You are on page 1of 9

Twenty-Fourth Symposium (International) on Combustion/The Combustion Institute, 1992/pp.

833-841

A STUDY ON ETHANOL OXIDATION KINETICS IN LAMINAR PREMIXED FLAMES, FLOW REACTORS, AND SHOCK TUBES F. N. EGOLFOPOULOS Department of Mechanical Engineering University of Southern California, Los Angeles, CA 90089-1453
AND

D. X. DU AND C. K. LAW Department of Mechanical and Aerospace Engineering Princeton University, Princeton, N] 08544-5263

A comprehensive experimental and numerical study on ethanol oxidation kinetics has been conducted. The laminar flame speeds of ethanol/air mixtures were determined by using the counterflow twin-flame technique at 1 atm pressure and for initial mixture temperatures between 363 and 453 K. A detailed kinetic scheme was subsequently compiled by grafting the latest information on ethanol kinetics onto a previously developed methanol scheme, and was found to be self-consistent in that it closely predicts not only the experimental laminar flame speeds of ethanol, but also those of methane, methanol, and all the C2-hydrocarbons. Further recognizing that prediction of the laminar flame speeds is not sufficient for the satisfactory validation of a kinetic mechanism, the present scheme has also been tested against experimental data in the literature on the species and temperature profiles in flow reactors and on the ignition delay times in shock tubes. Such studies demonstrate the importance of the CH:3 and HO2 radical chemistry, and the present results suggest that the rate of CH3 + HO2 CHaO + OH may be slower while that of CHa + HO2 ~ CH4 + O~ may be faster than values frequently used in recent literature.

Introduction
Based on economic and environmental considerations, in the past decade or so the potential and concern with the use of alcohols as alternate fuels have stimulated extensive fundamental studies on their combustion behavior. 1-7 Methanol, being the simplest of the alcohols, has been extensively studied 1-6 and its oxidation and pyrolysis steps can be considered to be reasonably well understood. Ethanol, the next simplest alcohol, has also been extensively studied experimentally. 4'5'7 Its kinetics, however, was much less understood until the recent contribution of Norton 4 and Norton and Dryer z'7 from their turbulent flow reactor studies. Their studies demonstrated the importance of the branching ratios for the initial fuel reaction to produce the three C.2H50 isomers, which subsequently react through different routes to form different products. It was emphasized, however, that since these kinetic understandings were obtained from flow reactor studies, some of the steps in the proposed reaction scheme need to be further tested for different reaction conditions. Flame studies can contribute towards the com-

prehensive validation of a mechanism because of the large variations of temperature and species concentrations associated with the flame structure. Such studies, however, are useful only if the experimental data are obtained from configurations which can be modelled with confidence. For example, it has been well established that determination of the laminar flame speed can be significantly and systematically 'affected by the presence of stretch effects, s In view of the above considerations, and further recognizing that ethanol flame studies are limited 9-1z while detailed modeling of the flame propagation and structure have not been conducted, the first objective of the present investigation is to provide accurate experimental data on the laminar flame speed by using the counterflow twin-flame techniquesA3'14 which systematically eliminates the stretch effects. These data are useful not only for the fundamental study of ethanol kinetics, but their quantitative values are also needed for the simulation of practical combustion situations. Our second objective is then to compile an ethanol kinetic mechanism and compare the calculated laminar flame speeds with the experimental data.

833

834

LAMINAR FLAMES--KINETIC STRUCTURES of C2H~O were taken from Burcat. 23 The transport properties of CzHsOH were taken from Monchick and Mason z4 and those of the three isomers of C2H~O were calculated by using the approach of Prausnitz, z5 in which the Lennard-Jones parameters for the radical complexes were estimated by averaging two structurally similar molecules with the appropriate mixing rules. For example, the structure of C2H50 was assumed to be "'between" those of C2HsOH and C2H5 and its potential well depth was determined as: (e/Kn)c~HsO = [(E/KB)czHsOH >( (e / KB)czlas] 1/2 and the collision diameter as: (O-)czH~o = (O'czHsOH+ O'C2Hs)/2. For the laminar flame speed determination, the freely-propagating flame version of the one-dimensional code lsA9 was used, and thermal diffusion was included. For the simulation of flow reactors and shock tubes, the SENKIN code2~ was used which calculates the homogeneous gas phase chemical kinetics.

