.; Drysdale, D.D.; Horne, D.G. An assessment of rate data for high-temperaturesystems. Symp. (Int.) Combust. 1973, 14, 107–118. [Google Scholar] [CrossRef]
  116. Dixon-Lewis, G. Spherically Symmetric Flame Propagation in Hydrogen-Air Mixtures. Combust. Sci. Technol. 1983, 34, 1–29. [Google Scholar] [CrossRef]
  117. Du, H.; Hessler, J.P. Rate coefficient for the reaction H+O2→OH+O: Results at high temperatures, 2000 to 5300 K. J. Chem. Phys. 1992, 96, 1077–1092. [Google Scholar] [CrossRef]
  118. Kaufman, F.; Del Greco, F.P. Fast reactions of OH radicals. Symp. (Int.) Combust. 1963, 9, 659–668. [Google Scholar] [CrossRef]
  119. Masten, D.A.; Hanson, R.K.; Bowman, C.T. Shock tube study of the reaction hydrogen atom + oxygen → hydroxyl + oxygen atom using hydroxyl laser absorption. J. Phys. Chem. 1990, 94, 7119–7128. [Google Scholar] [CrossRef]
  120. Miller, J.A.; Pilling, M.J.; Troe, J. Unravelling combustion mechanisms through a quantitative understanding of elementary reactions. Proc. Combust. Inst. 2005, 30, 43–88. [Google Scholar] [CrossRef]
  121. Pirraglia, A.N.; Michael, J.V.; Sutherland, J.W.; Klemm, R.B. A flash photolysis-shock tube kinetic study of the hydrogen atom reaction with oxygen: H + O2 ⇄ OH + O (962 K ≤ T ≤ 1705 K) and H + O2 + Ar → HO2 + Ar (746 K ≤ T ≤ 987 K). J. Phys. Chem. 1989, 93, 282–291. [Google Scholar] [CrossRef]
  122. Semenov, N. On the Constants of the Reactions H+ O2== OH+ O and H2+ O2= 2OH. Acta Physicochim. URSS 1945, 20, 292. [Google Scholar]
  123. Turányi, T.; Nagy, T.; Zsély, I.G.; Cserháti, M.; Varga, T.; Szabó, B.T.; Sedyó, I.; Kiss, P.T.; Zempléni, A.; Curran, H.J. Determination of rate parameters based on both direct and indirect measurements. Int. J. Chem. Kinet. 2012, 44, 284–302. [Google Scholar] [CrossRef] [Green Version]
  124. Yang, H.; Gardiner, W.C.; Shin, K.S.; Fujii, N. Shock tube study of the rate coefficient of H + O2 → OH + O. Chem. Phys. Lett. 1994, 231, 449–453. [Google Scholar] [CrossRef]
  125. Yu, C.L.; Frenklach, M.; Masten, D.A.; Hanson, R.K.; Bowman, C.T. Reexamination of Shock-Tube Measurements of the Rate Coefficient of H + O2.fwdarw. OH + O. J. Phys. Chem. 1994, 98, 4770–4771. [Google Scholar] [CrossRef]
  126. Aganesyan, K.T.; Nalbandyan, A.B. The determination of the rate constants for the reactions between hydrogen and oxygen atoms and ammonia molecules. Dokl. Phys. Chem. (Engl. Transl.) 1965, 160, 18. [Google Scholar]
  127. Aleksandrov, E.N.; Arutyunov, V.S.; Dubrovina, I.V.; Kozlov, S.N. Kinetics of hydrogen oxidation near the lower explosion limit. Int. J. Chem. Kinet. 1984, 16, 817–834. [Google Scholar] [CrossRef]
  128. Azatyan, V.V. Combustion limitation method in the heterogeneous termination of chains in diffusion range. Arm. Khim. Z. 1967, 20, 577–584. [Google Scholar]
  129. Azatyan, V.V.; Nalbandyan, A.B.; Ts’ui, M.-Y. Determination of the rate constant of the reaction between atomic oxygen and ethane. Dokl. Akad. Nauk SSR 1962, 147, 361–364. [Google Scholar]
  130. Azatyan, V.V.; Voevodskii, V.V.; Nalbandyan, A.B. Determination of the rate constant of the reaction of atomic oxygen with molecular hydrogen. Kinet. Katal. 1961, 2, 340–349. [Google Scholar]
  131. Balakhnin, V.P.; Gershenzon, Y.M. A quantitative study of the mechanism of hydrogen combustion close to the lower explosion limit. Dokl. Phys. Chem. (Engl. Transl.) 1967, 170, 659–662. [Google Scholar]
  132. Baldwin, R.R. The first limit of the hydrogen + oxygen reaction in potassium chloride-coated vessels. Trans. Faraday Soc. 1956, 52, 1344–1354. [Google Scholar] [CrossRef]
  133. Belles, F.E.; Brabbs, T.A. Experimental verification of effects of turbulent Boundary layers on chemical-kinetic measurements in a shock tube. Symp. (Int.) Combust. 1971, 13, 165–175. [Google Scholar] [CrossRef] [Green Version]
  134. Biordi, J.C.; Lazzara, C.P.; Papp, J.F. Molecular beam mass spectrometry applied to determining the kinetics of reactions in flames II. A critique of rate coefficient determinations. Combust. Flame 1976, 26, 57–76. [Google Scholar] [CrossRef]
  135. Bowman, C.T. Non-equilibrium radical concentrations in shock-initiated methane oxidation. Symp. (Int.) Combust. 1975, 15, 869–882. [Google Scholar] [CrossRef]
  136. Brabbs, T.A.; Belles, F.E. Experimental Study of Effects of Laminar Boundary Layers on Chemical-Kinetic Measurements in a Shock Tube. Proc. Int. Symp. Shock Tubes Waves 1971, 67, 24. [Google Scholar]
