• Keine Ergebnisse gefunden

OOQOOH

iC 8 H 18 /air φ =1

6.1 Reduced Mechanism

6.0.3 RedMaster-S

Similar to RedMaster-R the program automatically performs ignition-delay calculations with the CHEMKIN package for selected experimental data. It reads the output CHEMKIN files and selects automatically the time points and produces input files for a CONNECT analysis in the KINALC package as described in subsection 6.0.1. The user defines an initial group of important species (i.e. fuel components). The program finds the species most “connected”

to this group and adds them to the bottom of the initial group of species (list of species). It repeats the iterations until all species are in the list. The final list contains species according to their importance (their connection to the initial group) - most important species on the top, least important species in the bottom part of the list. Then the program produces similar lists for other timepoints and experiments. Then it compares bottom portion of the lists and chooses unimportant species - the species common simultaneously for all collected lists. The bottom portion of the list, which is used in that comparison is a user defined parameter. Then the program removes unimportant species and their reactions from the mechanism. The flow chart of the program is shown in Fig. 6.3.

Similar to RedMaster-R, as input the user provides the kinetic mechanism file, the ther-modynamical data file, the file with the list of experiments, and program control file. The file with the list of experiments contains all data necessary for ignition delay calculation: temper-ature, pressure, and mixture composition. The program control file contains the number of experiments which should be used for the further analysis, calculation data for the Chemkin II package (relative and absolute error, calculation time) and parameters for the unimportant reactions analysis (number of time points at which the analysis is made before and after the autoignition; the initial group of species; theportion of the list of connected species which is taken as unimportant). As output the user gets the file with calculated ignition delays and a reduced mechanism.

Table 15: Reduced mechanisms obtained with RedMaster-S.

Variant Experiments Used Portion of the list Reduced Mechanism

Substances Reactions Pressure Temperature

1.1 91-92 0.5 148 1047

9 bar 1148 K, 1427 K

1.2 91-92 0.65 115 829

1.3 91-92 0.75 98 646

1.4 91-92 0.77 91 528

1.5 91-92 0.8 77 452

2.1 2-17 0.5 152 119

25 bar 880-1220 K

2.2 2-17 0.6 129 859

2.3 2-17 0.75 89 519

3.1 4-6 0.5 141 986

25 bar 952-963 K

3.2 4-6 0.7 111 725

3.3 4-6 0.75 96 543

3.4 4-6 0.8 73 407

4.1 91-92 0.65 80 670

9 bar 1148 K, 1427 K

4.2 91-92 0.7 79 577

4.3 91-92 0.75 69 468

calculations with mechanisms 1.4 and 1.5 have growing discrepancy from original mechanism.

The second set of reduced mechanisms is made for a wide range of initial conditions.

Experiments 2-17 cover temperatures from 880 to 1220 K, and pressures around 25 bar.

Figures 6.6 and 6.7 show the change in discrepancy vs. degree of reduction. A large number of used experiments does not provide high quality reduction at these points and deviations at high temperatures are observed.

The third set of reduced mechanisms is derived for temperatures, where transition from low to high temperature combustion occurs. This process requires substances and reactions for both paths of combustion process. Therefore only a moderate degree of reduction can be observed. Figure 6.8 shows that the reduced mechanism 3.3 has a low deviation from the experiments 4-6, but provides high errors in the high temperature region. Mechanism 3.4 has very high errors for all experimental points.

During the reduction of the mechanism one of the main problems is the erroneous elimi-nation of some important species. Each of 5 surrogate model components have very different combustion properties. In the developed model methylnaphthalene has a very low influence on the ignition delay of the mixture. This is also shown in the sensitivity analysis in subsec-tion 5.8. Thus some species necessary for the methylnaphthalene oxidasubsec-tion are automatically deleted, so that this component cannot oxidate. This usually happens when the number of species goes under 100 or 90 and it limits the reduction to moderate degrees.

In order to get a maximal reduction, first the low temperature reactions are deleted and then the RedMaster-S is used. The set of mechanisms 4.1-4.3 is obtained in such way.

Figs. 6.9-6.10 represent the deviation from the full mechanism. Mechanism 4.3 has only 69 species, but its deviation has not significantly increased in comparison to mechanisms 1.5 or 2.3. Mechanisms 4.2 and 4.3 are further reduced with RedMaster-R. The same experimental

Table 16: Reduced mechanisms obtained with RedMaster-R.

Variant Experiments Used TREAC Reduced Mechanism Substances Reactions

4.2.1 91-92 80 79 303

4.2.2 91-92 10 79 348

4.2.3 91-92 1 79 402

4.2.4 91-92 0.1 79 445

4.2.5 91-92 0.01 79 488

4.3.1 91-92 80 69 230

4.3.2 91-92 10 69 256

4.3.3 91-92 1 69 300

4.3.4 91-92 0.1 69 343

4.3.5 91-92 0.01 69 381

points 91-92 are used. Table 16 summarizes all variants of the RedMaster-R reduction, that are made for each mechanism.

Figures 6.11 and 6.12 show that even at high reduction degrees there is no significant increase in the discrepancy between the reduced 4.2.1-4.2.5 and the full mechanisms. The difference is mainly seen in the high temperature region. The same is true for the set of reduced mechanisms 4.3.1-4.3.5, see Figs. 6.13 and 6.14. Mechanism 4.3.1 is further tested.