Through such comparisons, and by using sensitivity and species consumption path analyses, the accuracy of the mechanism will be tested and enhanced insight into the controlling kinetic processes will be gained. The third objective is to further test the mechanism against data obtained from flow reactors and shock tubes. These comparisons allow the assessment of certain aspects of the mechanism which are insensitive to laminar flame speed studies. These further comparisons involve the temperature range from 1000 K up to 1700 K and the stoichiometry range from fuel lean to fuel rich conditions. In the following we shall first specify the experimental and numerical aspects of the present investigation. This will be followed by comparison and discussion of the results from the various studies involving laminar flame speeds, flow reactors, and shock tubes. Experimental Methodology The counterflow twin flame technique for the determination of laminar flame speeds is well documented. 13A4 It involves the establishment of two symmetrical, planar, nearly-adiabatic flames in a nozzle-generated counterflow configuration, and the subsequent determination of the axial velocity profile along the centerline of the flow by laser Doppler velocimetry. The minimum point of the velocity profile is identified as a reference upstream flame speed, S,, while the velocity gradient ahead of this point characterizes the imposed stretch rate, K. Thus by plotting Su versus K, the stretch-free flame speed S~ is obtained through linear extrapolation to zero stretch. Some recent theoretical studies 15'1~ have indicated that the linear extrapolation may yield higher values for S~. These concerns appear to be based on higher order considerations whose effects seem to be less than 5 to 10%. Since there has not been any systematic experimental verification of the extent of these effects, it is believed that the present data are still the most accurate and as such will be used in comparisons with the numerical calculations. The flow control of ethanol, which is liquid under standard conditions, was similar to that used in the previous experimental work on methanol and the details can be found in Ref. 6.

Study Based on Laminar Flame Speed Results


The experimentally determined laminar flame speeds, S~, of ethanol/air mixtures at atmospheric pressure and for unburned mixture temperatures, Tu, of 363, 428, and 453 K are shown in Fig. 1 for a wide range of stoichiometry. The flame speeds at 298 K were not measured but were obtained by extrapolating the data obtained at higher Tu data to 298 K. For the temperature range investigated, the results show that S~ basically increases linearly with Tu at a given stoichiometry, as expected. The maximum S~ for ethanol/air was found to occur around ~b = i. 15, which is higher than the corresponding values of about ~b = 1.05 for alkane/ air flamesIs but is similar to the value for methanol/air flames. 6 The reason for such a difference with the alkane flames is the presence of the extra O atom in the ethanol molecule which provides an additional amount of "oxidizer" as the fuel concentration becomes richer. Figure 1 shows the numerically calculated results and their close agreement with the experimental data. The kinetic scheme includes 35 species and 196 reversible reactions for the C1, C2, CH3OH, and C2HsOH submechanisms. Its compilation was conducted iteratively since the validity of certain reactions could not be assessed by using the flame speed data alone. The C j, C2, and CH3OH submechanisms, involving 30 species and 171 reversible reactions, were initially compiled in Egolfopoulos et al. 6 This mechanism was further revised

Numerical Methodology
Numerical simulations of all the reaction systems were conducted by using CHEMKIN-based programs. 17-21 The thermodynamic data base developed by Kee et al. ~z was used and the thermodynamic properties of C2HsOH and the three isomers