  137. Brabbs, T.A.; Belles, F.E.; Brokaw, R.S. Shock-tube measurements of specific reaction rates in the branched-chain H2-CO-O2 system. Symp. (Int.) Combust. 1971, 13, 129–136. [Google Scholar] [CrossRef] [Green Version]
  138. Buneva, V.N.; Kabasheva, G.N.; Panfilov, V.N. Determination of the reaction rate constant of hydrogen splitting in a rarefied flame by means of isotope exchange. Kinet. Catal. 1969, 10, 1007. [Google Scholar]
  139. Chiang, C.-C.; Skinner, G.B. Resonance Absorption Measure- ments of Atom Concentrations in Reacting Gas Mixtures. 4. Measurements of H and D Atoms in Oxidationof H2, D2 and CD4. Proc. Int. Symp. Shock Tubes Waves 1980, 12, 629. [Google Scholar]
  140. Dean, A.M.; Kistiakowsky, G.B. Oxidation of Carbon Monoxide/Methane Mixtures in Shock Waves. J. Chem. Phys. 1971, 54, 1718–1725. [Google Scholar] [CrossRef]
  141. Dixon-Lewis, G.; Sutton, M.M.; Williams, A. Some reactions of hydrogen atoms and simple radicals at high temperatures. Symp. (Int.) Combust. 1965, 10, 495–502. [Google Scholar] [CrossRef]
  142. Eberius, K.H.; Hoyermann, K.; Wagner, H.G. Experimental and mathematical study of a hydrogen-oxygen flame. Symp. (Int.) Combust. 1971, 13, 713–721. [Google Scholar] [CrossRef]
  143. Fenimore, C.P.; Jones, G.W. Rate of Reaction in Hydrogen, Nitrous Oxide and in Some Other Flames. J. Phys. Chem. 1959, 63, 1154–1158. [Google Scholar] [CrossRef]
  144. Frank, P.; Just, T. High Temperature Reaction Rate for H + O2 = OH + O and OH + H2 = H2O + H. Ber. Bunsenges. Phys. Chem. 1985, 89, 181–187. [Google Scholar] [CrossRef]
  145. Fujii, N.; Shin, K.S. Rate constant for H+O2→O+OH by laser absorption spectroscopy of OH in shock-heated H2-O2-Ar mixtures. Chem. Phys. Lett. 1988, 151, 461–465. [Google Scholar] [CrossRef]
  146. Gutman, D.; Hardwidge, E.A.; Dougherty, F.A.; Lutz, R.W. Shock-Tube Study of the Recombination Rate of Hydrogen Atoms with Oxygen Molecules. J. Chem. Phys. 1967, 47, 4400–4407. [Google Scholar] [CrossRef]
  147. Gutman, D.; Schott, G.L. Shock-Tube Study of Chain Branching during the Induction Period of the Hydrogen—Oxygen Reaction. J. Chem. Phys. 1967, 46, 4576–4584. [Google Scholar] [CrossRef]
  148. Jachimowski, C.J.; Houghton, W.M. Shock-tube study of the induction-period kinetics of the hydrogen-oxygen reaction. Combust. Flame 1970, 15, 125–132. [Google Scholar] [CrossRef]
  149. Karmilova, L.V.; Nalbandyan, A.B.; Semenov, N.N. Hydrogen combustion kinetics with oxygen below the lower autoignition temperature. Z. Fiz. Khim. 1958, 32, 1193–1204. [Google Scholar]
  150. Kochubei, V.F.; Moin, F.B. Kinetics of the Reaction of Atomic Hydrogen with Oxygen. Sov. Prog. Chem. (Engl. Transl.) 1973, 39, 88. [Google Scholar]
  151. Kurzius, S.C.; Boudart, M. Kinetics of the branching step in the hydrogen-oxygen reaction. Combust. Flame 1968, 12, 477–491. [Google Scholar] [CrossRef]
  152. Lovachev, L.A. The rate constant of reaction H+O2 → OH+O. Combust. Flame 1963, 7, 388–389. [Google Scholar] [CrossRef]
  153. Miller, J.A. Nonstatistical effects and detailed balance in quasiclassical trajectory calculations of the thermal rate coefficient for O+OH→O2+H. J. Chem. Phys. 1986, 84, 6170–6177. [Google Scholar] [CrossRef]
  154. Myerson, A.L.; Watt, W.S. Atom-Formation Rates behind Shock Waves in Hydrogen and the Effect of Added Oxygen. J. Chem. Phys. 1968, 49, 425–433. [Google Scholar] [CrossRef]
  155. Namoradze, M.A.; Troshin, A.F.; Azatyan, V.V.; Dzotsenidze, Z.G.; Museridze, M.D. Effect of Initial Conditions on the Intensity of Chemiluminescence during Carbon Monoxide Oxidation. Soobshch. Akad. Nauk Gruz. SSR 1974, 73, 377. [Google Scholar]
  156. Pamidimukkala, K.M.; Skinner, G.B. Resonance Absorption Measurements of Atom Concentrations in Reacting Gas Mixtures. 9. Measurements of O atoms in Oxidation of H/sub 2/ and D/sub 2; Wright State Univ., Dept. of Chemistry: Dayton, OH, USA, 1981. [Google Scholar]
  157. Ripley, D.L.; Gardiner, W.C. Shock-Tube Study of the Hydrogen—Oxygen Reaction. II. Role of Exchange Initiation. J. Chem. Phys. 1966, 44, 2285–2296. [Google Scholar] [CrossRef]
  158. Schott, G.L. Further studies of exponential branching rates in reflected-shock heated, nonstoichiometric H2COO2 systems. Combust. Flame 1973, 21, 357–370. [Google Scholar] [CrossRef]