Figs. 6.15-6.18 represent a comparison of the reduced mechanism with the full mechanism and experimental data. The discrepancy with the full mechanism in the high temperature region is generally less than a factor of 2, which is very good for an almost threefold reduction in the number of species and six-fold reduction of the number of reactions.

0,60 0,65 0,70 0,75 0,80 0,85 0,90 0,95 1,00

1 10 100 1000

P = 9 bar Jet-A = 1.257%

O2 = 20.745%, N2 = 77.998%

φ=1

calculation, full mechanism, this work calculation, reduced mechanism 1.1 calculation, reduced mechanism 1.2 calculation, reduced mechanism 1.3 calculation, reduced mechanism 1.4 calculation, reduced mechanism 1.5

Ignition Delay, µs

1000/T, 1/K

Figure 6.4: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 1.1-1.5 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approximately 9 bar.

0,80 0,85 0,90 0,95 1,00 1,05 1,10 100

1000 10000

P = 25 bar Jet-A = 1.276%

O2 = 20.74%

N2 = 77.98%

φ=1.0

calculation, full mechanism, this work calculation, reduced mechanism 1.1 calculation, reduced mechanism 1.2 calculation, reduced mechanism 1.3 calculation, reduced mechanism 1.4 calculation, reduced mechanism 1.5

Ignition Delay, µs

1000/T, 1/K

Figure 6.5: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 1.1-1.5 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approximately 25 bar.

0,80 0,85 0,90 0,95 1,00 1,05 1,10 1,15

100 1000 10000

calculation, full mechanism, this work calculation, reduced mechanism 2.1 calculation, reduced mechanism 2.2 calculation, reduced mechanism 2.3

Ignition Delay, µs

1000/T, 1/K

Jet-A = 1.276%

O2 = 20.74%

N2 = 77.98%

φ=1.0

P = 25 bar

Figure 6.6: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 2.1-2.3 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approximately 25 bar.

0,60 0,65 0,70 0,75 0,80 0,85 0,90 0,95 1,00 1

10 100 1000 10000

calculation, full mechanism, this work calculation, reduced mechanism 2.1 calculation, reduced mechanism 2.2 calculation, reduced mechanism 2.3

Ignition Delay, µs

1000/T, 1/K

Jet-A = 1.257%

O2 = 20.745%, N2 = 77.998%

φ=1

P = 9 bar

Figure 6.7: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 2.1-2.3 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approximately 9 bar.

0,80 0,85 0,90 0,95 1,00 1,05 1,10 1,15

100 1000 10000

calculation, full mechanism, this work calculation, reduced mechanism 3.1 calculation, reduced mechanism 3.2 calculation, reduced mechanism 3.3 calculation, reduced mechanism 3.4

Ignition Delay, µs

1000/T, 1/K

Jet-A = 1.276%

O2 = 20.74%

N2 = 77.98%

φ=1.0

P = 25 bar

Figure 6.8: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 3.1-3.4 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approximately 25 bar.

0,80 0,85 0,90 0,95 1,00 1,05 1,10 1,15 100

1000 10000

Jet-A = 1.276%

O2 = 20.74%

N2 = 77.98%

φ=1.0

calculation, full mechanism, this work calculation, reduced mechanism 4.1 calculation, reduced mechanism 4.2 calculation, reduced mechanism 4.3

Ignition Delay, µs

1000/T, 1/K

P = 25 bar

Figure 6.9: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 4.1-4.3 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approximately 25 bar.

0,70 0,75 0,80 0,85 0,90 0,95 1,00

10 100 1000 10000

calculation, full mechanism, this work calculation, reduced mechanism 4.1 calculation, reduced mechanism 4.2 calculation, reduced mechanism 4.3

Ignition Delay, µs

1000/T, 1/K

Jet-A = 1.257%

O2 = 20.745%, N2 = 77.998%

φ=1

P = 9 bar

Figure 6.10: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 4.1-4.3 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approximately 9 bar.

0,80 0,85 0,90 0,95 1,00 1,05 1,10 1,15 1,20 100

1000

calculation, full mechanism, this work calculation, reduced mechanism 4.2.1 calculation, reduced mechanism 4.2.2 calculation, reduced mechanism 4.2.3 calculation, reduced mechanism 4.2.4 calculation, reduced mechanism 4.3.5

Ignition Delay, µs

1000/T, 1/K

Jet-A = 1.276%

O2 = 20.74%

N2 = 77.98%

φ=1.0

P = 25 bar

Figure 6.11: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 4.2.1-4.2.5 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approxi-mately 25 bar.

0,60 0,65 0,70 0,75 0,80 0,85 0,90 0,95 1,00

1 10 100 1000

calculation, full mechanism, this work calculation, reduced mechanism 4.2.1 calculation, reduced mechanism 4.2.2 calculation, reduced mechanism 4.2.3 calculation, reduced mechanism 4.2.4 calculation, reduced mechanism 4.2.5

Ignition Delay, µs

1000/T, 1/K

Jet-A = 1.257%

O2 = 20.745%, N2 = 77.998%

φ=1

P = 9 bar

Figure 6.12: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 4.2.1-4.2.5 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approxi-mately 9 bar.