STUDY ON ETHANOL OXIDATION KINETICS


100

835

8O

ethanol/air flame is shown in Fig. 2. These paths were determined by integrating all reactions through the flame and determining the fraction of each species consumed by a specific reaction; this fraction is indicated next to each species. From such a study the following observations can be made: (1) CzHsOH: It is mainly attacked by H, O, and OH which abstract one H and create the three CzH50 isomers. The presence of the O atom in the molecule clearly alters the initial products and thereby the subsequent kinetic steps as compared to those involving the corresponding alkane, C2H6. Significant differences are also observed between CzH5OH and CH3OH in that the main initial products are CHzOH and CH30 for methanol while they are not as important in ethanol oxidation. For fuel rich concentrations it is found that the reaction of fuel with H increases due to reduced concentrations of OH and O. (2) CH3CHOH: This is the major product of the initial fuel reactions. For all stoichiometries it reacts only with Oz to produce acetaldehyde (CH3HCO) and HOe. Therefore, similar to methanol oxidation, a large amount of HOe is produced during ethanol oxidation by means other than the termination reaction H + 02 + M ~ HOe + M. As such, the HOe chemistry is expected to be important. Furthermore, the substantial production of CH3HCO constitutes a potential emission problem, similar to the that of formaldehyde in methanol oxidation. CH3HCO is subsequently completely converted to the methyl radical (CH3). (3) CeH4OH: It thermally decomposes to C2H4 which is the primary production step of ethene in ethanol oxidation. Subsequent reactions of ethene lead to 60% conversion to C2 species through H and OH reactions and 40% conversion to CH3 through O and OH reactions. The production of Ce species becomes about 85% for rich flames due to the abundance of H radicals and the production of CH3 becomes 65% for lean flames due to the abundance of O radicals. (4) CH3CHeO: It exists in smaller concentrations as compared to the other isomers. It thermally decomposes to CH3 with the simultaneous production of CHeO which, however, exists in much lower concentration levels as compared to those in methanol flames. (5) CH3: It is the major product from all three C2H50 isomer reaction channels and its subsequent reactions are expected to be important in ethanol oxidation. This is similar to methane oxidation in which CH 3 is produced directly from the fuel. The main difference from the methane mechanism is that since CH3 and HOe both exist in large concentrations in ethanol oxidation, the interaction of these two radicals is expected to be important. It may also be noted that reactions between CH3 and HOe

o~

o 1o0

o~ ~4o

o 0.4

T of6

ol

1~' 112 ',i, '1~ EquivalenceRatio,

lls

2.0

FIG. 1. Experimentally and numerically determined laminar flame speeds, S~ T,,), at p = 1 arm, using the present kinetic scheme.

in order to provide better agreements on methanol/air laminar flame speeds and involves 31 species and 175 reversible reactions. In the new CH3OH submechanism lower rates for the reactions CH3 + OH~------CH30 + H 7 and HOe + H - - O H + OH z6 were used. Furthermore, stronger collision efficiency for H20 in the H + O2 + M = H O e + M reactione7 was used as well as detailed ICHe and 3CHe submechanisms, es Suffice to note that this mechanism has been shown to satisfactorily predict a number of oxidation properties of C1- and C2-hydrocarbons and methanol mixtures with oxygen and inert. The only modifications of the scheme implemented herein are reactions (R1) and (-R2) which will be discussed later. For the CzHsOH submechanism the latest flow reactor oxidation result of Norton and Dryer 7 was used. Falloff corrections and appropriate third body efficiencies have been used for all pressure dependent reactions. In order to utilize the flame speed comparisons for validation of the kinetic scheme, it is important to identify the main species consumption paths in these flames and conduct sensitivity analysis18 ' 19 by determining the influence of all reaction rates on the mass burning rate m ~ = p~,S ~ where pu is the density of the unburned mixture. The species consumption paths for stoichiometric

836

LAMINAR FLAMES--KINETIC STRUCTURES C2HsOH I

I
I c: 4o. I
M(1.0) ~
, ,

OH (0.26)

H (0,05) O (0,05) OH (0.43)

O (0.06) / OH (0.09)

CH3CHOH ] 02 (1.0)

CH3CH20

I C2HsOH/ rai1 p=l atm i ~=1.0 Tu=363 K

M (1.0)

H(0.34)~O (0.29)

[H (0.41) | .19) M (1.0)

H (0.16) OH (0.19)

OH (0.I 1)