  159. Schott, G.L.; Kinsey, J.L. Kinetic Studies of Hydroxyl Radicals in Shock Waves. II. Induction Times in the Hydrogen-Oxygen Reaction. J. Chem. Phys. 1958, 29, 1177–1182. [Google Scholar] [CrossRef]
  160. Shin, K.S.; Michael, J.V. Rate constants for the reactions H+O2→OH+O and D+O2→OD+O over the temperature range 1085–2278 K by the laser photolysis–shock tube technique. J. Chem. Phys. 1991, 95, 262–273. [Google Scholar] [CrossRef]
  161. Shub, F.S.; Apel’baum, L.P.; Temkin, M.I. The accelerating effect of oxygen in a homogeneous reaction of water gas. Kinet. Catal. 1971, 11, 931. [Google Scholar]
  162. Yuan, T.; Wang, C.; Yu, C.; Frenklach, M.; Rabinowitz, M. Determination of the rate coefficient for the reaction H + O2 → OH + O by a shock tube/laser absorption/detailed modeling study. J. Phys. Chem. 1991, 95, 1258–1265. [Google Scholar] [CrossRef]
  163. Baldwin, R.R.; Cowe, D.W. The inhibition of the hydrogen + oxygen reaction by formaldehyde. Trans. Faraday Soc. 1962, 58, 1768–1781. [Google Scholar] [CrossRef]
  164. Browne, W.G.; Porter, R.P.; Verlin, J.D.; Clark, A.H. A study of acetylene-oxygen flames. Symp. (Int.) Combust. 1969, 12, 1035–1047. [Google Scholar] [CrossRef]
  165. Dixon-Lewis, G.; Linnett, J.W. Flame structure and flame reaction kinetics—V. Investigation of reaction mechanism in a rich hydrogen+nitrogen+oxygen flame by solution of conservation equations. Proc. R. Soc. Lond. A. Math. Phys. Sci. 1970, 317, 235–263. [Google Scholar] [CrossRef]
  166. Germann, T.C.; Miller, W.H. Quantum Mechanical Pressure-Dependent Reaction and Recombination Rates for O + OH → H + O2, HO2. J. Phys. Chem. A 1997, 101, 6358–6367. [Google Scholar] [CrossRef]
  167. Miller, J.A. Collision dynamics and the thermal rate coefficient for the reaction H+O2→OH+O. J. Chem. Phys. 1981, 74, 5120–5132. [Google Scholar] [CrossRef]
  168. Miller, J.A.; Garrett, B.C. Quantifying the non-RRKM effect in the H + O2 ⇄ OH + O reaction. Int. J. Chem. Kinet. 1997, 29, 275–287. [Google Scholar] [CrossRef]
  169. Westenberg, A.A.; Fristrom, R.M. H and O atom profiles measured by ESR in C2 hydrocarbon-O2 flames. Symp. (Int.) Combust. 1965, 10, 473–487. [Google Scholar] [CrossRef]
  170. Dixon-Lewis, G.; Wilson, W.E.; Westenberg, A.A. Studies of Hydroxyl Radical Kinetics by Quantitative ESR. J. Chem. Phys. 1966, 44, 2877–2884. [Google Scholar] [CrossRef]
  171. Dixon-Lewis, G.; Linnett, J.W. Flame structure and flame reaction kinetics. VII. Reactions of traces of heavy water, deuterium and carbon dioxide added to rich hydrogen + nitrogen + oxygen flames. Proc. R. Soc. of Lond. A. Math. Phys. Sci. 1972, 330, 219–245. [Google Scholar] [CrossRef]
  172. Dixon-Lewis, G.; Sutton, M.M.; Williams, A. Reactions contributing to the establishment of the water gas equilibrium when carbon dioxide is added to a hydrogen-oxygen flame. Trans. Faraday Soc. 1965, 61, 255–262. [Google Scholar] [CrossRef]
  173. Jost, W.; Schecker, H.G.; Wagner, H.G. Messungen der Geschwindigkeit der Einstellung des Wassergas-Gleichgewichtes. Z. Phys. Chem. 1965, 45, 47–57. [Google Scholar] [CrossRef]
  174. Herron, J.T. Mass-Spectrometric Study of the Rate of the Reaction CO + OH. J. Chem. Phys. 1966, 45, 1854–1855. [Google Scholar] [CrossRef]
  175. Heath, G.A.; Pearson, G.S. Perchloric acid flames: Part III. Chemical structure of methane flames. Symp. (Int.) Combust. 1967, 11, 967–977. [Google Scholar] [CrossRef]
  176. Porter, R.P.; Clark, A.H.; Kaskan, W.E.; Browne, W.E. A study of hydrocarbon flames. Symp. (Int.) Combust. 1967, 11, 907–917. [Google Scholar] [CrossRef]
  177. Wilson, W.E.; O’Donovan, J.T. Mass-Spectrometric Study of the Reaction Rate of OH with Itself and with CO. J. Chem. Phys. 1967, 47, 5455–5457. [Google Scholar] [CrossRef]
  178. Greiner, N.R. Hydroxyl Radical Kinetics by Kinetic Spectroscopy. V. Reactions with H2 and CO in the Range 300–500 °K. J. Chem. Phys. 1969, 51, 5049–5051. [Google Scholar] [CrossRef]
  179. Kochubei, V.F.; Moin, F.B. Kinetics of the reaction of CO2 with hydrogen. Kinet. Catal. 1969, 10, 992. [Google Scholar]
  180. Baldwin, R.R.; Walker, R.W.; Webster, S.J. The carbon monoxide-sensitized decomposition of hydrogen peroxide. Combust. Flame 1970, 15, 167–172. [Google Scholar] [CrossRef]