0,80 0,85 0,90 0,95 1,00 1,05 1,10 100

1000

Jet-A = 1.276%

O2 = 20.74%

N2 = 77.98%

φ=1.0

P = 25 bar

calculation, Full Mechanism calculation, Reduced Mechanism 4.3 calculation, Reduced Mechanism 4.3.1 calculation, Reduced Mechanism 4.3.2 calculation, Reduced Mechanism 4.3.3 calculation, Reduced Mechanism 4.3.4 calculation, Reduced Mechanism 4.3.5

Ignition Delay, µs

1000/T, 1/K

Figure 6.13: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 4.3.1-4.3.5 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approxi-mately 25 bar.

0,60 0,65 0,70 0,75 0,80 0,85 0,90 0,95 1,00

1 10 100 1000

P = 9 bar Jet-A = 1.257%

O2 = 20.745%, N2 = 77.998%

φ=1

Calculation, Full mechanism, this work Calculation, Reduced mechanism 4.3 Calculation, Reduced mechanism 4.3.1 Calculation, Reduced mechanism 4.3.2 Calculation, Reduced mechanism 4.3.3 Calculation, Reduced mechanism 4.3.4 Calculation, Reduced mechanism 4.3.5

Ignition Delay, µs

1000/T, 1/K

Figure 6.14: Comparison of ignition delay times calculated using the full and set of reduced mechanisms 4.3.1-4.3.5 for a Jet-A/air mixture. Stoichiometric mixture. Pressure approxi-mately 9 bar.

0,90 0,95 1,00

Jet-A = 0.615%

O2 = 10.0%, N2 = 89.38%

φ=1

experiment, 18 bar, Vasu et al., 2008 calculation, 18 bar, this work

calculation, 18 bar, reduced mechanism 4.3.1

Ignition Delay, µs

1000/T, 1/K

Figure 6.15: Comparison of experimental [118] ignition delay times and calculated ignition delay times for a Jet-A/O2/N2 mixture. Stoichiometric mixture. Pressure approximately 18 bar.

0,80 0,85 0,90 0,95 1,00

100 1000

Jet-A = 1.276%

O2 = 20.74%

N2 = 77.98%

φ=1.0

calculation, 25 bar, Full Mechanism

calculation, 25 bar, Reduced Mechanism 4.3.1 experiment, 25 bar, Vasu et al., 2008

Ignition Delay, µs

1000/T, 1/K

Figure 6.16: Comparison of experimental [118] ignition delay times and calculated ignition delay times for a Jet-A/air mixture. Stoichiometric mixture. Pressure approximately 25 bar.

0,85 0,90 0,95 1,00 10

100 1000 10000

Jet-A = 0.642%

O

2

= 20.87%, N

2

= 78.48%

φ

=0.5

experiment, 21 bar, Vasu et al., 2008 calculation, 21 bar, this work

calculation, 21 bar, reduced mechanism 4.3.1

Ignition Delay, µs

1000/T, 1/K

Figure 6.17: Comparison of experimental [118] ignition delay times and calculated ignition delay times for a Jet-A/air mixture. Lean mixture, φ=0.5. Pressure approximately 21 bar.

0,85 0,90 0,95 1,00

10 100 1000 10000

Jet-A = 2.48%

O2 = 20.49%, N2 = 77.03%

φ=2

experiment, 9 bar, Vasu et al., 2008 calculation, 9 bar, this work

calculation, 9 bar, reduced mechanism 4.3.1

Ignition Delay, µs

1000/T, 1/K

Figure 6.18: Comparison of experimental [118] ignition delay times and calculated ignition delay times for a Jet-A/air mixture. Rich mixture, φ=2. Pressure approximately 9 bar.

7 Conclusions and Future Work

The objective of this dissertation is to develop a methodology for modeling surrogate fuels and the development of a reduced kinetic mechanism for these fuels. Several programs are developed for the calculation of the physical properties (critical points, phase diagrams, dis-tillation curves), for calculation of thermodynamic data, and for the automatic reduction of the kinetic mechanisms.

As an example the methodology is applied to aviation kerosene Jet-A producing a model consisting of 5 hydrocarbons. The types of hydrocarbons and their percentages are chosen in such way that at the same time it met the physical properties and general chemical com-position requirements of Jet-A kerosene. The model obtained is also compared to a number of other Jet-A surrogates suggested in other works.

In order to compare the combustion properties of the surrogate with Jet-A kerosene 3 submechanisms (n-decane, n-dodecane, n-hexadecane) are developed and added to other kinetic models, developed at DLR. The resulting kinetic model of proposed surrogate is extensively tested over a wide range of initial conditions. The work focused on the conditions typical for jet engines: pressures from 10 to 40 bar, equivalence ratios between 0.5 and 2, and a wide range of temperatures. The proposed mechanism consists of 183 species and 1239 reversible reactions. Calculated ignition delays show very good agreement with experimental data and deviations are generally less than a factor of two. Another feature of the mechanism is that it includes the chemistry of aromatic compounds. This allows the model to be easily extended for the predictions of soot precursors.

The programs for automatic mechanism reduction are extensively tested. Several reduced models are obtained and compared with the initial model. It is established, that the reduction of the mechanisms produces small errors until some certain number of components, after which a further reduction causes very high deviations from the initial model. This minimal number of components depends on the numerical experiments, over which the reduction is made. For high temperatures a higher degree of reduction can be achieved within the same deviation from the original mechanism. This is due to the fact, that low-temperature chemistry becomes unimportant and a larger number of species can be deleted.

A higher degree of reduction of the number of species can be obtained in the high temper-ature region. This allowed the reduction of an already short initial model to only 69 species and 230 reactions. This reduced mechanism is also extensively tested by comparison with the initial mechanism. At temperatures above 1100 K it showed little deviation from the original mechanism.

The methodology developed here is independent on the type of the fuel. This could facilitate the process of internal combustion engine development for these types of fuels. For example it permits relatively simple modeling of other types of fuels, such as JP-8 kerosene, diesel, petrol, etc. Also, the hierarchical structure of hydrocarbons combustion mechanisms allows an easy extension of the combustion model for other types of hydrocarbons.

Future work should be conducted in two directions. First, the submechanisms of each model fuel component should be extended to improve the agreement with experimental re-sults. Second, the reduction methods should be improved to provide a higher degree of reduction. One of the possibilities could be the application of genetic algorithms for the op-timization of the parameters. An automatic selection of the reaction importance index and the control points where analysis is performed would almost completely remove the need for user intervention in the reduction process.

References

[1] Westbrook, C. K. and Dryer, F. L., “Simplified reaction mechanisms for the oxidation of hydrocarbon fuels in flames.,”Combustion Science and Technology 27(1-2), 31–43 (1981).

[2] Molnar, M. and John Marek, C., “Simplified two-time step method for calculating combustion and emission rates of Jet-A and methane fuel with and without water injection,”AIAA 2005-0549, 13107–13137 (2005).

[3] Slavinskaya, N., Zizin, A., and Aigner, M., “On model design of a surrogate fuel formu-lation,”Journal of Engineering for Gas Turbines and Power 132, 1–11 (2010).

[4] Colket, M., Edwards, T., Williams, S., Cernansky, N. P., Miller, D. L., Egolfopoulos, F., Lindstedt, P., Seshadri, K., Dryer, F. L., Law, C. K., Friend, D., Lenhert, D. B., Pitsch, H., Sarofim, A., Smooke, M., and Tsang., W., “Development of an experimental database and kinetic models for surrogate jet fuels,”AIAA 2007-770 (2007).

[5] Bacha, J., Barnes, F., Franklin, M., Gibbs, L., Hemighaus, G., Hogue, N., Lesnini, D., Lind, J., Maybury, J., and Morris, J., [Aviation fuels technical review (FTR-3)] (2000).

[6] Dagaut, P. and Cathonnet, M., “The ignition, oxidation, and combustion of kerosene:

A review of experimental and kinetic modeling,”Progress in Energy and Combustion Science 32(1), 48 – 92 (2006).

[7] Lindstedt, P. and Maurice, L., “Detailed chemical-kinetic model for aviation fuels,”

Journal of Propulsion and Power 16, 187–195 (2000).

[8] Mawid, M., Park, T., Sekar, B., and Arana, C., “Importance of surrogate JP-8/Jet-A fuel composition in detailed chemical kinetics development,”AIAA 2004-4207 (2004).

[9] Agosta, A., Cernansky, N. P., Miller, D. L., Faravelli, T., and Ranzi, E., “Reference com-ponents of jet fuels: kinetic modeling and experimental results,”Experimental Thermal and Fluid Science28(7), 701 – 708 (2004).

[10] Colket, M., Edwards, T., Williams, S., Cernansky, N. P., Miller, D. L., Egolfopoulos, F., Dryer, F. L., Bellan, J., Lindstedt, P., Seshadri, K., Pitsch, H., Sarofim, A., Smooke, M., and Tsang, W., “Identification of target validation data for development of surrogate jet fuels,”AIAA 2008-972 (2008).

[11] Edwards, T. and Maurice, L. Q., “Surrogate mixtures to represent complex aviation and rocket fuels,”Journal of Propulsion and Power 17(2), 461 – 466 (2001).

[12] Montgomery, C. J., Cannon, S. M., Mawid, M. A., and Sekar, B., “Reduced chemical kinetic mechanisms for JP-8 combustion,”AIAA 2002-0336 (2002).

[13] Violi, A., Yan, S., Eddings, E. G., Sarofim, A. F., Granata, S., Faravelli, T., and Ranzi, E., “Experimental formulation and kinetic model for JP-8 surrogate mixtures,”

Combustion Science and Technology 174(11-12), 399 – 417 (2002).

[14] Farrell, J., Cernansky, N., Dryer, F., Friend, D., Hergart, C., Law, C. K., McDavid, R., Mueller, C., Patel, A., and Pitsch, H., “Development of an experimental database and kinetic models for surrogate diesel fuels,” technical report, SAE (2007).

[15] de Oliveira, F. S., Teixeira, L. S. G., Araujo, M. C. U., and Korn, M., “Screening analysis to detect adulterations in Brazilian gasoline samples using distillation curves,”

Fuel 83(7-8), 917 – 923 (2004).

[16] Smith, B. L. and Bruno, T. J., “Improvements in the measurement of distillation curves.

3. Application to gasoline and gasoline + methanol mixtures,”Industrial and Engineer-ing Chemistry Research 46(1), 297–309 (2007).

[17] Huber, M. L., Lemmon, E. W., and Bruno, T. J., “Surrogate mixture models for the thermophysical properties of aviation fuel Jet-A,”Energy and Fuels 24(6), 3565–3571 (2010).

[18] Prausnitz, J. M., Lichtenthaler, R. N., and de Azevedo, E. G., [Molecular thermody-namics of fluid-phase equilibria], 3 ed. (1999).

[19] Andersen, V. F., Anderson, J. E., Wallington, T. J., Mueller, S. A., and Nielsen, O. J.,

“Distillation curves for alcohol-gasoline blends,”Energy and Fuels 24(4), 2683–2691 (2010).