FiG. 2. Species consumption path analysis for stoichiometric ethanol/air flame (~b = 1.0) at T, = 363 K and p = 1 atm. arc of interest in the high pressure combustion of methane and ethane, although repurted rates have high uncertainty. The first order normalized sensitivity coefficientslS.t9 of selected reactions on the mass burning rate, m", of lean, stoichiometric, and rich flames are shown in Fig. 3. Similar to the hvdrocarbon and methanol flames, the main branching reaction H + O2 ---> OH + O and the CO oxidation reaction CO + OH ~ CO2 + H have dominant effect on m". Furthermore, similar to the methane flames above 5 atm, and methanol flames at atmospheric pressure, the HOz chemistry appears to be important, with noticeable sensitivity for the CIt3 + HO2 ----> CI|30 + OH reaction. Theonly ethanol-specific reaction appearilag to be important is C2HsOH + Oil ---) C21t401t + H20, which favors propagation of lean and stoichiometric flames but retards the propagation of rich flames. The reason for such a qualitative difference is that C2H4OH i s subsequently converted to CzH4 which is consumed by O radicals, in addition to tt and OH, and O radicals exist in very small quantities for fuel rich mixtures. The main intermediates CH3HCO and CH3CHzO of the other two possible paths of the initial fuel consumption reactions do not require O for their oxidation. The sensitivities of HCO + M ---) H + CO + M and HCO + O 2 ~ C O + HO2 arc similar to those in methanol flames. Reference 6 gives a detailed explanation of the rule of these reactions and their sensitivity sign reversal as stoichiometry changes.

Study Based on Flow Reactor Results

Turbulent flow reactors typically operate in temperature ranges of about 1000 to 1200 K, and the system is near-adiabatic before any significant amount of heat is released. In the present investigation the data of Norton and Dryers'r under oxidation conditions were simulated by using SENKIN 24 for a constant pressure, adiabatic system which closely resembles the experimental configuration. When the original (;1, C2, and CH3OH submechanisms were used in the simulations, the predicted ethanol profile was too "steep" while the temperature also increases too fast as compared to the experimental

STUDY ON ETHANOL OXIDATION KINETICS

837

HO2+H=OH+OH

llI~H+HO2=H2+O2 - - ' 1 Illll~ H+O2+M=HO2+M l l ~ H+OH+M=H20+M H+O2___OH~ . ~ C2H3+H=C2H2+H2 CH3+HO2=CH30+OH


co+oHio2 ~

~l [ C2HsOH/air /I / p=l atm / [ [ Tu=363K /I " - "l l ~ I I f"

1.2

~ ! 1.o ! ~ 0.8 '~'~0.6 -~ 0.4

f | A Experunentel by NortonandDryer ..~.~.. Numerical withModified Ratesfor - - CH3+HO2__>products NuraeticalwithLiterature Rates for
/

CH_3_3+H~-~l~roducts

I
I * = 1.8 I ll~ = 1.0 ] l~= 0.6 9 I~ 0.4 0.5

0~ .
0.0 1.2 1.0 r 0.8 O *~0.6

\ N~ ;,, ~ 4

/-FlowReactor-Oxidation'~ l C2H5OHProfile /

-0.2

-0.1

HCO+O2=CO+HO 2 HCO+M=H+CO+M C2H5OH+O C2HsOH+O H H4OH+H20 0.0 0.1 0.2 0.3

"'-..'~..."

~=0.81

Normalized Sensitivity on Mass Burning Rate FIG. 3. Normalized first order sensitivity coefficients of the most important reactions on the mass burning rate of lean, stoichiometric, and rich ethanol/air flames at T, = 363 K and p = 1 atm. data. This behavior was observed for both lean (~b = 0.81) and rich (~b = 1.24) conditions. In order to resolve this disagreement, the sensitivities of the reactions involving CzHsOH, CeH4OH, CH3CHOH, CH3CHeO, CH3HCO, C2H4, C2H3, CH4, CH3, CHeO, HCO, HOz, and HeOe to the fuel decay profile were examined in detail. These species were chosen because they are the important intermediates subsequent to fuel decay. It was consequently found that the reactions having the greatest effect on the slope of the ethanol profile were those between CH3 and HO2: CH3 + HOe ~ CH30 + OH CH3 + HOe--> CH4 + O2 (R1) (-R2)

-~ o.4
0.2

\',.~

['FlowReactor-Oxidation) / C2H'OHPr~ /

20

40

60

80

100

Time,msec FIG. 4. Comparison between the experimental (symbols) profiles for C2HsOH at various q~'s as determined by Norton and Dryer7 in a flow reactor, and numerical calculations (lines) using the present scheme with the rates for reactions CH3 + HO2 "---> Products taken from Tsang and Hampson29 as well as the present scheme with the modified rates.