  181. Dean, A.M.; Kistiakowsky, G.B. Oxidation of Carbon Monoxide by Oxygen in Shock Waves. J. Chem. Phys. 1970, 53, 830–838. [Google Scholar] [CrossRef]
  182. Shub, F.S.; Apel’baum, L.O.; Temkin, M.I. The kinetics of the homogeneous water gas reaction. Kinet. Catal. 1970, 11, 462. [Google Scholar]
  183. Izod, T.P.J.; Kistiakowsky, G.B.; Matsuda, S. Oxidation of Carbon Monoxide Mixtures with Added Ethane or Azomethane Studied in Incident Shock Waves. J. Chem. Phys. 1971, 55, 4425–4432. [Google Scholar] [CrossRef]
  184. Stuhl, F.; Niki, H. Pulsed Vacuum-uv Photochemical Study of Reactions of OH with H2, D2, and CO Using a Resonance-Fluorescent Detection Method. J. Chem. Phys. 1972, 57, 3671–3677. [Google Scholar] [CrossRef]
  185. Eberius, K.H.; Hoyermann, K.; Wagner, H.C. Structure of lean acetylene-oxygen flames. Symp. (Int.) Combust. 1973, 14, 147–156. [Google Scholar] [CrossRef]
  186. Gardiner, W.C.; Mallard, W.G.; McFarland, M.; Morinaga, K.; Owen, J.H.; Rawlins, W.T.; Takeyama, T.; Walker, B.F. Elementary reaction rates from post-induction-period profiles in shock-initiated combustion. Symp. (Int.) Combust. 1973, 14, 61–75. [Google Scholar] [CrossRef]
  187. Smith, I.W.M.; Zellner, R. Rate measurements of reactions of OH by resonance absorption. Part 2.—Reactions of OH with CO, C2H4 and C2H2. J. Chem. Soc. Faraday Trans. 2 Mol. Chem. Phys. 1973, 69, 1617–1627. [Google Scholar] [CrossRef]
  188. Westenberg, A.A.; de Haas, N. Rates of CO + OH and H2 + OH over an extended temperature range. J. Chem. Phys. 1973, 58, 4061–4065. [Google Scholar] [CrossRef]
  189. Howard, C.J.; Evenson, K.M. Laser magnetic resonance study of the gas phase reactions of OH with CO, NO, and NO2. J. Chem. Phys. 1974, 61, 1943–1952. [Google Scholar] [CrossRef]
  190. Trainor, D.W.; Vonrosenberg, C.W. Energy partitioning in the reaction OH + CO → CO2 + H. Chem. Phys. Lett. 1974, 29, 35–38. [Google Scholar] [CrossRef]
  191. Biordi, J.C.; Lazzara, C.P.; Papp, J.F. Flame structure studies of CF3Br-Inhibited methane flames: II. Kinetics and mechanisms. Symp. (Int.) Combust. 1975, 15, 917–932. [Google Scholar] [CrossRef]
  192. Trainor, D.W.; von Rosenberg, C.W. Energy partitioning in the products of elementary reactions involving OH-radicals. Symp. (Int.) Combust. 1975, 15, 755–764. [Google Scholar] [CrossRef]
  193. Vandooren, J.; Peeters, J.; Tiggelen, P.J.V. Rate constant of the elementary reaction of carbon monoxide with hydroxyl radical. Symp. (Int.) Combust. 1975, 15, 745–753. [Google Scholar] [CrossRef]
  194. Atkinson, R.; Perry, R.A.; Pitts, J.N. Kinetics of the reactions of OH radicals with CO and N2O. Chem. Phys. Lett. 1976, 44, 204–208. [Google Scholar] [CrossRef]
  195. Atri, G.M.; Baldwin, R.R.; Jackson, D.; Walker, R.W. The reaction of OH radicals and HO2 radicals with carbon monoxide. Combust. Flame 1977, 30, 1–12. [Google Scholar] [CrossRef]
  196. Chan, W.H.; Uselman, W.M.; Calvert, J.G.; Shaw, J.H. The pressure dependence of the rate constant for the reaction: HO + CO → H + CO2. Chem. Phys. Lett. 1977, 45, 240–244. [Google Scholar] [CrossRef]
  197. Overend, R.; Paraskevopoulos, G. The question of a pressure effect in the reaction OH + CO at room temperature. Chem. Phys. Lett. 1977, 49, 109–111. [Google Scholar] [CrossRef]
  198. Biermann, H.W.; Zetzsch, C.; Stuhl, F. On the Pressure Dependence of the Reaction of HO with CO. Ber. Bunsenges. Phys. Chem. 1978, 82, 633–639. [Google Scholar] [CrossRef]
  199. Butler, R.; Solomon, I.J.; Snelson, A. Pressure dependence of the CO + OH rate constant in O2 + N2 mixtures. Chem. Phys. Lett. 1978, 54, 19–24. [Google Scholar] [CrossRef]
  200. Clyne, M.A.A.; Holt, P.M. Reaction Kinetics Involving Ground X2π and Excited A2Σ+ Hydroxyl Radicals. Part 1.—Quenching Kinetics of OH A2Σ+ and Rate Constants for Reactions of OH X2π with CH3CCl3 and CO. J. Chem. Soc. Faraday Trans. 2 1979, 75, 569. [Google Scholar] [CrossRef]
  201. Ravishankara, A.R.; Thompson, R.L. Kinetic study of the reaction of OH with CO from 250 to 1040 K. Chem. Phys. Lett. 1983, 99, 377–381. [Google Scholar] [CrossRef]
  202. Demore, W.B. Rate constant for the OH + CO reaction: Pressure dependence and the effect of oxygen. Int. J. Chem. Kinet. 1984, 16, 1187–1200. [Google Scholar] [CrossRef]
  203. Hofzumahaus, A.; Stuhl, F. Rate Constant of the Reaction HO + CO in the Presence of N2 and O2. Ber. Bunsenges. Phys. Chem. 1984, 88, 557–561. [Google Scholar] [CrossRef]