[20] Smith, B. and Bruno, T., “Application of a composition-explicit distillation curve metrology to mixtures of Jet-A and S-8,” Journal of Propulsion and Power 24(3), 618 (2008).

[21] Rachner, M., [Die Stoffeigenschaften von Kerosin Jet A-1], DLR-Mitteilung. 98-01 (1998).

[22] Demirbas, A., “Distillation properties of various diesel oils,”Energy Sources, Part A:

Recovery, Utilization, and Environmental Effects 30(16), 1484–1490 (2008).

[23] Bugorkova, L. N., Burova, M. O., Kvasova, V. A., Leont’eva, S. A., Yanovskii, S. M., and Demidenko, K. A., “Distillation of petroleum crudes and products by ”Chromadis-tillation” method,”Chemistry and Technology of Fuels and Oils 24(7), 314–316 (1988).

[24] Polling, B., Prausnitz, J., and O’Connell, J., [The properties of gases and liquids]

(2001).

[25] Panagiotopoulos, A. Z. and Reid, R. C., [New mixing rule for cubic equations of state for highly polar, asymmetric systems], ch. 29, 571–582.

[26] Stryjek, R. and Vera, J. H., “PRSV: An improved Peng-Robinson equation of state for pure compounds and mixtures,”The Canadian Journal of Chemical Engineering, Can.

J. Chem. Eng.64(2), 323–333 (1986).

[27] Stryjek, R. and Vera, J. H., “PRSV - An improved Peng-Robinson equation of state with new mixing rules for strongly nonideal mixtures,”The Canadian Journal of Chemical Engineering, Can. J. Chem. Eng. 64(2), 334–340 (1986).

[28] Schwartzentruber, J., Galivel-Solastiouk, F., and Renon, H., “Representation of the vapor-liquid equilibrium of the ternary system carbon dioxide-propane-methanol and its binaries with a cubic equation of state : a new mixing rule,”Fluid Phase Equilib-ria38(3), 217 – 226 (1987).

[29] Martin, J. J., “Cubic equations of state-which?,”Industrial and Engineering Chemistry Fundamentals 18(2), 81–97 (1979).

[30] Trebble, M. A. and Bishnoi, P. R., “Accuracy and consistency comparisons of ten cubic equations of state for polar and non-polar compounds,”Fluid Phase Equilibria 29, 465 – 474 (1986).

[31] Span, R. and Wagner, W., “Equations of state for technical applications. I. Simultane-ously optimized functional forms for nonpolar and polar fluids,”International Journal of Thermophysics 24, 1–39 (2003).

[32] Sandler, S., ed., [Models for Thermodynamic and Phase Equilibria Calculations] (1994).

[33] Adachi, Y. and Sugie, H., “A new mixing rule - modified conventional mixing rule,”

Fluid Phase Equilibria28(2), 103 – 118 (1986).

[34] Sandoval, R., Wilczek-Vera, G., and Vera, J., “Prediction of ternary vapor-liquid equi-libria with the PRSV equation of state,”Fluid Phase Equilibria 52, 119 – 126 (1989).

[35] Wei, Y. and Sadus, R., “Equations of state for the calculation of fluid-phase equilibria,”

AIChE Journal46(1), 169–196 (2000).

[36] Michelsen, M., “Calculation of phase envelopes and critical points for multicomponent mixtures,”Fluid Phase Equilibria4(1-2), 1–10 (1980).

[37] Michelsen, M. L., “The isothermal flash problem. Part II. Phase-split calculation,”Fluid Phase Equilibria9(1), 21 – 40 (1982).

[38] Asselineau, L., Bogdanic, G., and Vidal, J., “A versatile algorithm for calculating vapour–liquid equilibria,”Fluid Phase Equilibria 3(4), 273 – 290 (1979).

[39] Soave, G., “Equilibrium constants from a modified Redlich-Kwong equation of state,”

Chemical Engineering Science 27(6), 1197–1203 (1972).

[40] Heidemann, R. A. and Khalil, A. M., “Calculation of critical points,” AIChE Jour-nal 26(5), 769–779 (1980).

[41] Michelsen, M. L. and Heidemann, R. A., “Calculation of critical points from cubic two-constant equations of state,”AIChE Journal27(1547-5905), 521–253 (1981).

[42] Boukouvalas, C., Magoulas, K., Tassios, D., and Kikic, I., “Comparison of the perfor-mance of the LCVM model (an EoS/GE model) and the PHCT EOS (the perturbed hard chain theory equation of state) in the prediction of the Vapor - Liquid equilibria of binary systems containing light gases,”Journal of Supercritical Fluids 19(2), 123–132 (2001).

[43] Lemmon, E., Huber, M., and McLinden, M., “NIST Reference Fluid Thermodynamic and Transport Properties Database (REFPROP). Version 8.0.” National Institute of Standards and Technology, Standard Reference Data Program.

[44] Laugier, S. and Richon, D., “Vapor-liquid equilibria for hydrogen sulfide + hexane, + cyclohexane, + benzene, + pentadecane, and + (hexane + pentadecane),”Journal of Chemical and Engineering Data40(1), 153–159 (1995).

[45] Olds, R. H., Reamer, H. H., Sage, B. H., and Lacey, W. N., “The n-butane-carbon dioxide system,”Industrial and Engineering Chemistry 41(3), 475–482 (1949).

[46] Richon, D., Laugier, S., and Renon, H., “High-pressure vapor-liquid equilibria for binary mixtures containing a light paraffin and an aromatic compound or a naphthene in the range 313-473 K,”Journal of Chemical and Engineering Data36(1), 104–111 (1991).

[47] Jangkamolkulchai, A. and Luks, K., “Partial miscibility behavior of the methane + ethane + n-docosane and the methane + ethane + n-tetradecylbenzene ternary mix-tures,”Journal of Chemical and Engineering Data 34(1), 92–99 (1989).

[48] Boukouvalas, C. J., Magoulas, K. G., Stamataki, S. K., and Tassios, D. P., “Prediction of vapor-liquid equilibria with the LCVM model: Systems containing light gases with medium and high molecular weight compounds,”Industrial and Engineering Chemistry Research36(12), 5454–5460 (1997).

[49] Hissong, D. W., A thermodynamic analysis of the critical properties of hydrocarbon mixtures, PhD thesis, Ohio State University (1968).

[50] Lemmon, E. W., Huber, M. L., and McLinden, M. O., “NIST standard reference database 23: Reference fluid thermodynamic and transport properties-REFPROP, Ver-sion 8.0,” (2007).

[51] Heneghan, S., Locklear, S., Geiger, D., Anderson, S., and Schulz, W., “Static tests of jet fuel thermal and oxidative stability,”Journal of Propulsion and Power 9(1), 5–9 (1993).

[52] Mawid, M. A. and Sekar, B., “Development of a detailed JP-8/Jet-A chemical kinetic mechanism for high pressure conditions in gas turbine combustors,”ASME Conference Proceedings 2006(42363), 369–388 (2006).

[53] Gauthier, B., Davidson, D., and Hanson, R., “Shock tube determination of ignition delay times in full-blend and surrogate fuel mixtures,”Combustion and Flame139(4), 300 – 311 (2004).

[54] Slavinskaya, N., Haidn, O., Donato, F., and Aigner, M., “Modeling of reference fuel combustion with PAH formation,”AIAA 2006-1511 (2006).

[55] Troe, J., “Theory of thermal unimolecular reactions in the fall-off range. I. Strong colli-sion rate constants,”Berichte der Bunsengesellschaft fuer physikalische Chemie87(2), 161–169 (1983).

[56] Gilbert, R. G., Luther, K., and Troe, J., “Theory of thermal unimolecular reactions in the fall-off range. II. Weak collision rate constants,”Berichte der Bunsengesellschaft fuer physikalische Chemie 87, 169–177 (1983).

[57] Benson, S. W. and Buss, J. H., “Additivity rules for the estimation of molecular prop-erties. Thermodynamic propprop-erties.,”Journal of Chemical Physics29, 546–572 (1958).

[58] Wilhoit, R. C., “Thermodynamics research center current data news,” Technical Report 3,2, Texas A & M University (1975).

[59] Curran, H. J., Gaffuri, P., Pitz, W. J., and Westbrook, C. K., “A comprehensive mod-eling study of iso-octane oxidation,”Combustion and Flame 129(3), 253 – 280 (2002).

[60] Curran, H., Gaffuri, P., Pitz, W., and Westbrook, C., “A comprehensive modeling study of n-heptane oxidation,”Combustion and Flame 114(1-2), 149 – 177 (1998).

[61] Wallington, T., Dagaut, P., and Kurylo, M., “Ultraviolet-absorption cross-sections and reaction-kinetics and mechanisms for peroxy-radicals in the gas-phase,”Chemical Re-views 92, 667–710 (1992).

[62] Tsang, W. and Hampson, R., “Chemical kinetic database for combustion chemistry.

1. Methane and related-compounds,” Journal of Physical and Chemical Reference Data 15(3), 1087–1279 (1986).

[63] Lutz, A., Kee, R., and Miller, J., “SENKIN: A Fortran Program for predicting homoge-neous gas phase chemical kinetics with sensitivity analysis,” Technical Report Report No. SAND 87-8248, Sandia National Laboratories (1988).

[64] Warnatz, J., [Rate coefficients in the C/H/O system], Combustion Chemistry (1984).

[65] Vajda, S., Valko, P., and Turanyi, T., “Principal component analysis of kinetic models,”

International Journal of Chemical Kinetics 17(1), 55–81 (1985).

[66] Zsely, I. G., Zador, J., and Turanyi, T., “Uncertainty analysis of updated hydrogen and carbon monoxide oxidation mechanisms,” Proceedings of the Combustion Insti-tute 30(1), 1273 – 1281 (2005).

[67] Zsely, I. G., Zador, J., and Turanyi, T., “Uncertainty analysis of NO production dur-ing methane combustion,”International Journal of Chemical Kinetics 40(11), 754–768 (2008).

[68] Saltelli, A., Andres, T. H., and Homma, T., “Sensitivity analysis of model output : An investigation of new techniques,”Computational Statistics and Data Analysis 15(2), 211 – 238 (1993).

[69] Saltelli, A. and Homma, T., “Sensitivity analysis for model output : Performance of black box techniques on three international benchmark exercises,”Computational Statistics and Data Analysis 13(1), 73 – 94 (1992).

[70] Turanyi, T., Berces, T., and Vajda, S., “Reaction rate analysis of complex kinetic systems,”International Journal of Chemical Kinetics 21(2), 83–99 (1989).

[71] Wei, J. and Kuo, J., “A lumping analysis in monomolecular reaction systems: Analysis of the exactly lumpable system,” Industrial and Engineering Chemistry Fundamen-tals 8(1), 114–123 (1969).

[72] Li, G. and Rabitz, H., “Reduced kinetic equations of a CO/H2/air oxidation model by a special perturbation method,”Chemical Engineering Science 52(23), 4317 – 4327 (1997).

[73] Li, G. and Rabitz, H., “Combined symbolic and numerical approach to constrained non-linear lumping - With application to an H2/O2 oxidation model,”Chemical Engineering Science 51(21), 4801 – 4816 (1996).

[74] Li, G. and Rabitz, H., “A lumped model for H2/O2 oxidation in the oscillatory regime,”

The Journal of Chemical Physics 102(18), 7006 – 7016 (1995).

[75] Li, G., Rabitz, H., and Toth, J., “A general analysis of exact nonlinear lumping in chemical kinetics,”Chemical Engineering Science 49(3), 343 – 361 (1994).

[76] Li, G., Tomlin, A. S., Rabitz, H., and Toth, J., “A general analysis of approximate nonlinear lumping in chemical kinetics. I. Unconstrained lumping,” The Journal of Chemical Physics 101(2), 1172 – 1187 (1994).

[77] Li, G. and Rabitz, H., “The direct lumping approach: an application to a catalytic reforming model,”Chemical Engineering Science 48(10), 1903 – 1909 (1993).

[78] Li, G., Tomlin, A. S., Rabitz, H., and Toth, J., “Determination of approximate lumping schemes by a singular perturbation method,”The Journal of Chemical Physics 99(5), 3562 – 3574 (1993).

[79] Li, G. and Rabitz, H., “Determination of constrained lumping schemes for nonisother-mal first-order reaction systems,” Chemical Engineering Science 46(2), 583 – 596 (1991).

[80] Li, G. and Rabitz, H., “New approaches to determination of constrained lumping schemes for a reaction system in the whole composition space,” Chemical Engineer-ing Science 46(1), 95 – 111 (1991).

[81] Li, G. and Rabitz, H., “A general lumping analysis of a reaction system coupled with diffusion,”Chemical Engineering Science 46(8), 2041 – 2053 (1991).

[82] Li, G. and Rabitz, H., “A general analysis of approximate lumping in chemical kinetics,”

Chemical Engineering Science 45(4), 977 – 1002 (1990).

[83] Li, G. and Rabitz, H., “A general analysis of exact lumping in chemical kinetics,”

Chemical Engineering Science 44(6), 1413 – 1430 (1989).

[84] Peters, N. and Kee, R., “The computation of stretched laminar methane-air diffusion flames using a reduced four-step mechanism,” Combustion and Flame 68(1), 17–29 (1987).

[85] Peters, N. and Williams, F., “The asymptotic structure of stoichiometric methaneair flames,”Combustion and Flame 68(2), 185–207 (1987).

[86] Sung, C., Law, C., and Chen, J.-Y., “Augmented reduced mechanism for methane oxida-tion with comprehensive global parametric validaoxida-tion,”Proceedings of the Combustion Institute 27(1), 295–304 (1998).

[87] Maas, U. and Pope, S., “Implementation of simplified chemical kinetics based on intrin-sic low-dimensional manifolds,”Proceedings of the Combustion Institute24(1), 103–112 (1992).

[88] Maas, U., “Efficient calculation of intrinsic low-dimensional manifolds for the simplifi-cation of chemical kinetics,”Computing and Visualization in Science 1, 69–81 (1998).

[89] Bykov, V., Goldfarb, I., and Gol’dshtein, V., “Novel numerical decomposition ap-proaches for multiscale combustion and kinetic models,”Journal of Physics: Conference Series 22(1), 1–29 (2005).

[90] Bykov, V., Goldfarb, I., and Gol’dshtein, V., “Singularly perturbed vector fields,” Jour-nal of Physics: Conference Series55(1), 28–44 (2006).

[91] Lu, T. and Law, C. K., “A directed relation graph method for mechanism reduction,”

Proceedings of the Combustion Institute 30(1), 1333–1341 (2005).

[92] Mauss, F., Peters, N., Rogg, B., and Williams, F. A., [Reduced kinetic mechanisms for applications in combustion systems], Lecture Notes in Physics (1993).

[93] Bilger, R., Starner, S., and Kee, R., “On reduced mechanisms for methane-air combus-tion in nonpremixed flames,”Combustion and Flame 80(2), 135–149 (1990).

[94] Jones, W. and Lindstedt, R., “Global reaction schemes for hydrocarbon combustion,”

Combustion and Flame 73(3), 233–249 (1988).

[95] Andersen, J., Rasmussen, C., Giselsson, T., and Glarborg, P., “Global combustion mechanisms for use in CFD modeling under oxy-fuel conditions,” Energy and Fu-els23(3), 1379–1389 (2009).

[96] Li, S. C. and Williams, F. A., “Reaction mechanisms for methane ignition,”Journal of Engineering for Gas Turbines and Power 124, 471–480 (2002).

[97] Novosselov, I. and Malte, P., “Development and application of an eight-step global mechanism for CFD and CRN simulations of lean-premixed combustors,”Journal of Engineering for Gas Turbines and Power 130(2) (2008).