This result is not surprising because CH3 and HOe are key radicals in ethanol oxidation with CH3 being the main product of all reaction channels. More specifically, it was found that a slower rate for (R1) and a faster rate for (-R2) would lead to milder ethanol profiles because (R1) produces less stable species as compared to (-R2). In the initial scheme 6 the rates of (R1) and (R2) were those recommended by Tsaug and Hampson 29 with uncertainty factors of 3 and 5 respectively, as well as the rate of Norton and Dryer ~ for (R1). Consequently we adopted the rate of Dagaut et al.30 for (R1) which is 6 and 3 times slower at 1000 K as compared to the rates of Tsang and Hampsone9 and Norton and Dryer7 respectively. Furthermore, we also used a rate which is 5 times faster for (R2) than the rate of Tsang and Hampson. Specifically, these reaction rates are expressed as: CH3 + HOe--> CH30 + OH [k = 4.00E13 T~176 CH4 + O2--->CH3 + HO2 [k = 20.3E13 T~176 (R1) (a2)

where k is the specific reaction rate, T the temperature, R the universal gas constant, and the units are moles, cubic centimeters, seconds, Ke|vins, and calories/mole. Results of the calculations based on both the original and modified mechanisms are shown in Fig. 4 together with the experimental data. Improved agreement is obtained for both lean and rich conditions. Variation of the branching ratio leading to the three C2H~O isomers results in different induction times in some cases, but the slope of the ethanol profile is influenced minimally. The initial temperature of the mixture was also reduced by 10 K, which is the experimental uncertainty, and no significant change in the ethanol profile was found. In Fig. 5 we compare the experimental data of Norton and Dryer 57 9 for lean conditions (~b = 0.81) with the calculated results obtained by using the modified scheme. Due to the uncertainty in the experimental induction time, the numerical results had to be "time shifted "'2'3'6 in order to match the experimental data at the 50% fuel decay point. The amount of shift is reported in the figure caption. The results show that there is good agreement for

838
1.6 [ 1.4 f~. . . . . . . . . . . . . - - - I1 CO

LAMINAR FLAMES--KINETIC STRUCTURES


CO2~ - - - O H2 1400 1350 I

1.2k "-~
0

l .....

/ ........9 M~o~0s ~
~

5OHx2 "'"" " T,KI


/

J. -'?" .:'_':'-'-'-'A -u A
/

....
A

1300 1250

~40o

Shock Tube-Oxidation (Natarajan and B h a s k a r a n ) )

.~

' ,.

-1201) ~

*i
300
u

l.
0.2 -10 0 10 20

..........."6"

4,
O
9 9 9

I150 [~

1050
100

"r'

nn

n a a

30

40 50 60 Time, msec

70

80

90

n Experimental --Numerical at T i ~ o 12oo 1300 1400

o a u

n [] nao [] 1700

1500

1600

FIG. 5. Comparison between experimental (symbols) flow reactor oxidation data for ~b = 0.81 as determined by Norton and Dryer5'7 and numerical calculations (lines) using the present scheme. The numerical results have been "shifted" by - 9 msec. (0.588% CzHsOH, 2.185% O~, 97.227% N2, T~, = 1090 K, p~, = 1.0 atm, adiabatic, isobaric; all concentrations on per mole basis) most of the species until the fuel is consumed and before the CO oxidation is initiated. In this regime heat loss to the wall (maintained at -1000 K) and/ or molecular diffusion become significant and the adiabatic, homogeneous model is probably not a good representation of the experimental conditions and thus discrepancies in the CO oxidation region immediately following the fuel consumption zone were not further assessed. The calculated profiles for the minor species such as CgH4, CH4, CgH0, and CgH2 agree reasonably with the experimental data. The calculated profiles for CHaHCO, while agreeing closely with the experiments during the production stage, the subsequent consumption rates are found to be higher than the experimental ones. Similar results for the major and minor species are obtained for the rich conditions (ok = 1.24). The modification in the rate of the reaction CH3 + HOg ~ Products has a small effect (0.5-1.0 cm/s) on the calculated laminar flame speeds.

InitialMixtureTemperature(Ti),K Fic. 6. Comparison between experimental (symbols) shock tube ignition delay data as determined by Natarajan and Bhaskaran 31 and numerical calculations (lines) using the present scheme. (2.5% C2HsOH, 7.5% 02, 90.0% Ar, P,n = 1.0 atm, adiabatic, isochoric; all concentrations on per mole basis) abatic, constant volume system, which allows for both temperature and pressure increases during the course of the reaction. In Fig. 6 the experimental data of Natarajan and Bhaskaran31 for a stoichiometric mixture of C2H3OH/ O2/Ar at 1 atm are compared with the numerical calculations. The experimental data were determined by monitoring the OH concentration and ignition was assumed to occur when the OH concentration reached 2 10-9 mole/cc, The comparison in Fig. 6 can be considered to be quite favorable, especially in view of the +-50 K uncertainty in the temperature determination33'34 in shock tube experiments. Similar agreements were obtained by modeling mixtures at different stoichiometries and pressures. It may be noted that previous modeling efforts31'35'36 resulted in significant disagreements with the data of Natarajan and Bhaskaran, 31 even allowing for the uncertainty in the temperature. The experimental data of Cooke et al. 32 were also compared favorably with the numerical calculations. Sensitivity analysis showed that among the fuel consumption reactions the ignition delay time is mostly sensitive to the decomposition reaction C2HsOH ---> CH3 + CH2OH. Noticeable sensitivity was also found for reactions (R1) and (-R2).

Study Based on Shock Tube Results While reactor-type experiments provide information for the low- and intermediate-temperature regimes, shock tube studies operate at the higher temperature range of 1300 K to 2500 K and thereby complete t h e temperature range of relevance in combustion. We have thus tested the present ethanol mechanism against the high temperature ignition delay data of Natarajan and Bhaskaransl and Cooke et al. 32 Since in both investigations the measurements were conducted behind reflected shock waves, this configuration was numerically simulated as an adi-

Concluding Remarks
In the present investigation, we first determined the laminar flame speeds of atmospheric ethanol/ air mixtures by using the counterflow technique. These data were obtained for various mixture initial temperatures and stoichiometries ranging from very

STUDY ON ETHANOL OXIDATION KINETICS lean to very rich, and can be used with reasonable confidence for the study of ethanol kinetics as well as for practical applications. An ethanol oxidation kinetic scheme was then compiled by simultaneously testing it against the experimental data from flame speed measurements, flow reactors, and shock tubes. Modeling of the various experimental configurations provided complementary information on the kinetics, which are essential for the comprehensive validation of the mechanism. The comparison indicates a slower rate for the reaction CH3 + HOz CH30 + OH and a faster rate for the reaction CH3 + HO2 ---) CH4 + Oz than those frequently used in recent literature. It is clear that such an observation, while still tentative, could not be obtained from studies on laminar flames speeds alone. Furthermore, by using the present kinetic scheme, sensitivities and species consumption paths in ethanol flames were identified and the importance of the CHa-chemistry demonstrated. The shock tube calculations also yielded satisfactory comparisons. It must be emphasized that caution is still needed when applying the present scheme to burning and reacting conditions which deviate substantially from the range of parameters and conditions investigated herein. Further research is also needed in the various kinetic aspects including branching ratios and pyrolysis steps.

839

7. NORTON, T. S. AND DRYER, F. L.: An Experi-

8.

9.

10.
11.

12.

13.

14.

15.

16. 17.

Acknowledgments
This research was supported in part by the Army Research Office and the National Science Foundation. It is a pleasure to acknowledge the helpful discussions with and constructive comments by Dr. K. Brezinsky, Professor F. L. Dryer, Professor W. C. Gardiner, and Dr. T. S. Norton. We also wish to specially thank Professor F. L. Dryer for an advance copy of Ref. 7 and Drs. H-H. Grotheer and Th. Just for their comments on the kinetic scheme. 18.

19.

20.

REFERENCES 1. WESTBROOK, C. K. AND DRYER, F. L.: Combust. Flame 37, 171 (1980). 2. NORTON, T. S. AND DRYER, F. L.: Combust. Sci. Tech. 63, 107 (1989). 3. NORTON,T. S. AND DRYER, F. L.: Int. J. Chem. Kinet. 22, 219 (1990). 4. NORTON, T. S.: The Combustion Chemistry of Simple Alcohol Fuels, Ph.D. Thesis, Princeton University, Princeton, N.J. (1990). 5. NORTON, T. S. AND DRYER, F. L.: Twenty-Third Symposium (International) on Combustion, p. 179, The Combustion Institute, 1991. 6. EGOLFOPOULOS,F. N., DU, D. X. AND LAW, C. K.: Combust. Sci. Tech. 83, 33 (1992). 21.

22.

23.

24. 25.

mental and Modeling Study of Ethanol Oxidation Kinetics in an Atmospheric Pressure Flow Reactor, to appear in Int. J. Chem. Kinet. LAW, C. K.: Twenty-Second Symposium (International) on Combustion, p. 1381, The Combustion Institute, 1989. SMITH, S. R. AND GORDON, A. S.: J. Phys. Chem. 60, 1059 (1956). LIEB, D. F. AND ROBLEE, L. H. S. JR.: Cornbust. Flame 14, 285 (1970). PANDYA,T. P. AND SRIVASTAVA,N. K.: Combust. Sci. Tech. 11, 165 (1975). G(3LDER, (~. L.: Nineteenth Symposium (International) on Combustion, p. 275, The Combustion Institute, 1982. EGOLFOPOULOS, F. N., CHo, P. AND LAW, C. K.: Combust. Flame 76, 375 (1989). ZHU, D. L., ECOLFOPOULOS, F. N. aND LAW, C. K.: Twenty-Second Symposium (International) on Combustion, p. 1537, The Combustion Institute, 1989. DIXON-LEWIS, G.: Twenty-Third Symposium (International) on Combustion, p. 305, The Combustion Institute, 1991. TIEN, J. H. AND MATALON,M.: Combust. Flame 84, 238 (1991). KEE, R. J., WARNATZ, J. AND MILLER, J. A.: A FORTRAN Computer Code Package for the Evaluation of Gas-Phase Viscosities, Conductivities, and Diffusion Coefficients. Sandia Report SAND83-8209, 1983. KEE, R. J., GRCAR, J. F., SMOOKE, M. D. AND MILLER, J. A.: A Fortran Program for Modeling Steady Laminar One-Dimensional Premixed Flames. Sandia Report SAND85-8240, 1985. GREAR, J. F., KEE, R. J., SMOOr(E, M. D. AND MILLER, J. A.: Twenty-First Symposium (International) on Combustion, p. 1773, The Combustion Institute, 1986. LUTZ, A. E., KEE, R. J. AND MILLER, J. A.: SENKIN: A Fortran Program for Predicting Homogeneous Gas Phase Chemical Kinetics with Sensitivity Analysis. Sandia Report SAND878248, 1987. KEE, R. J., RUPLEY, F. M. AND MILLER, J. A.: Chemkin-II: A Fortran Chemical Kinetics Package for the Analysis of Gas-Phase Chemical Kinetics. Sandia Report SAND89-8009, 1989. KEE, R. J., RUPLEY, F. M. AND MILLER, J. A.: The CHEMKIN Thermodynamic Data Base. Sandia Report SAND87-8215, 1987. BURCAT, A.: Combustion Chemistry (W. C. Gardiner, Ed.), p. 455, Springer-Verlag, NY, 1984. MONCHICK, L. AND MASON, E. A.: J. Chem. Phys. 35, 1676 (1961). PRaUSNITZ, J. M.: Molecular Thermodynamics

840

LAMINAR F L A M E S - - K I N E T I C STRUCTURES of Fluid Phase Equilibria, Prentice-Hall, NJ, 1969. KEYSER, F. L.: J. Phys. Chem. 90, 2994 (1986). BAULCH, D. L., DRYSDALE, D. D., HORNE, D. G. AND LLOYD, A. C.: Evaluated Kinetic Data for High Temperature Reactions, vol. 1 and 2, Butterworths, London (1973). FRENKLACH, M., WANG, a . AND RABINOWlTZ, M. J.: Prog. Energy Combust. Sci. 18, 47 (1992). TSANC, W. AND HAMPSON, R. F.: J. Phys. Chem. Ref. Data 15, 1087 (1986). DACAUT, P., BOEaINEa, J-C AND CATHONNET, M.: Combust. Sci. Tech. 77, 127 (1991). NATARAJAN,K. AND BHASKARAN,K. A.: 13th Int. Shock Tube Symposium, Niagara Falls, p. 834, 1981. COOKE, D. F., DODSON, M. G. AND WILLIAMS, A.: Combust. Flame 16, 233 (1971). BOWMAN,C. T.: Combust. Flame 25, 343 (1975). GARDINER, W. C.: Personal communications. BHASKARAN,K. A., RAVIK~MAR,R., KARUPPANAN, K. M. AND NATAaAJAN, K.: 1st Specialists' Meeting (International), Vol. 2, p. 278, The Combustion Institute, 1981. BOalSOV, A. A., ZAMANSKII, V. M., KONNOV, A. A., LISYANSKII, V. V., RUSAKOV, S. A. AND SKACHKOV, C. I.: Khimicheskaya Fizika 4, 1543 (1985).

26. 27.

32. 33. 34. 35.

28. 29. 30. 31.

36.

COMMENTS
R. Ramaprabhu, Anna University, India. In your paper, as I see, you have drawn on results from both shock tube and flow reactor experiments, besides your own experiments, to validate your mechanism for ethanol oxidation. I want to know why you have not thought of arriving at global or quasi global schemes, and validate the same against your experimental results. This would help practical designers to arrive at heat release much more easily and confidentally. Besides, it would have been better to study flame structures for these type of fuels to gain more understanding. Author's Reply. It is not the objective of the present study to propose any global and semi-global schemes. We feel that since the detailed mechanism of ethanol oxidation has yet to be established, it is premature to attempt for simplified schemes, even if such attempts are themselves meaningful. A meaningful flame structure study would need for comparison and guidance a comprehensive set of experimental data, which however, does not exist. Horst-Henning Grotheer, DLR Stuttgart, Germany. With regard to the kinetic data a lot of unknowns remain. For the reactions of primary ethanol attack, branching fractions into the three isomeric radicals have never been measured directly and even the total rate coefficients are not precisely enough known, in particular for CzH~OH + H. For the radicals produced and their reactions, rate coefficients are even more unclear. For instance, CH3CHOH + M, or C2H4OH + 02 have not been measured directly. How can you know that CH3CHOH reacts exclusively with O3 and that C~H4OH or C2H~O are only lost by their decomposition?

P. Van Tiggelen, Universite Catholique de Louvain, Belgium. Although the flame burning velocity
is not very sensitive to the detailed mechanism, it should be pointed out that flame structure studies can cast some light on specific elementary steps provided that the experimental conditions are chosen accordingly (equivalence ratio, fuel, and so on).

Author's Reply. In the paper we have mentioned the need for further research on the branching ratios and pyrolysis kinetics, as you have also emphasized. We have also cautioned on the extent of potential applicability of the present kinetic model. More fundamental kinetic studies are obviously needed, and a study of the present nature can provide useful guidelines for such further investigations. Although space limitations does not permit us to provide a complete listing of the rate coefficients, the paper is self-contained in that all these information are either stated or referenced.

Michael Tanoff, Chalmers University of Technology, Sweden. The Combustion Chemistry Group at
Chalmers has been probing the structure of lowpressure, laminar premixed, fiat ethanol/oxygen/ argon flames, using continuous microprobe sampiing mass spectrometry. We have observed discrepancies, identical to those that you report from

Author's Reply. We completely agree with the usefulness and essential importance of flame structure studies.

STUDY ON ETHANOL OXIDATION KINETICS your flow reactor studies, between measured and computed (using detailed chemical mechanisms applied to fiat flame burner simulations) ethanol consumption rates. Specifically, the modeling predicts a faster rate of ethanol consumption than observed in the flame reactor. Although the disagreement may be explained, partially, by uncertainties in the flame's temperature profile, we are anxious to include your suggestions for the CH3 + HOz rates, and to observe their effect on the computed ethanol concentration profiles.

841

present data and all your earlier results. If correct, what are the implications for modelling?

David Smith, British Gas plc, U.K. Various people have speculated that your burning velocities may be high, possibly by 10%. This would affect the

Author's Reply. We are well aware of some recent theoretical studies which indicate that the counterflow technique could yield 5-10% overestimates of the laminar burning velocity. In fact, we have specifically mentioned this point in the paper, and have been trying to verify it experimentally. On the other hand, one should recognize that a 5 10% inaccuracy amounts to only a few cm/s in the measured burning velocity, which is frequently within the experimental uncertainty anyway. As such, it is not appropriate to attempt to extract kinetic information based on differences which are within the range of these inaccuracies and/or uncertainties. We have followed this guideline in our studies.

You might also like