  204. Niki, H.; Maker, P.D.; Savage, C.M.; Breitenbach, L.P. Fourier transform infrared spectroscopic study of the kinetics for the hydroxyl radical reaction of 13C16O-carbon monoxide and 12C18O-carbon monoxide. J. Phys. Chem. 1984, 88, 2116–2119. [Google Scholar] [CrossRef]
  205. Smith, I.W.M.; Williams, M.D. Kinetics of OH(v = 0 and 1) and OD(v = 0 and 1) studied by time-resolved laser-induced fluorescence. Ber. Bunsenges. Phys. Chem. 1985, 89, 319. [Google Scholar] [CrossRef]
  206. Larson, C.W.; Stewart, P.H.; Golden, D.M. Pressure and temperature dependence of reactions proceeding via a bound complex. An approach for combustion and atmospheric chemistry modelers. Application to HO + CO → [HOCO] → H + CO2. Int. J. Chem. Kinet. 1988, 20, 27–40. [Google Scholar] [CrossRef]
  207. Frost, M.J.; Sharkey, P.; Smith, I.W.M. Energy and structure of the transition states in the reaction OH + CO → H + CO2. Faraday Discuss. Chem. Soc. 1991, 91, 305–317. [Google Scholar] [CrossRef]
  208. Frost, M.J.; Sharkey, P.; Smith, I.W.M. Reaction between hydroxyl (deuteroxyl) radicals and carbon monoxide at temperatures down to 80 K: Experiment and theory. J. Phys. Chem. 1993, 97, 12254–12259. [Google Scholar] [CrossRef]
  209. Wooldridge, M.S.; Hanson, R.K.; Bowman, C.T. A shock tube study of the CO+OH→CO2+H reaction. Symp. (Int.) Combust. 1994, 25, 741–748. [Google Scholar] [CrossRef]
  210. Forster, R.; Frost, M.; Fulle, D.; Hamann, H.F.; Hippler, H.; Schlepegrell, A.; Troe, J. High pressure range of the addition of HO to HO, NO, NO2, and CO. I. Saturated laser induced fluorescence measurements at 298 K. J. Chem. Phys. 1995, 103, 2949–2958. [Google Scholar] [CrossRef]
  211. Lissianski, V.; Yang, H.; Qin, Z.; Mueller, M.R.; Shin, K.S.; Gardiner, W.C. High-temperature measurements of the rate coefficient of the H + CO2 → CO + OH reaction. Chem. Phys. Lett. 1995, 240, 57–62. [Google Scholar] [CrossRef]
  212. Wooldridge, M.S.; Hanson, R.K.; Bowman, C.T. A shock tube study of CO + OH → CO2 + H and HNCO + OH → products via simultaneous laser absorption measurements of OH and CO2. Int. J. Chem. Kinet. 1996, 28, 361–372. [Google Scholar] [CrossRef]
  213. Bohn, B.; Zetzsch, C. Formation of HO2 from OH and C2H2 in the presence of O2. J. Chem. Soc. Faraday Trans. 1998, 94, 1203–1210. [Google Scholar] [CrossRef]
  214. Chen, W.-C.; Marcus, R.A. On the theory of the reaction rate of vibrationally excited CO molecules with OH radicals. J. Chem. Phys. 2006, 124, 024306. [Google Scholar] [CrossRef] [Green Version]
  215. Li, J.; Zhao, Z.; Kazakov, A.; Chaos, M.; Dryer, F.L.; Scire, J.J., Jr. A comprehensive kinetic mechanism for CO, CH2O, and CH3OH combustion. Int. J. Chem. Kinet. 2007, 39, 109–136. [Google Scholar] [CrossRef]
  216. Browne, W.G.; White, D.R.; Smookler, G.R. A study of the chemical kinetics of shock heated H2/CO/O2 mixtures. Symp. (Int.) Combust. 1969, 12, 557–567. [Google Scholar] [CrossRef]
  217. Day, M.J.; Thompson, K.; Dixon-Lewis, G. Some reactions of hydroperoxyl and hydroxylradicals at high temperatures. Symp. (Int.) Combust. 1973, 14, 47–59. [Google Scholar] [CrossRef]
  218. Mozurkewich, M.; Lamb, J.J.; Benson, S.W. Negative activation energies and curved Arrhenius plots. 2. Hydroxyl + carbon monoxide. J. Phys. Chem. 1984, 88, 6435–6441. [Google Scholar] [CrossRef]
  219. Dzegilenko, F.N.; Bowman, J.M. Recovering a full dimensional quantum rate constant from a reduced dimensionality calculation: Application to the OH+CO→H+CO2 reaction. J. Chem. Phys. 1996, 105, 2280–2286. [Google Scholar] [CrossRef]
  220. Zsély, I.G.; Nagy, T.; Simmie, J.M.; Curran, H.J. Reduction of a detailed kinetic model for the ignition of methane/propane mixtures at gas turbine conditions using simulation error minimization methods. Combust. Flame 2011, 158, 1469–1479. [Google Scholar] [CrossRef]
  221. Varga, T.; Olm, C.; Nagy, T.; Zsely, I.G.; Valko, E.; Palvolgyi, R.; Curran, H.J.; Turanyi, T. Development of a Joint Hydrogen and Syngas Combustion Mechanism Based on an Optimization Approach. Int. J. Chem. Kinet. 2016, 48, 407–422. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Simulated pressure curve and ignition delay time (IDT)-definition for rapid compression machine (RCM) (left) and shock tube (ST) (right).