[98] Nicol, D., Malte, P., Hamer, A., Roby, R., and Steele, R., “Development of a five-step global methane oxidation-NO formation mechanism for lean-premixed gas turbine combustion,” Journal of Engineering for Gas Turbines and Power 121(2), 272–280 (1999).

[99] Meredith, K. and Black, D., “Automated global mechanism generation for use in CFD simulations,”AIAA 2006-1168, 14161–14173 (2006).

[100] Chen, J.-Y., “A general procedure for constructing reduced reaction mechanisms with given independent relations,”Combustion Science and Technology57, 89–94 (1988).

[101] Silin, I., “FUMILI,” technical report, CERN Program Library D 510 (1983).

[102] Sirjean, B., Dames, E., Sheen, D. A., You, X.-Q., Sung, C., Holley, A. T., Egolfopoulos, F. N., Wang, H., Vasu, S. S., Davidson, D. F., Hanson, R. K., Pitsch, H., Bowman, C. T., Kelley, A., Law, C. K., Tsang, W., Cernansky, N. P., Miller, D. L., Violi, A., and Lindstedt, R. P., “A high-temperature chemical kinetic model of n-alkane oxidation, JetSurF version 1.0,” (2009).

[103] Goos, E., Burcat, A., and Ruscic, B., “Ideal Gas Thermochemical Database with up-dates from Active Thermochemical Tables,” (2010).

[104] Scott, D., “Chemical thermodynamic properties of hydrocarbons and related sub-stances. Properties of the alkane hydrocarbons, C1 through C10 in the ideal gas state from 0 to 1500 K.,” Bulletin 666, U.S. Bureau of Mines (1974).

[105] Dagaut, P. and Gail, S., “Chemical kinetic study of the effect of a biofuel additive on Jet-A1 combustion,”Journal of Physical Chemistry A111(19), 3992–4000 (2007).

[106] Slavinskaya, N. and Haidn, O., “Modeling of n-heptane and iso-octane oxidation in air,”

Journal of Propulsion and Power 19(6), 1200–1216 (2003).

[107] Slavinskaya, N. A., Zizin, A., and Riedel, U., “Towards kerosene reaction model devel-opment: propylcyclohexane, cyC9H18, n-dodecane, C12H26, and hexadecane C16H34 combustion,”AIAA 2010-605 (2010).

[108] Fieweger, K., Blumenthal, R., and Adomeit, G., “Self-ignition of S.I. engine model fuels:

A shock tube investigation at high pressure,”Combustion and Flame109(4), 599 – 619 (1997).

[109] Wang, H., Warner, S. J., Oehlschlaeger, M. A., Bounaceur, R., Biet, J., Glaude, P.-A., and Battin-Leclerc, F., “An experimental and kinetic modeling study of the autoignition of alpha-methylnaphthalene/air and alpha-methylnaphthalene/n-decane/air mixtures at elevated pressures,”Combustion and Flame 157(10), 1976–1988 (2010).

[110] Mati, K., Ristori, A., Gail, S., Pengloan, G., and Dagaut, P., “The oxidation of a diesel fuel at 1-10 atm: Experimental study in a JSR and detailed chemical kinetic modeling,”

Proceedings of the Combustion Institute 31, 2939 – 2946 (2007).

[111] Bikas, G. and Peters, N., “Kinetic modelling of n-decane combustion and autoigni-tion: Modeling combustion of n-decane,”Combustion and Flame126(1-2), 1456 – 1475 (2001).

[112] Nehse, M., Warnatz, J., and Chevalier, C., “Kinetic modeling of the oxidation of large aliphatic hydrocarbons,”Proceedings of the Combustion Institute26(1-2), 773–780 (1996).

[113] Ranzi, E., Frassoldati, A., Granata, S., and Faravelli, T., “Wide-range kinetic modeling study of the pyrolysis, partial oxidation, and combustion of heavy n-alkanes,”Industrial and Engineering Chemistry Research44(14), 5170 – 5183 (2005).

[114] Pfahl, U., Fieweger, K., and Adomeit, G., “Self-ignition of diesel-relevant hydrocarbon-air mixtures under engine conditions,”Proceedings of the Combustion Institute 26(1), 781 – 789 (1996).

[115] Dean, A., Penyazkov, O., Sevruk, K., and Varatharajan, B., “Autoignition of surrogate fuels at elevated temperatures and pressures,” Proceedings of the Combustion Insti-tute 31(2), 2481–2488 (2007).

[116] Shen, H.-P. S., Steinberg, J., Vanderover, J., and Oehlschlaeger, M. A., “A shock tube study of the ignition of n-heptane, n-decane, n-dodecane, and n-tetradecane at elevated pressures,”Energy and Fuels 23, 2482–2489 (2009).

[117] Vasu, S. S., Davidson, D. F., Hong, Z., Vasudevan, V., and Hanson, R. K., “n-Dodecane oxidation at high-pressures: Measurements of ignition delay times and OH concentra-tion time-histories,”Proceedings of the Combustion Institute 32(1), 173–180 (2009).

[118] Vasu, S. S., Davidson, D. F., and Hanson, R. K., “Jet fuel ignition delay times: Shock tube experiments over wide conditions and surrogate model predictions,”Combustion and Flame 152(1-2), 125 – 143 (2008).

[119] “KINALC program package.” http://garfield.chem.elte.hu/Combustion/kinalc.htm.

[120] Slavinskaya, N. and Lenfers, C., “Skeletal mechanism production for n-decane,” 2nd Int. Workshop on Model Reduction in Reacting Flow (2007).