Figure 1. Simulated pressure curve and ignition delay time (IDT)-definition for rapid compression machine (RCM) (left) and shock tube (ST) (right).
Energies 13 00778 g001
Figure 2. Schematic representation of the mechanism development process presented in the actual study.
Figure 2. Schematic representation of the mechanism development process presented in the actual study.
Energies 13 00778 g002
Figure 3. Experimental data (symbols) [12,13] versus model prediction (lines) using the mechanism MechA (solid lines) and MechB (dashed lines). The respective content of methane, propane, and n-heptane is given in mol% and corresponds to the values in Table 1.
Figure 3. Experimental data (symbols) [12,13] versus model prediction (lines) using the mechanism MechA (solid lines) and MechB (dashed lines). The respective content of methane, propane, and n-heptane is given in mol% and corresponds to the values in Table 1.
Energies 13 00778 g003
Figure 4. Comparison of the reaction rate coefficient of reaction 42 (CH4 + OH ↔ H2O + CH3) of MechA to MechC as well as the fitted rate coefficient using reference data from the NIST (National Institute of Standards and Technology) database.
Figure 4. Comparison of the reaction rate coefficient of reaction 42 (CH4 + OH ↔ H2O + CH3) of MechA to MechC as well as the fitted rate coefficient using reference data from the NIST (National Institute of Standards and Technology) database.
Energies 13 00778 g004
Figure 5. Comparison of the reaction rate coefficient of reaction 44 (CH4 + O2 ↔ CH3 + HO2) of MechA to MechC as well as the fitted rate coefficient using reference data from the NIST database.
Figure 5. Comparison of the reaction rate coefficient of reaction 44 (CH4 + O2 ↔ CH3 + HO2) of MechA to MechC as well as the fitted rate coefficient using reference data from the NIST database.
Energies 13 00778 g005
Figure 6. Comparison of the reaction rate coefficient of reaction 45 (CH4 + HO2 ↔ CH3 + H2O2) of MechA to MechC as well as the fitted rate coefficient using reference data from the NIST database.
Figure 6. Comparison of the reaction rate coefficient of reaction 45 (CH4 + HO2 ↔ CH3 + H2O2) of MechA to MechC as well as the fitted rate coefficient using reference data from the NIST database.
Energies 13 00778 g006
Figure 7. Simulated time-dependent sensitivity towards temperature of the reaction 274 (C7H16 + OH ↔ H2O + n-C7H15) and the pressure profile in the area of ignition calculated at 800 K and 60 bar, using mix 9 from Table 1.
Figure 7. Simulated time-dependent sensitivity towards temperature of the reaction 274 (C7H16 + OH ↔ H2O + n-C7H15) and the pressure profile in the area of ignition calculated at 800 K and 60 bar, using mix 9 from Table 1.
Energies 13 00778 g007
Figure 8. Reaction sensitivity towards temperature simulated with mix 9 from Table 1.
Figure 8. Reaction sensitivity towards temperature simulated with mix 9 from Table 1.
Energies 13 00778 g008
Figure 9. Comparison of the reaction rate coefficient of reaction 302 (n-C7-OQOOH ↔ OH + CH2O + CO + C2H4 + n-C3H7) of MechA to MechC.
Figure 9. Comparison of the reaction rate coefficient of reaction 302 (n-C7-OQOOH ↔ OH + CH2O + CO + C2H4 + n-C3H7) of MechA to MechC.
Energies 13 00778 g009
Figure 10. Sensitivity towards temperature simulated with MechA and mix 4 from Table 1 (left); measured IDTs (symbols) and simulated values (lines) using MechA for different methane–propane mixtures (right). The content of methane and propane is given in mol% and corresponds to mix 1 to 4 in Table 1.
Figure 10. Sensitivity towards temperature simulated with MechA and mix 4 from Table 1 (left); measured IDTs (symbols) and simulated values (lines) using MechA for different methane–propane mixtures (right). The content of methane and propane is given in mol% and corresponds to mix 1 to 4 in Table 1.
Energies 13 00778 g010
Figure 11. Extract from the propane oxidation scheme derived from MechA. Reactions with modified Arrhenius parameters to improve the calculation of the propane influence on the IDT are labeled in red.
Figure 11. Extract from the propane oxidation scheme derived from MechA. Reactions with modified Arrhenius parameters to improve the calculation of the propane influence on the IDT are labeled in red.
Energies 13 00778 g011
Figure 12. Comparison of the reaction rate coefficient of reaction 235 (C3H8 + OH ↔ n-C3H7 + H2O) of MechA and MechD as well as the fitted rate coefficient using reference data from the NIST database.
Figure 12. Comparison of the reaction rate coefficient of reaction 235 (C3H8 + OH ↔ n-C3H7 + H2O) of MechA and MechD as well as the fitted rate coefficient using reference data from the NIST database.
Energies 13 00778 g012
Figure 13. Comparison of the reaction rate coefficient of reaction 238 (C3H8 + HO2n-C3H7 + H2O2) of MechA, MechD and AramcoMech 3.0 as well as the fitted rate coefficient using reference data from the NIST database.
Figure 13. Comparison of the reaction rate coefficient of reaction 238 (C3H8 + HO2n-C3H7 + H2O2) of MechA, MechD and AramcoMech 3.0 as well as the fitted rate coefficient using reference data from the NIST database.
Energies 13 00778 g013
Figure 14. Comparison of the reaction rate coefficient of reaction 244 (n-C3H7 + O2 ↔ C3H6 + HO2) of MechA and MechD as well as the fitted rate coefficient using reference data from the NIST database.
Figure 14. Comparison of the reaction rate coefficient of reaction 244 (n-C3H7 + O2 ↔ C3H6 + HO2) of MechA and MechD as well as the fitted rate coefficient using reference data from the NIST database.
Energies 13 00778 g014
Figure 15. Comparison of the reaction rate coefficient of reaction 247 (C3H6OOH + O2 ↔ OC3H5OOH + OH) of MechA and MechD.
Figure 15. Comparison of the reaction rate coefficient of reaction 247 (C3H6OOH + O2 ↔ OC3H5OOH + OH) of MechA and MechD.
Energies 13 00778 g015
Figure 16. Comparison of the reaction rate coefficient of reaction 248 (OC3H5OOH ↔ CH2CHO + CH2O + OH) of MechA and MechD.
Figure 16. Comparison of the reaction rate coefficient of reaction 248 (OC3H5OOH ↔ CH2CHO + CH2O + OH) of MechA and MechD.
Energies 13 00778 g016
Figure 17. Measured laminar flame speeds (symbols) of methane–air mixtures [112] as function of the fuel–air equivalence ratio ϕ at different pressures and temperatures in comparison with simulated values (lines) using MechD.
Figure 17. Measured laminar flame speeds (symbols) of methane–air mixtures [112] as function of the fuel–air equivalence ratio ϕ at different pressures and temperatures in comparison with simulated values (lines) using MechD.
Energies 13 00778 g017
Figure 18. Analysis of the most sensitive reactions regarding the laminar flame speed as function of the fuel–air equivalence ratio ϕ using a methane–air mixture at 300 K and 10.133 (left) and 20.265 bar (right).
Figure 18. Analysis of the most sensitive reactions regarding the laminar flame speed as function of the fuel–air equivalence ratio ϕ using a methane–air mixture at 300 K and 10.133 (left) and 20.265 bar (right).
Energies 13 00778 g018
Figure 19. Analysis of the most sensitive reactions regarding the laminar flame speed as function of the fuel–air equivalence ratio ϕ using a methane–air mixture at 300 (left), 360 (middle), and 400 K (right) and 10 bar.
Figure 19. Analysis of the most sensitive reactions regarding the laminar flame speed as function of the fuel–air equivalence ratio ϕ using a methane–air mixture at 300 (left), 360 (middle), and 400 K (right) and 10 bar.
Energies 13 00778 g019
Figure 20. Comparison of the reaction rate coefficient of reaction 1 (H + O2 ↔ OH + O) of MechA and TU Wien dual fuel mechanism 2.0 as well as the fitted rate coefficient using reference data from the NIST database.
Figure 20. Comparison of the reaction rate coefficient of reaction 1 (H + O2 ↔ OH + O) of MechA and TU Wien dual fuel mechanism 2.0 as well as the fitted rate coefficient using reference data from the NIST database.
Energies 13 00778 g020
Figure 21. Comparison of the reaction rate coefficient of reaction 25 (CO + OH ↔ CO2 + H) of MechA, TU Wien dual fuel mechanism 2.0, Zsely et al. [220], Varga et al. [221] as well as the fitted rate coefficient using reference data from the NIST database.
Figure 21. Comparison of the reaction rate coefficient of reaction 25 (CO + OH ↔ CO2 + H) of MechA, TU Wien dual fuel mechanism 2.0, Zsely et al. [220], Varga et al. [221] as well as the fitted rate coefficient using reference data from the NIST database.
Energies 13 00778 g021
Figure 22. Experimental data (symbols) [12,13] versus model prediction (lines) using the mechanism MechB (dashed lines) and the final TU Wien dual fuel mechanism 2.0 (solid lines). The diagram in the upper left corner highlights the comparison between measured and calculated IDTs of methane–propane mixtures. The respective content of methane, propane, and n-heptane is given in mol% and corresponds to the values in Table 1.
Figure 22. Experimental data (symbols) [12,13] versus model prediction (lines) using the mechanism MechB (dashed lines) and the final TU Wien dual fuel mechanism 2.0 (solid lines). The diagram in the upper left corner highlights the comparison between measured and calculated IDTs of methane–propane mixtures. The respective content of methane, propane, and n-heptane is given in mol% and corresponds to the values in Table 1.
Energies 13 00778 g022
Figure 23. Measured laminar flame speeds of methane–air mixtures [112] as function of the fuel–air equivalence ratio ϕ in comparison with simulated values before (MechD; dashed lines) and after the Arrhenius parameter adaption of reactions 1, 25, and 56 (TU Wien dual fuel mechanism 2.0; solid lines).
Figure 23. Measured laminar flame speeds of methane–air mixtures [112] as function of the fuel–air equivalence ratio ϕ in comparison with simulated values before (MechD; dashed lines) and after the Arrhenius parameter adaption of reactions 1, 25, and 56 (TU Wien dual fuel mechanism 2.0; solid lines).
Energies 13 00778 g023
Figure 24. IDT simulation with the TU Wien dual fuel mechanism 2.0 (lines) and the TU Wien dual fuel mechanism 2.0 without the nitrogen sub-mechanism (symbols). The respective content of methane, propane and n-heptane is given in mol% and corresponds to the values in Table 1.
Figure 24. IDT simulation with the TU Wien dual fuel mechanism 2.0 (lines) and the TU Wien dual fuel mechanism 2.0 without the nitrogen sub-mechanism (symbols). The respective content of methane, propane and n-heptane is given in mol% and corresponds to the values in Table 1.
Energies 13 00778 g024
Figure 25. Simulated laminar flame speed values as function of the fuel–air equivalence ratio ϕ using the TU Wien dual fuel mechanism 2.0 (lines) and the TU Wien dual fuel mechanism 2.0 without the nitrogen sub-mechanism (symbols).
Figure 25. Simulated laminar flame speed values as function of the fuel–air equivalence ratio ϕ using the TU Wien dual fuel mechanism 2.0 (lines) and the TU Wien dual fuel mechanism 2.0 without the nitrogen sub-mechanism (symbols).
Energies 13 00778 g025
Table 1. Overview of the homogeneous mixtures used for the optimization of the reaction mechanism including the respective proportion of methane, propane and n-heptane in the mixture, each in mol%, the fuel–air equivalence ratio ϕ, the pressure p in bar, and the temperature T in K.
Table 1. Overview of the homogeneous mixtures used for the optimization of the reaction mechanism including the respective proportion of methane, propane and n-heptane in the mixture, each in mol%, the fuel–air equivalence ratio ϕ, the pressure p in bar, and the temperature T in K.
Test FacilityFuel CompositionpTRef.
MixCH4 [mol%]C3H8 [mol%]C7H16 [mol%]ϕ [−][bar][K]
Rapid compression machine1100000.526100906–941[13]
295500.526100888–916[13]
3901000.526100803–898[12]
4703000.526100826–865[13]
592.684.882.440.59460701–877[12]
690.484.764.760.66160671–781[12]
792.684.882.440.594100709–817[12]
897.5602.440.59960720–869[12]
Shock tube986.364.559.090.79660748–1187[12]
1090.9109.090.81660785–1284[12]
Table 2. Overview of methane reactions used for the adjustment in MechB.
Table 2. Overview of methane reactions used for the adjustment in MechB.
Reaction Number Reaction Equation
42CH4 + OH ↔ H2O + CH3
44CH4 + O2 ↔ CH3 + HO2
45CH4 + HO2 ↔ CH3 + H2O2
Table 3. Overview of the reactions used to adjust the calculation of the propane influence on the IDT.
Table 3. Overview of the reactions used to adjust the calculation of the propane influence on the IDT.
Reaction NumberReaction Equation
235C3H8 + OH ↔ n-C3H7 + H2O
238C3H8 + HO2n-C3H7 + H2O2
244n-C3H7 + O2 ↔ C3H6 + HO2
247C3H6OOH + O2 ↔ OC3H5OOH + OH
248OC3H5OOH ↔ CH2CHO + CH2O + OH
Table 4. Reactions selected for adjusting the laminar flame speed calculation.
Table 4. Reactions selected for adjusting the laminar flame speed calculation.
Reaction NumberReaction Equation
1H + O2 ↔ OH + O
25CO + OH ↔ CO2 + H
56CH3 + H (+M) ↔ CH4 (+M)
Table 5. Overview of the reactions used for the mechanism adaption, the purpose of the adaptation and the resulting mechanism.
Table 5. Overview of the reactions used for the mechanism adaption, the purpose of the adaptation and the resulting mechanism.
StepReact.Reaction EquationPurposeMech.
142CH4 + OH ↔ H2O + CH3Update of adjustments made in [7]MechC
44CH4 + O2 ↔ CH3 + HO2
45CH4 + HO2 ↔ CH3 + H2O2
302n-C7-OQOOH ↔ OH + CH2O + CO + C2H4 + n-C3H7
2235C3H8 + OH ↔ n-C3H7 + H2OCorrecting propane influence on IDTMechD
238C3H8 + HO2n-C3H7 + H2O2
244n-C3H7 + O2 ↔ C3H6 + HO2
247C3H6OOH + O2 ↔ OC3H5OOH + OH
248OC3H5OOH ↔ CH2CHO + CH2O + OH
31H + O2 ↔ OH + OCorrecting flame speed calculationTU Wien dual fuel mech. 2.0
25CO + OH ↔ CO2 + H
56CH3 + H (+M) ↔ CH4 (+M)

Share and Cite

MDPI and ACS Style

Schuh, S.; Winter, F. Dual Fuel Reaction Mechanism 2.0 including NOx Formation and Laminar Flame Speed Calculations Using Methane/Propane/n-Heptane Fuel Blends. Energies 2020, 13, 778. https://doi.org/10.3390/en13040778

AMA Style

Schuh S, Winter F. Dual Fuel Reaction Mechanism 2.0 including NOx Formation and Laminar Flame Speed Calculations Using Methane/Propane/n-Heptane Fuel Blends. Energies. 2020; 13(4):778. https://doi.org/10.3390/en13040778

Chicago/Turabian Style

Schuh, Sebastian, and Franz Winter. 2020. "Dual Fuel Reaction Mechanism 2.0 including NOx Formation and Laminar Flame Speed Calculations Using Methane/Propane/n-Heptane Fuel Blends" Energies 13, no. 4: 778. https://doi.org/10.3390/en13040778

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop