Anda di halaman 1dari 57

Proy. EnergyCombust. Sci. 1984, Vol. 10, pp. 1-57. Printed in Great Britain.

0360-1285/8450.00+ "50 Pergamon Press Ltd.

CHEMICAL KINETIC MODELING OF HYDROCARBON COMBUSTION


CHARLES K. WESTBROOK* a n d FREDERICK L. DRYERt *Lawrence Livermore National Laboratory, University of California, Livermore, California 94550, U.S.A. tDepartment of Mechanical and Aerospace Engineering, Princeton University, Princeton, New Jersey 08544, U.S.A. Abstract--Chemical kinetic modeling of high temperature hydrocarbon oxidation in combustion is reviewed. First, reaction mechanisms for specific fuels are discussed, with emphasis on the hierarchical structure of reaction mechanisms for complex fuels. The concept of a comprehensive mechanism is developed, requiring model validation by comparison with data from a wide range of experimental regimes. Fuels of increasing complexity from hydrogen to n-butane are described in detail, and further extensions of the general approach to other fuels are discussed. Kinetic modification to fuel oxidation kinetics is considered, including both inhibition and promotion of combustion. Simplifiedkinetic models are then described by comparing their features with those of detailed kinetic models. Finally, application of kinetic models to study real combustion systems are presented, beginning with purely kinetic-thermodynamic applications, in which transport effects such as diffusion of heat and mass can be neglected, such as shock tubes, detonations, plug flow reactors, and stirred reactors. Laminar flames and the coupling between diffusive transport and chemical kinetics are then described, together with applications of laminar flame models to practlcal combustion problems.

1. I N T R O D U C T I O N

In recent years, chemical kinetic modeling has become an important tool in the analysis of combustion systems. Availability of large amounts of elementary kinetic data, improved techniques for estimating specific reaction rates, development of efficient "stiff equation" solution techniques, and continual growth in the size, speed, and availability of large computers have contributed to the increasing application of detailed chemical kinetic modeling. For at least twenty years this approach has been employed in the simulation of controlled laboratory experiments. More recently, kinetics models have become useful in the analysis of practical energy conversion systems such as internal combustion engines. The influences of kinetic factors on other combustion fields such as the assessment of safety factors involved with large scale storage, transportation and use of liquid and gaseous fuels are only beginning to receive attention. Other important kinetic problems including particularly the oxidation of practical fuels in turbulent flows also need a great deal of examination. Although many fuel types are encountered in combustion environments, hydrocarbons comprise the vast majority. In this review we will devote most of our attention to the combustion of hydrocarbon fuels, but it should be pointed out that the same types of analyses can be applied to other fuels such as ammonia, hydrazine, carbon disulfide, and many others based on N - H , C-S, N - C - H , or other combinations rather than C - H systems. We will only briefly discuss the kinetics of the formation of chemical pollutants, a subject which has been treated in detail elsewhere.i -3 Detailed reviews of hydrocarbon oxidation have appeared quite recently,4-7 emphasizing those processes which are dominant at temperatures below about 700 K. Here we will deal primarily with
JP~CS IO:I-A

higher temperature conditions encountered in flames and explosions (T >/1000 K). High temperature combustion studies are complicated by the fact that typical time scales are very short, often of the order of microseconds. As a result, spatial scales are also very small, making experimental studies very difficult. On the other hand, high temperature reaction mechanisms can be conceptually simpler than those for combustion below 700 K. We will first discuss in detail the kinetic mechanisms which are used to describe the combustion of hydrocarbon fuels. Applications of these kinetic models to the analysis of selected types of problems will then be examined.
2. P R O B L E M F O R M U L A T I O N

The general mathematical formulation of the problem of chemically reactive flow systems s - l l consists of equations for conservation of mass, momentum, energy, and concentration of each chemical species, together with equation of state and other thermodynamic relationships. Chemical kinetics provides the coupling among the various chemical species concentrations, and with the energy equation through the heat of reaction. In many combustion problems the kinetics terms determine the characteristic space and time scales over which the equations must be solved. When spatial transport effects can be neglected, the conservation equations become a coupled set of ordinary differential equations (ODE's) for the species concentrations and the energy (or temperature) with time as the independent variable. If transport terms must be considered, then the equations are coupled partial differential equations involving derivatives with respect to both time and space. Examples of both types of systems will be presented in the second half of this paper.

C.K. WESTBROOK and F. L. DRYER to express the reaction rate in terms of other functions or in tabular form. Few reactions have received enough experimental attention to determine completely their behavior for the temperature range of concern here (1000K ~< T ~< 2500K). The use of transition-state theory ls'~9 and other theoretical methods can provide valuable estimates of elementary reaction rates, 2-22 and modeling studies can also provide rate estimates, ~4 particularly for reactions which are very difficult to observe directly. A number of important compilations and discussions of elementary rate data for reactions involved in hydrocarbon oxidation have appeared, 23 35 consisting of critical evaluations of experimental, theoretical, and modeling values for specific reaction rates. In the next section of this paper we will discuss many of these elementary reactions. An extensive list of these reactions has been collected into Appendix I, together with rate expressions and references indicating the source of the rate expression. These are not intended to be definitive evaluations of these rates but rather representative values which might serve as estimates with which to begin assembling a reaction mechanism. The critical evaluations cited above contain extensive references for the reactions being surveyed. However, reaction rate evaluation is a difficult and time-consuming process 36 which has not been carried out for many of the reactions which are important in hydrocarbon and other combustion reaction mechanisms. Direct experimental evaluations of reaction rates are the fundamental sources for the specific coefficients for use in eq. (1). Realistic kinetic models must adhere strictly to those rate expressions which have been measured experimentally, within their stated uncertainty limits. Occasionally, computed model results will be unable to reproduce observed data unless one or more rate expressions are assigned values which disagree dramatically with known experimental values. Invariably this indicates that the assumed reaction mechanism is incomplete or that other parts of the mechanism or physical model are incorrect. An illustration of this point is given in Fig. 1, using experimental data for CO oxidation in a turbulent flow reactor. 37 Measured values for temperature and concentrations of CO and CO2, indicated by the open circles, are plotted as functions of axial position from the inlet section of the reactor. An initial mechanism was postulated, 38 including reactions involving CO, CO2, O, OH, H, H2, and H20. If all of the reactions are assigned rates consistent with experimentally determined values, the best possible agreement between computed and measured results is shown by the upper set of curves. In order to force the model to agree with the observations, the rate of the key reaction C O + O H = CO 2 + H has to be reduced by a factor of ten below its known value. The magnitude of this disagreement indicates

The computational requirements for solving these differential equations depend very strongly on the details of the reaction mechanism. For problems without transport, one differential equation must be solved for each of N chemical species, together with the energy (or temperature) equation, producing (N + 1) equations. Both the computer storage requirements and the CPU time needed increase roughly as ( N + I ) z. When transport effects are included, the same ( N + 1) equations must then be solved for each spatial zone, in addition to equations for conservation of mass and each component of momentum. For many typical applications, 30-40 spatial zones are needed in each dimension, so that two-dimensional problems might contain 1000 zones, with 30,000 zones in three dimensions. Current reaction mechanisms for oxidation of methane or methanol, to be discussed later, contain 25-30 chemical species, so the use of these fuels could require the solution of approximately 30, 1000, 35,000, and 1,000,000 coupled differential equations for zero-dimensional (i.e. no transport), one-dimensional, two-dimensional, and three-dimensional models, respectively. Furthermore, these equations must be solved at each time step in the numerical solution of the combustion problem. The kinetics equations themselves form a set of coupled rate equations describing the reactions between chemical species. Each equation has its own characteristic time scale. The well known "stiffness" problem 12'~3 results when widely disparate time scales occur within a single problem. During the early development of solution methods for kinetics equations, these stiffness difficulties caused severe problems, but there are now many convenient techniques available ~4 for solving such equations. General computer software packages ~5'16 implementing these methods are in common use. 2.1. Reaction Rates For modeling chemical kinetics, it is necessary to adopt a uniform way of expressing the variation of reaction rates with temperature. Conventionally these are expressed in modified Arrhenius form
k = AT"exp(-E~/RT)

(eq. 1)

where the rate k depends on the temperature T and an activation energy E,, with a pre-exponential collision frequency factor A. In many cases all of the temperature dependence can adequately be incorporated into the single exponential term (i.e. with n = 0), since most binary elementary reactions exhibit classical Arrhenius behavior over modest ranges of temperature. However, over temperature ranges encountered in combustion, significant non-Arrhenius behavior can be exhibited, 17 and additional variation of the rate coefficient with temperature must be included. Most often this additional variation is taken into account by finding an appropriate value of n such that the resulting expression accurately fits the available experimental data. It is also possible but less common

Modeling of hydrocarbon combustion


o
0.010 o o o 0.010
o o o

1120 1100

ooO ooG o o0
o o o o o o o o

(b)

[ o o

O00~O0000G
--

1080 1060 1040

8
0.005

o
o

o o

0.005

o i o oo

Oo

[
1120 1100

I
0.012

~ .....

F
0.012 i

0.010 0.010 0.008 1080 1060 1040 20 40 Position 60 80

~ 0.00 6 0.00 4 0.002

0.008 0-006 i
0,004 0.002

20

40 Position

60

80

20

40 Position

60

80

FJ6.1. Comparisons betweenexperimentalflow reactor data (open circles)and computed values(continuous curves), showing CO and CO 2 mole fractions and temperature as functions of position. The top curves show results without HO2 chemistry, the bottom curves show results with HO 2 chemistry.

that the initial mechanism is inadequate. Subsequent inclusion of reactions with HO 2 and H202 dramatically improves the model predictions. The lower set of curves in Fig. 1 agree well with the data, using the more complete mechanism and rate expressions which are completely consistent with experimental determinations. Another example of the way in which problems with model predictions necessitate major revisions in reaction mechanisms is provided by the problem of CH 3 radical recombination reactions in methane oxidation, discussed in detail later in this paper. Non-Arrhenius temperature dependence can generally be treated adequately through the T" term in eq. (1). However, the pressure dependence of some elementary reactions has been somewhat more difficult to prescribe conveniently. Many unimolecular decomposition reactions and their associated recombination reactions exhibit significant pressure dependence in some experimental regimes. Other apparently bimolecular reactions actually proceed through an adduct state and display a dependence on pressure as well. Although a great deal of progress has been made in recent years in both understanding and predicting falloff behavior of important combustion reactions,39 42 further improvements in dealing with pressure dependence for modeling purposes are still needed.
3. REACTION MECHANISMS

Generally speaking, the reaction mechanism provides a description of the elementary steps which occur during the conversion of the fuel and oxidizer to final products. The term "elementary" is used

advisedly, since any mechanism is really an approximation in which all intermediate states with characteristic lifetimes too small to be resolved or to affect the observable features of the combustion are usually neglected. Often the mechanism to be used can depend on the combustion environment or on the types of results which are needed from the model. For example, ionized species or vibrationally excited states of molecules must be considered in some applications, while they can be completely neglected in other models. Further illustrations of these points will be presented below. In practical terms, the mechanism consists of all the chemical species which affect a given combustion event, together with the elementary reactions among those species. At first glance this might seem to require an enormous number of reactions. For N species, there would be N : reactant pairs, and for each pair there could be a number of possible products. Fortunately such a situation does not actually prevail. Many reactions which are mathematically possible either do not occur at all chemically or with rates which are vanishingly small. Therefore the construction of a realistic reaction mechanism involves principally the identification of those reactions which actually occur and are rapid enough to have an impact on the overall progress of the combustion event. The combustion of a hydrocarbon fuel consists primarily of the sequential fragmentation of the initial fuel molecule into smaller intermediate species which are ultimately converted to final products, usually dominated by H 2 0 and CO z. In many cases these intermediate species can be fuels themselves. For example, ethylene (C2H4) is an important inter-

C.K. WESTBROOK and F. L. DRYER

C3 species

I
CH4 C2H 6 - C2H 4 - C2H 2 C2H5OH

c.2o
1
CO

c.3o.

H2 - 0 2

FIG. 2. Hierarchical structure and overall interrelationships between oxidation mechanisms for simple hydrocarbon fuels.

mediate in the combustion of propane (C3H8) and other higher hydrocarbons, but ethylene can also be a primary fuel. Carbon monoxide (CO) and hydrogen (H2) are common species which are observed during the oxidation of all hydrocarbons, and the same radical species H, O, OH, HO 2, HCO, and others are common to all hydrocarbon combustion. These observations can be used to great advantage in establishing kinetic mechanisms for the reaction of complex practical fuels. A mechanism can be developed systematically, beginning with the simplest species and reactions which are common subelements in the combustion of more complex species, and sequentially constructed by incorporating new species and reactions in order of increasing complexity. At each level the newly added portions of the mechanism must be tested and validated by thorough comparison between numerically predicted and experimentally observed results. However, because of the sequential ordering, only those features which have been added need be examined closely. This hierarchical structure is summarized in Fig. 2. The validation of each level in a reaction mechanism is itself a complicated and involved process. It is not sufficient to test a mechanism by comparison with a single experiment, because different elementary reactions can be dominant in different experimental re-

gimes. For example, with many hydrocarbon systems, reactions between fuel molecules and H atoms are dominant in fuel-rich conditions but much less important in fuel-lean conditions where reactions between fuel and O or OH radicals are most important. Some reactions are negligible except under high temperature shock tube conditions, while others become completely unimportant for temperatures above 1000 K. Unless a mechanism is to be used only for a restricted class of applications under a specific set of system parameters, it should be validated by comparison with experimental data over wide ranges of physical conditions. Some of the experimental configurations which generate kinetics information useful in establishing combustion reaction mechanisms are summarized in Fig. 3, together with the parameter ranges most often covered by each type of experiment. We term reaction mechanisms which have been tested and validated for these wide parameter ranges "comprehensive reaction mechanisms".43-46 This type of mechanism represents a numerical modeling tool of exceptional generality. Unlike a reaction mechanism which has been developed only for a specific regime (e.g. for shock tube conditions), a comprehensive mechanism can be applied in a predictive fashion with some confidence that the important features of the fuel consumption process have not been omitted. For

Shock tube Plug flow reactor Flames

P high atmospheric low

T ~1300K 850-1300K 800-2500K

Dilution yes yes no

limits
none

Transport effects
no

none flammability limits

no yes

FIG. 3. Summary of some of the types of experimental regimes which provide high temperature kinetic rate data for detailed reaction mechanisms and for which experimental data exist for model validation. Typical parameter ranges are shown, together with an indication of whether or not transport effects are important in simulations.

Modeling of hydrocarbon combustion example, comprehensive mechanisms developed for oxidation of methanol43 and ethylene45 have been used to predict laminar flame47 and detonation4s 50 properties with the same fuels, with results in good agreement with experimental observations. Another essential element in mechanistic studies is the description of the thermochemical properties of all of the chemical species involved. This information, including heats of formation, temperature dependent specific heat data, and specific entropy and enthalpy, is generally as important a part of a kinetic model as the elementary reaction rate data. Many of these data are available in convenient tabulations.5~,52 However, thermochemical data for some minor intermediate and radical species are not well known, resulting in significant uncertainties in computed rates of important reactions. For example, the heat of formation of ketene is given by Benson ~8 as - 14.6 kcal/mol and as 4.0kcal/mol by Bahn. 52 The 1971 revision of the JANAF Thermochemical Tables 5~ revised the heat of formation of the formyl radical from - 2 . 9 to 10.4 kcal/mol, while Benson gives a value of 7.2 kcal/mol. These differences reflect a steady improvement in the precision with which such data are known, but they also can cause dilemmas when one is assembling a reaction mechanism. Other species for which large changes have been made recently or for which significant uncertainties still remain in the heat of formation include C2H, NH, and NH2. Since errors in the heat of formation are translated into errors in the activation energy of equilibrium constants and rates of reverse reactions, poorly known thermochemical data can be a serious problem in the modeling of some combustion systems. H+O 2 = O+OH O+H 2 = H+OH H 2 + O H = H20 + H O+H20 = OH+OH. The principal termination reactions include H+H+M O+O+M O+H+M H+OH+M = H2+M = O2+M = OH+M = HEO+M.

5 (1) (2) (3) (4)

(5) (6) (7) (8)

where M refers to any available third body species, required to conserve both momentum and energy in this type of reaction. The hydroperoxyl radical HO2, originally proposed to explain the explosion limit behavior of H 2 0 2 mixtures, 5'77 is formed primarily by H+O2+M = HO2+M (9)

and consumed by reactions with various radicals, including HO2+H = H2+O 2 HOz+H = OH+OH HO2+H = H20+O HO2 + O H = H 2 0 + O 2 HO2+O = O2+OH. (10) (11) (12) (13) (14)

The reaction of HO 2 with itself produces hydrogen peroxide (H20 2) HO 2 + HO 2 = H 2 0 2 + 02 (15)

which is consumed by reactions with radicals and by thermal decomposition H 2 0 2 + OH = H20 + HO2 H202+H = H20+OH H202+H = HO2+H 2 H202+M = OH+OH+M. (16) (17) (18) (19)

3.1. Specific Mechanisms


In this section we will summarize the important features in the reaction mechanisms for the combustion of hydrocarbon fuels. The hierarchical structure noted above will be emphasized by first discussing the H2-O2 submechanism, followed in order by CO, CH20, and others in order of increasing complexity. 3.1.1. H 2 - O 2 Historically, the hydrogen-oxygen system was the first combustion mechanism to be developed for a practical fuel. Analytical solutions were initially obtained for the rate equations. 53-57 This system was also used in the earliest detailed numerical models of fuel oxidation in shock tubes, stirred reactors, and laminar flames. 5s -63 Because the reaction mechanism required for hydrogen oxidation is much smaller than for hydrocarbon oxidation, H2-O 2 mechanisms have been used in the first models for multidimensional fluid mechanics-chemical kinetics,64'65 sensitivity analysis,66,67 detonations,6S 73 and ignition.74-76 The important reactions for H 2 oxidation and their rates have been extensively studied and documented. The radical species pool is evolved among OH, O, and H by the reactions

An additional reaction which is occasionally encountered involving HO2 is O + OH + M = HO 2 + M (20)

but there is no indication that it has any role in the H2-O 2 mechanism and there are no reliable data on its rate. 23 Rates of HO 2 reactions at combustion temperatures have been reviewed, 35"78,79 while others have recently been redetermined 8 84 at lower temperatures. In the high temperature combustion of hydrogen and hydrocarbons, reaction (1) is the single most important chain branching reaction, consuming one H atom and producing two radical species O and OH. Any type of kinetic perturbation which increases the production of H atoms will accelerate the overall rate of combustion by increasing the net amount of chain branching from reaction (1). Conversely, processes which reduce the H atom population and reactions which compete with reaction (1) for H atoms will tend

6
I 8 100 80 60 40
c .0

C.K. WESTBROOKand F. L. DRYER


I I

Stoiehiometric fuel-air mixtures

_~
8

.E

~*

20

I
0.1

I
1 Pressure (atm)

1
10

FlG. 4. Variation in laminar burning velocity with pressure for stoichiometric fuel-air mixtures.

to inhibit the c o m b u s t i o n . There are several good illustrations of this behavior. F o r example, reaction (9) competes directly with reaction (1) for H atoms, b u t the rate of the third order reaction (9) is m u c h more pressure d e p e n d e n t t h a n t h a t of reaction (1). The variation of l a m i n a r flame speed in C H 4 - a i r , 47'85 C H a O H - a i r .7 a n d C z H 4 - a i r s6 flames c a n be linked directly to this c o m p e t i t i o n between reactions (1) and (9) and is illustrated in Fig. 4. In each case there is a very gradual decrease in l a m i n a r flame speed S , as the pressure is increased, for pressures below atmospheric. Then, between 1 a n d 5 a t m . pressure, the decrease in Su with increasing pressure becomes more p r o n o u n c e d . The reason for this change in b e h a v i o r is t h a t reaction (9) begins to compete effectively with reaction (1) for H a t o m s at pressures above 1 arm. Below a t m o s p h e r i c pressure, reaction (9) does n o t compete effectively a n d the effect of pressure on the flame speed is very small. Because the e x p l a n a t i o n for this p h e n o m e n o n is i n d e p e n d e n t of the specific fuel being used, any h y d r o c a r b o n fuel will exhibit the same non-linear d e p e n d e n c e of b u r n i n g velocity on pressure. F u r t h e r m o r e , the same a r g u m e n t can be extended to n o n - h y d r o c a r b o n fuels such as a m m o n i a (NH3) , hydrazine ( N z H 4 ) a n d other N - H species for

which reaction (1) still provides the majority of the chain branching. In a d d i t i o n to their different d e p e n d e n c e o n pressure, the rates of reactions (1) a n d (9) have distinctly different activation energies. While E 1 is approximately 16.8 kcal/mol, 23 E 9 has a small negative value, so k 9 actually decreases with increasing temperature. Therefore reaction (1) will d o m i n a t e at high t e m p e r a t u r e a n d low pressure, while reaction (9) will be n'iore rapid at lower t e m p e r a t u r e a n d elevated pressure. These regimes will overlap in m a n y cases, a n d there is usually a region in which b o t h reactions have c o m p a r a b l e rates. While reaction (1) provides chain branching, reaction (9) produces chain termination u n d e r some conditions, while u n d e r other conditions it is p a r t of a chain p r o p a g a t i o n p a t h consisting in large part of reactions (9) and (11) or reactions (9), (15), a n d (19). A n o t h e r example of the way in which c o m p e t i t i o n for H a t o m s affects the rate of c o m b u s t i o n is provided by the o b s e r v a t i o n t h a t the addition of small a m o u n t s of various chemicals to flammable mixtures can dramatically reduce the overall rate of combustion. As we will discuss later in greater detail, the addition of m a n y h a l o g e n a t e d species inhibits the c o m b u s t i o n process by r e m o v i n g H a t o m s from the radical pool, forming H 2 by a catalytic cycle of reactions. The r e m o v a l of H a t o m s m e a n s that they are then unavailable for reaction with 0 2 t h r o u g h reaction (1), thereby reducing the rate of chain branching and slowing the overall rate of c o m b u s t i o n . Rates of reaction between H a t o m s and m a n y h y d r o c a r b o n species are considerably larger t h a n the rate of reaction (1) at the t e m p e r a t u r e s which are encountered in flames. These rate expressions are illustrated in Table 1 for a variety of simple hydrocarbons, with all the rates evaluated at 1000K. The rates of reaction between H a t o m s and the saturated hydroc a r b o n s as well as methanol, formaldehyde, and ethylene are larger t h a n the rate of reaction (1). Therefore, these reactions will compete effectively for H a t o m s with reaction (1), reducing the chain branching rate. This explains why m a n y h y d r o c a r b o n species act as inhibitors for the H 2 - O 2 system, sT'ss Even

TABLE 1. Rates of reactions between H atoms and 02 or hydrocarbon molecules in cm ~ mol sec kcal units Reaction H + O 2 -~ O + O H H + CH 4 -* C H 3 + H 2 H + C 2 H 6 - ' C2Hs +H2 H + C 2 H 4 -~ C2H3 + H 2 H+CH20 ~ HCO+H 2 H + CH3OH --, CHzOH + H 2 H +CH3OH ~ CH a + H 2 0 H + C 3 H 8 ~ iC3H~ + H 2 H + C 3 H s ~ nC3H 7+H2 H+C2H2 ~ C2H+H2 H + C4 H 10 ~ nCcH9 + H2 H q~C4Hlo --* sC4H 9 + H 2 Rate expression 5.13 1016T o.816exp ( - 16507/RT) 2.24 104T 3 exp ( - 8750/RT) 5.37 x 10aT 3"5exp ( - 5200/R T) 1.50 x 107 T 2 exp ( - 6000/R T) 3.30 x 1014 exp ( - 10500/RT) 3.00 1013 exp ( - 7000/RT) 5.25 1012 exp ( - 5340/RT) 1.46 x 107T 2 exp ( - 5000/RT) 9.38 107 T 2 exp ( - 7700/R T) 2.00 101*exp ( - 19000/RT) 1.30 x 1014 exp ( - 9700/RT) 2.00 1014 exp ( - 8300/RT) Rate at 1000 K 4.5 x 101 2.7 101 t 1.2 x 1012 7.3 1011 1.7 1012 8.9 x 101 l 3.6 l0 Is 1.2 1012 2.0 x 1012 1.4 x 101 9.9 1011 3.1 x 1012

Modeling of hydrocarbon combustion though the hydrocarbon provides additional fuel to the system and may in fact increase the final mixture temperature, the interference with reaction (1) results in an overall inhibition. At elevated pressures (P/> 20 atm.) and at relatively low temperatures ( T - 1000K), reaction (9) will dominate over reaction (1). In such environments the sequence of reactions (9), (15), and (19)provides chain propagation (i.e. two H atoms consumed, and two OH radicals produced). In contrast, the sequence of reaction (9) followed by Fuel + HO 2 = Radical + H20 2 and then reaction (19) results in chain branching (one H atom yields two OH radicals). Under these conditions increased fuel concentration will therefore result in an accelerated overall rate of reaction, while at lower pressures an increased fuel concentration will provide an inhibiting effect by competing with reaction (1). Reaction (2) provides additional chain branching in H 2 - O : systems, although its importance is not as great as that of reaction (1). Reaction (3) is responsible for the majority of H 2 consumption under normal circumstances, so when the OH level is low, H 2 oxidation will be slow. The reverse of reaction (10) provides a mechanism for initiation of H2-O 2 mixtures which accuratel)~ reproduces experimental ignition delay measurements in shock tube and detonation conditions. 89 Reaction (10)dominates over reaction (21), H2+O 2 =OH+OH (21)

occasionally considered in combustion environments since it is an oxidizer which contributes much more chemical potential energy to a given fuel-oxidizer mixture than does ordinary oxygen 02, and its use provides significantly higher product temperatures and pressures than does 02 . Mixtures of 0 3 and 02 will support a laminar "flame" consisting of a narrow decomposition front which will propagate at a fairly high speed. 92 The reaction mechanism for the ozone decomposition flame is the simplest one which has all of the requisite features of a truly detailed mechanism, consisting of only three species O, O 2, and 0 3, with three reactions, reaction (6) together with O3q-M = O 2 q - O + M 0 3 + O = 02 + 02. (22) (23)

Direct coupling between the 0 3 system and H 2 oxidation can be provided by reactions including H+O 3 = OH+O2 O H + O 3 = H O 2 + O 2. (24) (25)

Because of its simplicity, the ozone decomposition flame has been used frequently93-1o2 as a convenient vehicle for testing laminar flame models. The ozone mechanism is of minor practical importance in combustion environments, but its application to model validation can make it valuable. 3.1.3. CO The oxidation mechanism for carbon monoxide is very simple, consisting of CO+O+M = CO2+M (26) (27) C O + O 2 = CO2 +O.

suggested in the early modeling literature 5s as a possible alternative initiation mechanism. Recent evaluations 23 conclude that reaction (21) does not occur as written or has a very small rate. In their review of advances in kinetic rate data through 1979, Gardiner and Olson 33 state that all of the important mechanistic steps and rate expressions for the H2 Oz mechanism are now well known. They point out that important work still needs to be done, particularly in determining chaperon efficiencies for different third bodies in reaction (9). The importance of these effects can sometimes be very great, as we will demonstrate in the next section dealing with CO oxidation. Dixon-Lewis and his collaborators 9'9~ have devoted a great deal of attention to the problem of third body or chaperon efficiencies. As a general rule, the efficiency of each species which is present in quantities larger than about 5 ~ should be considered. Water molecules are particularly good chaperons, with efficiencies which can be 10-50 times larger than those of Ar or He. The chaperon efficiency increases with the number of degrees of freedom available to share collisional energy. Atomic species like argon or helium are comparatively inefficient, followed by diatomic species such as N 2 and O2, followed in turn by polyatomic species. 3.1.2. 03 An interesting submechanism of the H2-O 2 system is the ozone decomposition mechanism. Ozone is

The rates of both these CO oxidation reactions are quite small at combustion temperatures, so that CO oxidation in hydrogen-free environments is very slow. However, if H atoms are present, even in amounts as small as 20ppm H20,13 the CO mechanism becomes strongly coupled to that of H2-O2, primarily through CO+OH = CO2+H CO + HO 2 = CO 2 + OH. (28) (29)

Reaction (29) is rarely as important as reaction (28), 14'15 although at very high pressures or in the initial stages of hydrocarbon oxidation, the high HO 2 concentrations can make it competitive. All hydrocarbon oxidation eventually involves H2 and CO oxidation kinetics, and most of the CO2 that is produced results from reaction (28). The coupling between the CO and the H2 02 submechanisms is further complicated by the unusually large chaperon efficiency of water molecules for some reactions. For example, Dryer and Glassman 37 found that, at temperatures near i000 K and at atmospheric pressure, the rate of CO oxidation in a turbulent flow reactor depended on the concentrations of CO, 02, and H20, even though the water concentration was quite small (i.e. 2.5~). Modeling analysis 16 found that the role of H 2 0 as a third body in the recombina-

C. K , WESTBROOK a n d F. L. DRYER

TABLE2. Rates of reactions between OH radicals and selected other species in c m 3 mol sec kcal units Reaction CO+OH ~ CO2+H H2+OH ~ H20+H CH20 + O H ~ H C O + H 2 0 CHa + OH ~ CH 3 + H20 CH3OH + O H ~ CH2OH +HzO C2H6 + O H --* C2H5 + H20 C2H4 + OH ~ C2H3 + H20 CEHa + OH --, C H 3 + C H 2 0 C2H2 + OH -+ CH2CO + H C2H 2 + OH ~ C2H + H20 C2H 2 + OH --~ C H 3 + CO C3H 8 + OH ~ iC3H 7 + H20 C3H s + O H ~ nCaH 7 + H 2 0 Rate expression 1.50 107 T 1"3exp ( + 765/RT) 2.20 x 1013 exp ( - 5146/RT) 7.50 x 1012 exp ( - 170/RT) 3.50 x 103T 3's exp ( - 2000/RT) 4.00 1012 exp ( - 2000/RT) 1.12 1013 exp ( - 2450/RT) 4.80 x 1012 exp ( - 1230/RT) 2.00 x 1012 exp ( - 960/RT) 3.20 x 1011 exp ( - 200/RT) 6.30 x 1012exp ( - 7000/RT) 1.20 x 1012 exp ( - 500/RT) 4.80 I08T l'a exp ( - 850/RT) 5.76 10STl4exp ( - 850/RT) Rate at 1000 K 1.8 x 101~ 1.7 1012 6.9 1012 2.2 x 1012 1.5 1012 3.3 x 1012 2.6 x 1012 1.2 x 1012 2.9 101 t 1.9 x 1011 9.3 1011 5.0 x 1012 6.0 1012

tion reaction (9) is especially important. The chaperon efficiency for H 2 0 in this reaction was varied, showing that the sensitivity of the computed results to this one kinetic parameter is large. The best agreement was obtained by setting the third body efficiency of water molecules in reaction (9) equal to that determined from shock tube experiments by Getzinger and Schott. 1o 7 The importance of reaction (28) in all of hydrocarbon oxidation cannot be overemphasized. Like reaction (1), it plays a dominant role in the combustion of all hydrocarbon fuels. However, reaction (28) is an exceedingly complex "elementary" reaction, actually proceeding through a four-atom activated c o m p l e x ) v'x8 As a result, the reaction rate exhibits some pressure dependence and variation with the relative efficiency of different third bodies.~ 09 Because reaction (28) consumes nearly all of the C O which reacts to produce CO2, the rate of C O oxidation depends very much on the availability of O H radicals. It was observed earlier that the presence of hydrocarbon species inhibited chain branching because the rate of reaction (1) is considerably less than the rate of reaction between H atoms and hydrocarbons. In a very similar way the rate of reaction (28) is considerably less than the rate of reactions between O H radicals and typical hydrocarbon species, as shown in Table 2. As a result, the presence of most hydrocarbon species, even in quite small amounts, will effectively inhibit the oxidation of CO. During
I I I

the oxidation of hydrocarbons, C O is produced in substantial amounts, but the subsequent oxidation of C O to C O 2 is usually retarded until after the original hydrocarbon and the fragment intermediate hydrocarbon species have all been consumed. Only then does the O H concentration rise to higher levels, rapidly converting the C O to C O 2. This can be illustrated in Fig. 5, using computed results from three hydrocarbon air laminar flames. 86 The C O concentration grows steadily as fuel is consumed, but very little C O is consumed until all of the fuel has disappeared, whereupon the O H concentration (not shown) rises sharply and reaction (28) rapidly consumes the CO. 3.1.4. CH20 Formaldehyde is an intermediate in the oxidation of most hydrocarbon fuels. However, oxidation and pyrolysis studies with formaldehyde as the original fuel have been relatively rare, primarily because the preparation of a combustible mixture of formaldehyde and oxidizer or diluent is relatively difficult. Kinetic studies of formaldehyde pyrolysis and oxidation have appeared recently, 11 l la primarily in simulations of shock tube experiments. There is some debate concerning the thermal dissociation products of C H 2 0 . The two possibilities are CH20 + M = HCO + H + M C H 2 0 + M = C O + H 2 + M. (30) (30a)

1.0 0.8 __ o= 0.6 ----CH4 ///4 .J'/-

///"

//

!
0.4 E

o 0.2 O.C -1.0

....

./%z..

-0.5

0.0
Relative flame position (mm)

0.5

1.0

FIG. 5. Computed temperature and normalized fuel and CO profiles in laminar stoichiometric fuel-air flames at atmospheric pressure.

Most studies 11'11L~ 13 support the first channel, but other work ~~4,1 is favors the alternate path. The chain branching characteristics of the two paths are vastly different. Reaction (30) produces one H atom directly, and in most environments the formyl radical H C O rapidly decomposes further into C O and H. Thus the net difference between reactions (30) and (30a) is H + H as opposed to H 2. Together with reaction (1), reaction (30) is therefore a very active chain branching sequence in oxidizing environments while reaction (30a) is effectively a termination sequence. With these differences, it would at first appear that the resolution of the dispute should be straightforward. However, all of the mechanisms are strongly dependent on rates of other C H 2 0 reactions and reactions of formyl radicals.

Modeling of hydrocarbon combustion Uncertainties in these rates, particularly the reactions of H C O , have been so large that they have obscured resolution of this issue. It is probable that both paths actually occur at the same time x16 with different rates and different impacts on the chain properties of the overall combustion rate. It appears from modeling analysis of CH4, C H a O H and other flame and shock tube systems 47,4s that when formaldehyde exists only as an intermediate, the thermal decomposition reaction is not important and may in fact proceed in the reverse direction. The reactions between C H 2 0 and radical species are dominant in most situations. Formaldehyde is consumed primarily by reactions with OH, H, and O radicals to produce H C O . CH20 + OH = HCO + H20 CH20 + H = HCO + H 2 C H 2 0 + O = H C O + OH. (31) (32) (33) C H 2 0 + HO2 = H C O + H202.

ceilent example of the way in which an inadequate reaction mechanism can lead to incorrect estimates of rate data. In almost every case, the problem is caused by the omission of significant portions of the complete mechanism. Current C H , oxidation mechanisms in which CH3 recombination reactions and subsequent consumption of C 2 species are correctly included no longer require artificially high rates of reactions between C H 2 0 and radical species. The particular reaction environment being examined determines the relative importance of reactions (31-33). In fuel-rich mixtures, reaction (32) dominates, while in lean and stoichiometric conditions reactions (31) and (33) dominate. Another reaction which is less important but which can contribute to C H 2 0 consumption in some cases is (34)

The most recent evaluations of the rates of these reactions are by Dean and co-workers, lt'11~ who examined both pyrolysis and oxidation experiments in shock tubes, with 0 2 and N 2 0 as oxidizers over a wide range of fuel-oxidizer proportions and initial temperatures. This represents the closest approach yet to a comprehensive mechanism for formaldehyde combustion. Some confusion exists from early CH4 combustion modeling efforts, carried out before the importance of methyl radical recombination had been established. Methyl radicals were assumed to react directly to produce C H 2 0 , but C H 2 0 was not observed in large amounts. Therefore the reactions consuming C H 2 0 were required by these modeling studies to be very rapid. However, recent direct measurements of these C H 2 0 consumption rates have shown that the earlier values were much too high and that the low concentrations of formaldehyde were due instead to relatively low formation rates of formaldehyde. This is an ex-

The formyl radicals produced by reactions of C H 2 0 and other species are in turn consumed in a variety of elementary reactions. The two most important reactions of H C O are HCO+M = H+CO+M (35) (36) H C O + O 2 = C O + H O 2.

Benson and O'Neal 117 discuss reaction (35) in their summary of unimolecular gas phase reactions, pointing out that it is actually second order in most combustion environments. This pair of reactions for a hydrocarbon radical illustrates a pattern that will be repeated for many larger radicals; a reaction with 0 2 competes with a thermal decomposition reaction, and their relative rates determine much of the behavior of the overall mechanism. Typically the activation energy of the decomposition reaction is substantially higher than that of the reaction with 02. F o r the formyl r a d i c a l , E s 5 = 19kcal/mol and E 3 6 = 7.0kcal/mol. Therefore the decomposition reaction will dominate at high temperatures such as those encountered in

TABLE3. Rates of thermal dissociation reactions and reactions with 02 molecules for selected hydrocarbon radical species in cm 3 mol sec cal units Reaction HCO + 02 -~ CO + H O 2 HCO+M ~ CO+H+M CH2OH + O 2 ~ CH20 + HO 2 CH2OH + M ~ CH20 + H + M C2H 5 + 0 2 --"C2H 4 + HO 2 C2H5 + M --~ C2H4 + H + M C2H 3 + O 2 ~ C2H 2 + H O 2 C2H3 + M --* C2H2 + H + M C H 3 + O 2 --* CH30 + O C H 3 + M --* CH2 + H + M nC3H 7 + 0 2 -~ C 3 H 6 + H O nCaH 7 --~ C3H6 + H nC3n 7 ~ C2H 4 + CH 3 iC3H7 + 02 ~ C3H6 + H O 2 iC3H 7 ~ C3H 6 + H iC3H 7 ~ CzH 4 + CH 3
2

Rate expression 3.30 x 1012 exp (-- 7000/R T) 1.45 x 1014 exp ( - 19000/R T) 1.00 x 1012 exp ( - 6000/RT) 2.50 x 101a exp ( - 29000/R T) 1.00 x 1012 exp ( - 5000/R T) 2.00x 10Is exp ( - 30000/RT) 1.00 x 1012 exp ( - 10000/RT) 8.00X 1014 exp ( - 31500/RT) 4.80x 1013 exp ( - 29000/R T) 1.95 x 1016 exp (-91600/RT) 1.00 x 1012 exp ( - 5000/RT) 1.25 x 1014 exp ( - 37000/R T) 9.60 x 101 a exp ( - 31000/RT) 1.00 1012 exp ( - 5000/R T) 6.30 x 1013 exp ( - 36900/RT) 2.00 x 10 l exp ( - 29500/RT)

10

C.K. WESTBROOK and F. L. DRYER 3.1.5. CH4 More modeling work has been devoted to methane oxidation than to all other hydrocarbon fuels combined. The early mechanisms 12 124 have gradually been refined until current models 38'48'85'114'x25 144 can involve nearly one hundred elementary reactions. While methane has often been used in mechanistic studies because of its apparent simplicity, it is ironic that the methane oxidation mechanism exhibits subtleties which are not encountered in the oxidation of many larger fuel molecules. Methane is an exceedingly important practical fuel, constituting approximately 9 0 ~ of the composition of natural gas. It is also a significant byproduct in many industrial processes and is produced during the combustion of most other hydrocarbons. The construction of a complete reaction mechanism for methane combustion is complicated by a feature which will be observed below for other fuels. During the pyrolysis and oxidation of methane, radical recombination reactions produce significant amounts of larger hydrocarbons containing two or more carbon atoms.38,136-138 In particular, the C 2 species ethane, ethylene and acetylene are observed. The subsequent consumption reactions of these C a species must therefore be included in a complete CH 4 mechanism. Since the levels of C 2 species during CH~ combustion are considerably smaller than the CH 4 concentration, further growth of C 2 species to C 3 and C 4 hydrocarbons usually need not be considered. However, for very fuel-rich conditions under which soot production can occur, polymerization reactions forming species up to C4 levels have been proposed. 134 The thermal decomposition of CH 4 has been studied in detail, 1x5,145 147 yielding methyl radicals CH 4 + M = CH 3 + H + M. (40)

many shock tube experiments, while the reaction with 0 2 will dominate at lower temperatures. In Table 3 the rate expressions for analogous reaction pairs involving other hydrocarbon radicals are summarized, illustrating that the pattern noted here for HCO radicals applies to many other radical species as well. Most of the radical decomposition reactions have been written as second order reactions, which may not always be adequate. Still, the trends already noted, particularly the relative values of the activation energies, will follow the pattern shown in Table 3. In each case, the thermal decomposition reaction will be most important for rich mixtures and at high temperatures while the R + O 2 reaction will dominate for lean mixtures and at low temperatures. The boundary separating these two regimes will not always be very well defined, and there is usually a regime in which both reactions are important. The rates of the decomposition reactions increase as the radical species become larger, so that eventually the R + Oz reaction will become relatively unimportant. For example, in laminar flame environments Warnatz 118 showed that the decomposition reactions of propyl and butyl radicals were so fast that all their other reactions, including those with 02 molecules, could be neglected. However, for radicals smaller than propyl both reaction paths must be considered. The effects of these two reaction types on the chain properties of the mechanism are quite different. In nearly every case, the decomposition reaction yields a stable intermediate species and highly reactive H atoms which provide subsequent chain branching through reaction (1). In contrast, the radical-O2 reaction produces the same stable intermediate and HO 2 radicals which are usually much less reactive than H atoms. Therefore the overall chain behavior of the mechanism varies with temperature and equivalence ratio due to the balance between the two principal radical consumption paths. The formyl radical reacts also with other radical species, with rates that are close to collisional, having negligible activation energies. These include HCO+OH = CO+H20 HCO+H = CO+H 2 HCO+O = CO+OH. (37) (38) (39)

Another possible initiation reaction under oxidation conditions has been suggested ~24.~~s CH4+O 2 = CH3+HO 2 (41)

but has been shown 11s to be much less important than reaction (40), at least under shock tube conditions. Hydrogen abstraction from CH4 occurs by several reactions, including CH 4 + H = CH 3 + H 2 CH 4 + OH = CH 3 + H20 CH 4 + O = CH 3 + OH CH4 + HO2 = CH3 + H202. (42) (43) (44) (45)

Since radical concentrations are usually so much smaller than concentrations of the stable reactant and intermediate species, the radical radical reactions are rarely as significant as the radical decomposition or radical-O 2 reactions. Reactions (37-39) actually tend to inhibit the overall rate of combustion by competing with reaction (35) and by reducing the size of the available radical pool. In rich CO oxidation, formyl radical formation by reaction ( - 35) and consumption by reactions (36-39) can be important. 119 In such conditions, reaction ( - 3 5 ) is a termination reaction and its inclusion leads to a slight reduction in computed flame speeds in CO/H2/O z flames diluted by N 2 and Ar.

The rates of the first three of these reactions have been shown149-15a to exhibit substantial non-Arrhenius temperature dependence over the range required for combustion modeling. A major problem in methane combustion concerns the identification of the primary paths for methyl radical consumption. The earliest mechanisms assumed that methyl radicals reacted directly with O2 CH 3 -}-0 2 = HCO + H20 (46a)

Modeling of hydrocarbon combustion while later mechanisms replaced this "overall" reaction with CH 3 + 0 2 = CH20 + OH. (46b) CH 3 + HCO = CH,, + CO CH3 +HO2 = C H 3 0 + O H

11 (54) (55)

Bowman 129 pointed out that reaction (46b) was probably not an elementary reaction, but rather a convenient model for simulating CH 3 oxidation. Computed model results were found to be very sensitive to variations in the rate of reaction (46b), with a rather large rate and relatively low activation energy required in order to reproduce observed combustion rates. Two developments substantially changed this picture of CH3 oxidation. First, the importance of methyl radical recombination CH 3 + C H 3 = c 2 n 6 (47)

while at very high temperatures CH 3 may also dissociate 1s 5 CH 3 + M = CH 2 + H + M. (56)

as a major contributor to CH 3 radical consumption was established. 3s'13s In earlier models this path had been neglected, so in order to reproduce observed rates of CH 3 consumption an artificially high rate had been assigned to reaction (46b). The second major development was the direct measurement of the rate of reaction between methyl radicals and 0 2 molecules.~S2 At 1200 K the observed rate of reaction was at least two orders of magnitude smaller than that computed from commonly used rate expressions. The combination of these two factors effectively eliminated reaction (46b) from consideration in methane oxidation mechanisms. It now appears that the principal products of this reaction are CH 3 + 0 2 = CH30 + O. (46)

When methyl recombination reactions and the subsequent oxidation of the C 2 species are properly included in CH 4 oxidation mechanisms, it is possible to reproduce experimental data for methane oxidation without requiring an artificially large rate for reactions between CH 3 and 02 or for the reactions between formaldehyde and radical species as discussed earlier. This is another example of the way that inadequacies in the reaction mechanism (i.e. omission of CH 3 recombination) can lead to erroneous values for specific reaction rate expressions such as CH 3 + 0 2. It is also interesting to see how the use of sensitivity analysis, to be discussed in detail later, can assist in the development and validation of reaction mechanisms. Brute force sensitivity analysis of early mechanisms showed that computed results depended strongly on the rate used for reaction (46b), leading to an increased interest in that reaction. This attention led to the direct measurement of its rate and a reevaluation of the entire mechanism. Methoxy radicals (CH30) react primarily by means of CH30+M=CH20+H+M CH30 + 02 = CH20 + HO2 CH30+H = CHzO+H 2 (57) (58) (59)

This reaction path was initially suggested by Brabbs and Brokaw 125 and has been confirmed more recently.~ 53,154 The activation energy of reaction (46) is quite large (i.e. ~ 29 kcal/mol) and the rate expression in the Appendix agrees well with the observations at 1200 K. In addition to reaction (47), other methyl recombination reactions include CH 3 + CH 3 = C2H 5 + H CH 3 + CH 3 = C2H 4 + H 2. (48) (49)

At normal combustion temperatures reaction (47) dominates over reactions (48) and (49), but at high temperature shock tube conditions these alternative product distributions can contribute to the overall rate of reaction. Of these reactions, reaction (48) provides a great deal of chain branching, due to the H atom product as well as an additional H atom produced from ethyl radical decomposition. In contrast, reactions (47) and (49) produce relatively stable products. Reactions of methyl radicals include a variety of additional channels, CH 3 q- O = C H 2 0 q- H CH3 + O H = C H 2 0 + H z CH 3 + OH = C H 3 0 -I-H CH3 + C H 2 0 = CH4 + HCO (50) (51) (52) (53)

with the decomposition providing the major fraction. Other reactions between CH30 and radicals such as O and OH probably occur at fairly high rates, but such paths have not yet been incorporated into detailed mechanisms. The competition between reactions (57) and (58) follows the pattern discussed earlier for formyl radicals, with the reaction with O z competing only for very lean mixtures and at lower temperatures (see Table 3). Reaction (57) is generally very rapid, and a comparison between reaction (46b) and the combination of reactions (46) and (57) shows how a relatively low rate for reaction (46) can still result in a rapid consumption of (CH3) radicals. Both paths produce CH20, but the two-step methoxy radical route gives O and H separately while the one-step path produces OH. The radical multiplication, along with the very reactive H atom, in the methoxy radical path provides considerably more chain branching than the single step. Bhaskaran e t al. 126"153 in fact write reaction (46) as CH3q-O 2 = C H 2 0 + O + H which is equivalent to the combination of reactions (46) and (57). At shock tube temperatures, methyl radical recombination can yield C2H6, C2H s + H, and C2H 4 -[-H2, while at lower temperatures characteristic of flames and plug flow reactors only the C2H 6 product channel

12

C.K. WESTBROOK and F. L. DRYER C2H 6 + H = C2H 5 + H E


C2H 6 + O = C2H 5 + OH

is important. The C 2 species thus formed then react as discussed below. Methane oxidation therefore occurs through two roughly parallel paths, the first consisting of direct oxidation of methyl radicals to methoxy radicals and/or formaldehyde and the second of methyl recombination followed by oxidation of the resulting C 2 species. The balance between the two paths depends on the fuel-air equivalence ratio. For example, in atmospheric pressure methane-air flames, Warnatz x36,13v estimates that for lean or stoichiometric conditions, about 30 ~o of the methyl radicals are consumed by recombination, with the percentage increasing to 80% for fuel-rich flames. In other mechanisms methyl radical recombination is less dominant, but the importance of this path for methyl radical consumption is now firmly established. Methane oxidation is somewhat unique among simple hydrocarbon fuels because the primary radical product CH 3 is so difficult to oxidize. Unlike many radicals such as formyl, ethyl, vinyl, and propyl, methyl radicals do not decompose readily to provide additional H atoms (see Table 3). The rate of reaction between c n 3 and 0 2 is slow and does not simply abstract another H atom to produce HO2. Therefore recombination of CH 3 plays a proportionally greater role in CH4 oxidation than analogous reactions for other fuels. For these reasons, methane oxidation chemistry is not typical of most other hydrocarbon fuels. The laminar flame speed of stoichiometric CH 4 air mixtures is lower than that for other fuels such as propane or ethane, is 6,a 57 the effective activation energy for shock tube ignition of CH4 air mixtures is higher than that of other alkane-air mixtures, ass and many other properties of methane-air mixtures are also somewhat anomalous. Thus the use of methane as a reference or standard hydrocarbon fuel is to be discouraged, although its convenience and availability often make it a desirable fuel for both experimental and modeling studies. 3.1.6. C2H 6 Detailed models of ethane combustion have been developed for laminar flames, isothermal flow systems and shock tubes. 132'136'137'159-164 In addition, we have already seen that ethane oxidation is an important part of methane oxidation mechanisms. Ethane decomposition reaction ( - 4 7 ) is at neither its high nor low pressure limit in most combustion regimes, so some treatment or estimate of its falloff behavior is usually necessary. Another initiation reaction with 0 2 may also contribute,162A63 but only at low temperature and in oxidizing environments
C 2 H 6 + 0 2 = C 2 H 5 + H O 2.

(62) (63) (64)

C2H 6 + OH = C2H 5 + H20.

The rates of these reactions exhibit substantial nonArrhenius behavior. 149,165 Two principal reactions involving CzH5 occur, C2Hs(+M) = C 2 H 4 + H ( + M )
C 2 H s + O 2 = CzH 4 + HOE.

(65) (66)

In shock tube conditions, reaction (65) can be assumed to be second order) 66 while other studies at lower temperature a6" indicate first order behavior. It is clear that this reaction is in the falloff region in most typical combustion environments.33'44'136 Competition between reactions (65) and (66) (see Table 3) is also a key feature of the ethane oxidation mechanism, since the H atom from reaction (65) is highly reactive. Other reactions of the ethyl radicals include C2H 5 + O = C H 3 C H O + H C2H 5 + 0 = CH20 + C H 3 CzH5 + H = CH 3 + CH 3 (67) (68) ( - 48)

but again, because radical concentrations are usually quite low, these reactions which depend on two radical species concentrations are usually less important than reactions (65) and (66). However, like many other radical-radical reactions, their importance often lies in their role as chain termination steps rather than their direct effect on the rate of ethyl radical consumption. The ratio k67/k68 is approximately 5 : 1 at room temperature, a6s but high temperature rate data are not available for these reactions. The reaction of CzH 5 with OH is fast 169 but neither the absolute rate nor the primary products have been determined. When ethane is the primary fuel, another key feature of the oxidation mechanism is the recombination of C2 and Ca radicals to produce C3 and C4 species, including
C z H 5 + C a l l 5 = C,HI0 CEHs+CH 3 =C3H 8

(69)
(-70)

CH3 +C2H 3 = C3H 6 C 2 H 5 + C z H 3 = C4H s C 2 H 3 + C 2 H 3 = C 4 H 6.

(-71) (72)

(73)

(60)

A second, relatively faster reaction follows immediately upon production of methyl radicals,
C 2 H 6 + C H 3 = C 2 H 5 + C H 4.

(61)

This important step provides a path for the relatively non-reactive CH3 to produce reactive ethyl radicals. The key reactions involving C 2 H 6 include

For ethane oxidation, reaction ( - 7 0 ) is the most important of these, a36 The effect of these radical recombination reactions on the chain reaction properties of these mechanisms is primarily one of termination, with two radicals consumed and a relatively stable molecule being produced. The larger fuel species are eventually consumed, but the immediate effect of their production is one of inhibiting the overall oxidation process. Recombination reactions (69) and ( - 7 0 ) were included in a recent C2H 6 oxidation mechanism, a62'163 used to describe ethane pyrolysis and oxidation over a wide temperature range (900 K < T < 1800 K ) a n d

Modeling of hydrocarbon combustion for both shock tube and plug flow conditions. In this comprehensive mechanism, the importance of two reaction paths was demonstrated, one consisting of CH 3 radical oxidation leading to formaldehyde and formyl radicals, and the other involving a dehydrogenation sequence leading through C2H5, C2H4, C2H3, and C2H 2. 3.l.7. C 2 H 4 Ethylene is a primary fuel itself and is also produced in large amounts during the combustion of CH~., C2H 6 and other higher hydrocarbons. A recent study of ethylene combustion 45 has provided a detailed comprehensive reaction mechanism which is able to reproduce experimental data over wide ranges of operating conditions, including shock tubes, laminar flames, plug flow reactors, and detonations. The elementary reactions of C2H and its product species demonstrate some properties which are not encountered with alkane fuels. Although many features of ethylene combustion have begun to be resolved, other details remain uncertain.

13

Three initiation reactions can be important, including C2H4+M = CzH2+H2+M C2H4+M = C2H3+H+M C2H4+C2H4 = CzH 3 +C2H 5. (74) (75) (76)

Reaction (74) is fastest in most cases 17'171 while reaction (76) rarely competes except in conditions in which C 2 H 4 concentrations are very large. 172'j73 In typical shock tube experiments only dilute mixtures of C2H 4 in argon or other diluents are considered, and reaction (76) is negligible under those conditions. The two important reactions of the vinyl radical are C2H 3 + M = C2H 2 + H + M C2H 3 + 02 = C2H 2 + HO 2. (77) (78)

An interesting series of shock tube experiments by Just et al. 174 involved the pyrolysis of extremely dilute C2H 4 in argon over a temperature range of 1700-2200K. Hydrogen atom concentrations were measured as a function of time. Under these very

~ "

P = 1.95 bar

1.5

4
II

/"

T2P"~c ~H'
P = 1.77 bar

0.5

I 200

[ 400 Time - #s

I 600

FIG. 6. Hydrogen atom profiles from shock tube pyrolysis of C2H 4. Open circles show experimental data of Just el a]., 1 "14curves are computed resultsff 5

14

C.K. WESTBROOKand F. L. DRYER C2H s radicals 167 formed by reaction ( - 6 5 ) , reaction (67), or isomerization of the adduct in reactions (81) and (82). Additional reactions of vinyl radicals include
C z H 3 + H = C 2 H 2 + FI 2 C2H3+0 = CH2CO+H (84) (85)

dilute conditions, only the initiation reactions (74) and (75), together with the vinyl decomposition reaction (77) need be considered. Since the vinyl decomposition is rapid compared with the other two reactions, the rate of H atom production is very nearly twice the rate of reaction (75). The results of this analysis are illustrated in Fig. 6, showing the measured data as solid circles. With E75 = 98.16kcal/mol as determined originally by Just et al., the agreement between the model and measurements is fairly good (dashed curves), but when E75 is increased slightly to 108.72kcal/mol, the agreement (solid curves) is excellent. The extreme simplicity of the mechanism for this series of experiments made it possible to evaluate one of the key kinetic parameters. However, even within this context the value determined by the modeling study is dependent on the other parameters which were assumed. In particular, possible falloff behavior of reactions (74) and (75), not included explicitly in the analysis, may have a significant influence on the rate expressions finally derived from the modeling analysis. Reactions of radicals with ethylene fall into two general classes. One consists of H atom abstraction reactions,
C 2 H 4 + H = C2H 3 + H 2 C2H 4 + O H = C2H 3 + H 2 0

C2H 3 + O H = C2H 2 +H20.

(86)

The combined effects of reactions (84) and ( - 7 7 ) can provide an inhibitory effect on the oxidation rate of both ethylene and acetyleneff6 catalyzing the recombination of H atoms to form relatively inert H2. Otherwise reactions (84-86) are generally much less important than reactions (77) and (78), since radical concentrations are usually relatively small. A reaction which has been suggested 17'~s2 as a possible step leading to soot formation is
C 2 H 3 + C 2H 4 = C 4 H 6 + H

(87)

producing 1-3 butadiene. The generation of large molecules with C/H ratios close to or greater than unity, beginning with small species such as ethylene and acetylene, is currently a rather poorly understood kinetic problem which deserves further attention. 3.1.8. C2H 2 Acetylene is a practical fuel and is also believed to be an important contributor to the formation and growth of soot. In addition, CzH 2 and its pyrolysis and oxidation reactions are part of the reaction mechanisms for most other hydrocarbon fuels, particularly in fuel-rich conditions. A number of modeling studies143.18z -189 have examined pyrolysis and oxidation of acetylene, but there was still a great deal of uncertainty involved in identification of major reaction paths, product distributions, and rate expressions. Very recently, studies of acetylene oxidation in laminar flames, 19 191 and a combined modeling study of shock tube and laminar flame oxidation of acetylene by Miller et al. '.6 have begun to provide valuable additional information on the mechanism. The important initiation reactions include C2H2 + M = C 2 H + H + M
CzH 2 + CzH 2 = C4H 3 + H C 2 H 2 + 0 2 = H C C O + OH

(79)
(80)

but another family of reactions involves the formation of an activated complex followed by rearrangement and fragmentation. These include
C2H4+0 = CH3+HCO

(81) (82) (83)

C2H 4 + O = C H 2 0 + C H 2 C2H4+OH = CH3+CH20.

The addition reactions (81 83) generally have smaller activation energies than the abstraction reactions. Therefore, at room temperature the abstraction reactions are often negligibly slow, but under high temperature conditions such as in shock tube oxidation, the abstraction reactions can be very important. There also appears to be a temperature dependence for the product distribution from the two C z H , , + O reactions. The total rate of reaction has been observed 175 to be curved on an Arrhenius diagram, which could be explained if E8z is not equal to E 8 r At present there is still some uncertainty both in the rates and the product distributions of CzH4 + O and C2H 4 + OH reactions and their variations with temperature. 176 178 Furthermore, because some of these reactions involve the formation of an activated complex,179 the rate expressions are somewhat pressure dependent. One problem encountered in C2H 6 oxidation which is less important in C2H4 combustion is the formation of larger species by radical recombination. However, for rich conditions, recombination reactions such as reactions (71-73) can still provide significant chain termination. Acetaldehyde has been detected in C2H4 oxidation, ls'181 but the amounts are quite small and may be a product of reactions between 0 2 and excited

(88) (89) (90) (91)

C2H 2 + O z = H C O + HCO.

In acetylene pyrolysis, reaction (88) dominates under dilute conditions while reaction (89) is more important at high fuel concentrations. It is interesting to note that, unlike the other hydrocarbons already discussed, the pyrolysis of acetylene is exothermic, so that a decomposition "flame" or detonation are possible 193 even in pure C2H2. Under oxidation conditions, the reaction of C2H 2 with O2 is extremely important 46 but there is not yet positive identification of the product distribution. Miller et al. assumed that reaction (90) predominated while Jachimowski~S3 employed reaction (91 ).

Modeling of hydrocarbon combustion Reactions between C 2 H 2 and radical species are complex. As discussed earlier in C2H4, many reactions with C2H2 involve the formation of activated species. Miller et al. suggest that as a result, pressure dependence may be the rule rather than the exception for many elementary reactions of species with multiple carbon-carbon bonds such as acetylene and ethylene. Hydrogen atom abstraction
C2H 2 + H = C2H +H 2

15 (106)

C 4 H z + O H = C3H2 + H C O

(92)

competes with the recombination reaction ( - 7 7 ) only at high temperatures. Reactions between C2H 2 and O atoms include C2H 2 + O = C2H + O H C2H2 + O = CH 2 + C O
C 2 H 2 + O = HCCO + H

followed by reactions between C3H 2 and radical species to give a variety of other products. Warnatz et al. 19 attributed as much as 35 ~ of the fuel consumption in rich acetylene flames to this sequence, which avoids ketene formation entirely. Other intermediate species found in these mechanisms include C3H 4 and C3H3. Elementary reactions of these C 3 species and their rates are not yet well established and further work is needed. 3.1.9. C H 2 C O Ketene oxidation has not been studied extensively under combustion conditions. It has been examined in a number of studies 46'~36,191,192,194-197 in its role as an intermediate, primarily in acetylene oxidation. Identification of specific product channels is not very far advanced. A summary of the reactions which have been used in modeling studies includes C H 2 C O + M = CH 2 + C O + M CHzCOWOH = CHzO+HCO CH2CO+OH = HCCO+HzO CHzCO+O = HCCO+OH CHzCO + O = HCO + HCO C H 2 C O + H -- H C C O + H 2 CH2CO + H = CH 3 + CO. (107) (108) (109) (ll0) ( 111 ) (112) (113)

(93) (94) (95)

with reaction (94) being predominant under most conditions. 46 Miller et al. showed that in laminar flame and shock tube simulations, the most important means of C2H2 consumption for lean, stoichiometric and even slightly rich conditions involved reaction with O atoms. Reactions with OH radicals include
C 2 H 2 + OH = CHzCO + H C2H2+OH C2FI 2 + O H = C2H+H20 ~ CH 3+CO

(96) (97)
(98)

as well as an addition path proposed by Miller et al. 46 C2H z + OH = C2H2OH (99)

followed by the reactions involving C2H2OH, including C2HzOH+H = CH2CO+H 2 C2HzOH+O = CH2CO+OH C2H2OH + OH -~ CH2CO + H 2 0 CzH2OH + 02 = CH2CO + HO2 C 2 H z O H + M -- C H 2 C O + H + M . (100) (101) (102) (103) (104)

Reaction (98) was included in many early C 2 species oxidation mechanisms, but recent results 46'136 confirm that it does not occur directly but was rather a shorthand for reaction (96) followed by reaction (113) (see below). Warnatz 136 and Levy et al. 191A92 assumed that the only path leading to ketene was reaction (96). That is nominally a chain propagation step but in practice it accelerates the rate of fuel oxidation because the H atom produced is then available for reaction (I). The paths involving C2H2OH also lead to ketene, but at the cost of two radicals for reactions (I00) and (102), or the loss of one radical in the case of reaction (101). Therefore the sequence initiated by the addition reaction (99) is somewhat more inhibiting than the direct reaction (96). An alternative path for acetylene oxidation under rich conditions has been examined recently, 46'19'~91 proceeding by means of C2H 2 + C z H = C4H z + H (105)

Another possible product path for C H 2 C O + O is C H 2 0 + CO, although no modeling studies have considered it. Rate expressions for reactions (107) 198 and (108) 10 have been obtained from experimental studies, but rate expressions for the other reactions represent either estimates or results of modeling studies rather than direct measurements. Ketyl radicals produced from ketene, directly from acetylene by reaction (95), or by reaction (132) between C2H and O z discussed below, can be consumed by a number of reactions, all of which are somewhat speculative at present. Miller et al. 46 suggested a set of reactions
HCCOWO 2 = COWCO+OH

(114) (115) (116) (117)


(118)

HCCO+O = CO+CO+H HCCO+H = CHz+CO HCCO+OH = HCO+H +CO HCCO +CH2 = C 2 H 3 + C O

based on room temperature flow experiments, 199'~ but there is no information available on these reactions at the higher temperatures of flames or shock tubes. As a result, most detailed mechanisms avoid modeling the reactions involving the ketyl radical by including only reactions (107), (111), and (113). As we noted earlier for the reactions forming and consuming methoxy radicals, the ketyl radical can often be eliminated from viable reaction mechanisms. However, the chain reaction properties of the paths involving HCCO are considerably different from those

16

C.K. WESTBROOK and F. L. DRYER to match experimental observations. Commonly the data consist of shock tube ignition delay times or laminar flame speeds, which are not sufficient to determine rates or product distributions of specific reactions. Usually only the overall chain branching rate and rate of heat release are really needed in modeling such systems. If individual species' histories or spatial profiles have been measured, including radic.al and trace species, then the rates and product distributions for each series or family of reactions can be determined. Unfortunately such data are rarely available, so the modeling process can only select reaction rates which reproduce the correct branching rates. The selection of these rates is of course restricted by the known thermochemistry of each species involved in a given reaction. However, even the thermochemical data for these species are not always well established. For example, the standard heat of formation of C2H was recently found 25 to be 127 kcal/mol rather than the earlier value ~1 of 113 kcal/mol, resulting in substantial modifications to computed equilibrium coefficients and reverse rates for reactions involving C2H. An additional acetylene pyrolysis submechanism has been proposed lsLls6 as a possible factor in soot formation. This path, proceeding by means of formation of polyacetylene species, consists of C4H3 + M = C 4 H 2 + H + M C2H2 +CzH = C4H2 + H C4H 2 + M = C4H + H + M (138) (139) (140)

which eliminate it. Like the methoxy radical, inclusion of ketyl reactions usually increases the chain branching rate of the mechanism. Other species produced during the oxidation of ketene and acetylene include C2 H, CH2, and CH. The reactions of methylene near room temperature have been reviewed recently by Laufer 20~ and include CH 2 + 0 2 = CO 2 + H 2 CH2+O 2 = CO2+H+H CH2+O 2 = CO+H20 CH2+O 2 = CO+OH+H CH2 + 02 = HCO + OH CH 2 + O = CH + OH CH2+O = CO+H+H CH 2 + O = CO + H 2 CH 2 + OH = CH + H 2 0 CH 2 + H = CH + H 2 CH 2 + C H 2 = CzH 3 + H CH 2 + C H 2 = C2H 2 + H E CH 2 + C2H 3 = CH 3 + C2H 2. (119) (120) (121) (122) (123) (124) (125) (126) (127) (128) (129) (130) (131)

Although there have been kinetic studies of the reactions of methylene in low pressure flames,22 very little solid information is available concerning product distributions of given reactions. Most of the rates summarized in Appendix I represent the results of modeling studies in which the branching ratios of these reactions were varied in order to provide the needed chain branching rate. Some of these reactions are redundant under certain conditions. For example, reaction (123) followed by the formyl decomposition reaction (35) results in the same product distribution as reaction (122) alone. Under rich and sometimes stoichiometric conditions, this compression of the mechanistic sequence can be appropriate, but under lean conditions the formyl radical will instead react preferentially with 02 (reaction (36)), so that reaction (122) would be inappropriate. Much more information is needed, particularly on the product distributions of these reactions. Reactions of C2H and CH include C2H + 0 2 = HCCO + O C2H+O 2 = HCO+CO C2H + O = CO + CH C 2 H + C 2 H 3 = C2H 2 +C2H2 CH+O2 = HCO+O CH+O2 = CO+OH. (132) (133) (134) (135) (136) (137)

in addition to reactions (88), (89), (92) and others listed earlier. Mechanisms including these species 4L 1 "/0,18 5,18 6 have been used primarily t o model acetylene pyrolysis in shock tubes. Computed results, including H atom profiles and induction times, are relatively insensitive to variations in their rates. The submechanism itself is plausible on kinetic grounds, but the mechanism through which the resulting polyacetylenes eventually produce soot has not yet been established. 3.1.10. CH30H Several models for methanol combustion have appeared, including pyrolysis26 .and oxidation.43'4v' 126,207-217 The first mechanisms were developed for specific environments including shock tubes and turbulent flow reactors. Based on these specialized models, a comprehensive reaction mechanism was developed 43 which simultaneously reproduced all of the existing experimental data. Subsequently, that comprehensive mechanism was used to predict properties of methanol oxidation in laminar flames47'211.217 and detonations.48 Possible initiation reactions include CH3OH+M = CH3+OH+M CH3OH+M = CH30+H+M CH3OH+M = CH2OH+H+M CH3OH+O 2 = CH2OH+HO 2 (141) (141a) (141b) (142)

Reactions of C2H with alkanes 2a and with 02204. have been examined at low temperatures. However, at high temperatures, many of these reactions have not been observed directly or the product distributions are still uncertain. The rate expressions represent reasonable estimates which have then been adjusted

Modeling of hydrocarbon combustion but reactions (141a) and (141b) can be rejected on the basis of relative bond energies. The addition of reactions (141a) and (141b) to a mechanism already containing reaction (141) had no appreciable effect on computed induction times under shock tube oxidation conditions. 4a Reaction (141) provides a much different chain initiation function than do the alternate initiation reactions, since both the methoxy and hydroperoxyl radicals quickly decompose, producing an additional H atom. Reaction (141) yields only a relatively non-reactive methyl radical and the chain propagating radical OH. On this basis it might at first appear that, even with substantially slower rates, reactions (141a) and (141b) might have strong influences on computed combustion histories. However, as we shall describe later, the influence of the initiation reaction persists for only a small fraction of the induction period until a modest radical population has been established, after which only the fuel-radical reactions are significant. Therefore, the lower rates and the short period over which initiation reactions are important make reactions (141a) and (141b) negligible. Reaction (142) is also unimportant at high temperature conditions relative to reaction (141). Abstraction of H atoms proceeds primarily by breaking C - H bonds rather than the O - H bond, including CH3OH+H = CH3+HzO CHaOH+H = CH2OH+H z CH3OH+OH = CH2OH+H20 CH3OH+O = CH2OH+OH CH3OH+CH 3 = CHzOH+CH 4 C H 3 O H + H O z = C H / O H + H 2 0 2. (143) (144) (145) (146) (147) (148)

17

Although they are not considered in current mechanisms, reactions between CH2OH and OH or O must also occur. The decomposition reaction is dominant in rich conditions and reaction (150) is more important in lean mixtures. Both reactions produce CH20, consistent with experimental observations 22,22x of high formaldehyde concentrations during methanol oxidation. The oxidation path for methanol proceeds in a sequential manner through CHsOH --*CH2OH -0 C H 2 0 -0 HCO ~ CO ~ CO 2. Because only small amounts of CH3 are formed, subsequent oxidation of C2 species is of less importance than in CH4 oxidation. The validation of the methanol oxidation mechanism is greatly simplified by the linearity of the overall reaction path. Some of the techniques and principles involved in the validation of a comprehensive reaction mechanism can be illustrated very effectively for the case of methanol. The starting point for this mechanism was a previously established CH4 reaction mechanism, including a treatment of C z species oxidation. Elementary reactions were added dealing with CH3OH and its immediate product species CH2OH (reactions (141150)). These additional reactions were assembled after a search of the literature for data on specific reactions involving CHaOH and CHzOH. In some cases the initial rate expressions were estimated. This composite mechanism was then used to simulate a number of experiments dealing with methanol oxidation. Rate expressions for only these newly incorporated reactions were varied to obtain the best agreement between computed and experimental results. The experiments to be analysed were selected carefully so that each emphasized different parts of the new mechanism. Intermediate temperature (T ~ 1000K) turbulent flow reactor data 222 for lean conditions depended most on the rates of reactions (145) and (150), while similar data from a rich mixture depended on the rates of reactions (143), (144), and (149). Models for high temperature (1545 K < T < 2180K) shock tube data 2v in contrast emphasized the initiation reaction (141). In addition, the rich shock tube models depended on the rates of the same C H 3 O H radical reactions as the rich flow reactor calculations, and similarly for the lean mixtures. Therefore, an array of experimental data could be used to isolate one or two elementary reactions at a time, greatly simplifying the determination of reasonable rate expressions. The computed results also showed that some of the reactions, including reactions (142) and (147), contributed negligibly to the observed rate of methanol consumption under these conditions. This conclusion does not necessarily mean that such reactions should be omitted from future mechanisms. Reaction (147) is relatively unimportant in part due to the low methyl radical concentrations encountered in methanol oxidation. However, in fuels consisting of CH4-CH3OH mixtures for example, reaction (147) might be important.

For the reactions between C H 3 O H and H, reaction (143) is faster below 650 K, z's while the rates reverse at higher (T > 1000K) temperatures. 43 If reaction (143) or any other fast reaction produced CH a radicals in large amounts, then methyl radical recombination (reactions (47-49)) would result in considerable amounts of C2 species. However, very low concentrations of C2 species are observed in C H 3 O H - a i r flames. 219 The influences on the mechanism of reactions (143) and (144) are substantially different. Reaction (144) produces Hz and CH2OH which decomposes rapidly to yield H atoms CHzOH+M = CH20+H+M (149)

while reaction (143) produces the much less reactive C H 3 radical and the inert species HzO. In fuel-lean conditions, reaction (145) with OH is responsible for the majority of the CHaOH consumption, while in rich mixtures, the reactions with H atoms are most important. The CHzOH radicals which are produced then decompose by means of reaction (149) or react with 02 and H CH2OH+O2 = CH20+HO2 CH2OH + H = C H 2 0 + H 2.
JPEC$ I O : I - B

(150) (151 )

18

C.K. WESTBROOK and F. L. DRYER The construction of this ethanol reaction mechanism is another example of the use of the hierarchical structure of hydrocarbon oxidation mechanisms. Only reactions (152-161) are added to account for the new fuel molecule and its immediate product CH3CHOH , the same method used to develop a comprehensive mechanism for methanol oxidation. However, ethanol mechanisms are still somewhat speculative, and further development will be required, based on the availability of experimental data in a variety of environments. 3,1.12. C H 3 C H O The combustion of partially oxidized hydrocarbon species in general, and alcohols and aldehydes in particular, is an important part of the analysis of many practical kinetic systems. Although acetaldehyde is not an important fuel itself, it can be a significant intermediate in the combustion of higher hydrocarbon fuels and its oxidation properties can be expected to be similar in many ways to those of other aldehydes. Acetaldehyde pyrolysis and oxidation 22S'226 at high temperature have been examined using detailed kinetic models. These studies greatly simplified the C2H 6, C2H4, and CH 4 portions of the mechanisms, emphasizing the reactions involving CH3CHO, CH3CO, and CH2CHO. The principal initiation reactions are CH3CHO = CH 3 + H C O CH3CHO = CH3CO + H CH3CHO+O 2 = CH3CO+HO2 (162) (163) (164)

In a study of methane-oxygen combustion in flames, Harvey and MacCol1223 detected significant amounts of CH3OH. The reverse of reaction (141) is too slow to be able to account for the observed formation rate of methanol. However, an alternative mechanism in the cooler parts of these flames (800 K < T <.1200 K) may involve methoxy radicals, C H 3 0 + C H 3 0 = CH3OH + C H 2 0 CH30+RH = CH3OH+R where the methoxy radicals are produced primarily through reaction (55) between methyl and hydroperoxyl radicals. 3.1.11. C z H s O H Very little kinetic information is yet available on the species and elementary reactions involved in ethanol oxidation at high temperatures. Bhaskaran et al. 126 recently reported shock tube ignitions delay data for C z H s O H - O z - A r mixtures and constructed a detailed mechanism to simulate the experiments. Based in principle on mechanisms for methanol oxidation, the initiation reactions chosen are CEHsOH+M = CHa+CHzOH C z H 5 O H + O 2 = C H 3 C H O H + H O 2. (152) (153)

Reactions between radical species and ethanol included CEHsOH+OH=CH3CHOH+H20 C2HsOH+H = CH3CHOH+H 2 CzHsOH+O = CH3CHOH+OH (154) (155) (156)

C E H s O H + H O 2 = C H 3 C H O H + H z O 2 (157) C 2 H s O H + C H 3 = C H 3 C H O H + C H 4. (158) These are the analogs of reactions of methanol discussed earlier. The hydroxyethyl radical CH3CHOH then is consumed by the same paths as CHzOH, producing acetaldehyde CHaCHOH+M=CH3CHO+H+M C H 3 C H O H + O 2 = C H 3 C H O + H O 2. (159) (160)

with the decomposition reaction dominating at shock tube temperatures and reaction (162) more important than reaction (163). At flow reactor temperatures (T - 1000K) reaction (164) and reactions (9) and (36) which also produce HO2 are very important and lead to high concentrations of hydroperoxyl radicals. Because of these high HO 2 levels, much of the CH3 radical consumption can be attributed 226 to reaction

(55)
CH3+HO 2 =CH30+OH (55) but at higher temperatures this sequence should not be significant. The principal H atom abstraction reactions include CH3CHO+H = CH3CO+H 2 CH3CHO+OH = CH3CO+H2 O CH3CHO+O = CH3CO+OH CHaCHO+CH 3 = CH3CO+CH 4 CH3CHO + HO 2 =
C1L'I3CO

The production of acetaldehyde as the dominant intermediate is therefore somewhat analogous to the production of formaldehyde as the major intermediate in methanol combustion. The mechanism for the subsequent oxidation of acetaldehyde is discussed in the next section. More recently, a turbulent flow reactor study of ethanol pyrolysis and oxidation z24 has shown that, similar to the case for methanol, two hydrogen abstraction reactions are important at high temperature, consisting of reaction (155) as well as C 2 H s O H + H = C2Hs + H 2 0 . (161)

(165) (166) (167) (168) (169)

q- H202

followed by the decomposition reaction CH3CO = CH 3 + C O . (170)

The additional path to produce CzH 5 and then ethylene indicates that the oxidation and pyrolysis of both ethylene and acetaldehyde are essential submechanisms in the reaction mechanism for ethanol combustion.

Abstraction of the H atoms from the methyl group, followed by decomposition of CHzCHO to produce ketene has also been proposed :z6 but is likely to be unimportant relative to the above reaction sequence. The above mechanism leads to large concentrations of

Modeling of hydrocarbon combustion CH 3 radicals and CO. The methyl radicals produce CH4 and recombine with other methyl radicals to produce C2H6 . Subsequent reactions of C2H 6 lead to C2H4, C2H2 and other related species as already discussed. These trends are consistent with experimental data in which relatively high levels of CO, CH 4, C2H6, and C2H 4 are observed. Thus ketene is a minor product, perhaps arising as much from the secondary C2H 4 oxidation as directly from acetaldehyde. 3.1.13. C3H 8 As we have seen already, construction of a detailed reaction mechanism is a sequential or hierarchical process, beginning with the simplest species and building up to include larger and more complex molecules. At present the frontier in this development pattern is represented by the C 3 and C 4 hydrocarbons. Recently substantial progress has been made, with mechanisms describing the pyrolysis and oxidation 118'136'137'195-197'227-243 o f propane and propylene in flow reactors, shock tubes, and laminar flames. Many of the principal overall features of the combustion of propane have now been identified. The primary initiation step is
C3H 8 = CH 3 +C2H s

19 (182) (183)

C3Hs+C3H 5 = iC3HT+C3H 6 C3Hs+C3H 5 = r/C3H7 + C 3 H 6 and for oxidation C3H s + O = iC3H 7 + O H C3H s + O = nC3H 7 + OH C3H 8 + OH = iC3H 7 + H20 C3Hs + O H = nC3H7 + H 2 0 C 3 H s + H O 2 = iC3HT+H20 2 C 3 H s + H O 2 = nC3HT+H202.

(184) (185) (186) (187) (188) (189)

As observed previously for similar reactions with methane and ethane, the rate expressions for H atom abstraction by O, H, and OH all display significant non-Arrhenius temperature dependence. 19L244 Furthermore, reactions with OH are generally most important for propane consumption in lean and stoichiometric conditions, and reaction with H atoms and CH 3 radicals most important for rich conditions.195'235 The thermal decomposition of the propyl radicals gives nC3H 7 = C2H 4 + CH 3 1190)
nCaH 7 = C3H6+H

(191) (192) (193)

(70)

iC3H7 = C2H 4 + CH 3 iC3H 7 = C3H 6 + H

which is much faster than the reactions involving the breaking of a C H bond
C3H s = H + iC3H 7 or

C3Hs = H + nC3H 7 due to the smaller C - C bond energy. For oxidation conditions, other initiation reactions can include
C3H s + 0 2 = iC3H 7 + HO e (171)

C3H s + 02 = nC3H7 + HO 2

(172)

although they will be much slower than reaction (70) at shock tube temperatures. Propane is the first species to be encountered here in which distinctions between isomeric forms of a given species must be considered. For modeling purposes each isomeric form can conveniently be considered as a unique species, together with reactions involving both forms, such as
i C 3 H v + C 3 H 8 = n C 3 H T + C 3 H s.

although on structural grounds 245 the rate of reaction (192) will be very small. Reactions (191) and (193) have a somewhat accelerating overall effect on propane oxidation rates since they produce H atoms, while reaction (190) produces two relatively inactive species, CzH 4 and CH 3. For this reason, an important feature of the reaction mechanism for Call 8 consumption concerns the relative amounts of n-propyl and iso-propyl radicals which are produced in reactions (174-189). Some theoretical arguments have been presented 165 for these relative rates, and modeling results 195 can use the sensitivity to these branching rates to estimate them, but direct experimental determinations of the independent rates of abstraction of primary and secondary H atoms in C3H 8 at combustion temperatures would be very useful. Some models have included
nC3H7 + 0 2 = C3H 6 + H O 2 iC3H 7 + 0 2 = C3H 6 + HO 2

(173)

(194) (195)

Radical species reactions with C3H 8 include, for pyrolysis C3H s + H = iC3H 7 + H 2 C 3 H 8 + H = nC3H7 + H 2 C3H 8 + CH 3 = iC3H 7 + CH 4 C 3 H s + C H 3 = nC3HT+CH 4 C3Hs+C2H 3 = iC3H7+C2H 4 C3H8+C2H 3 = rtC3H7 +C2H 4 C3H8 +C2H 5 = iC3H7+C2H 6 C3Hs+C2H s = nC3H7+C2H 6 (174) (175) (176) (177) (178) (179) (180) (181)

with estimated rate expressions. Although analogous reactions involving HCO, C2Hs, C2H3, and CH2OH have been shown to be important in stoichiometric and lean mixtures, Warnatz 11s'~36 has shown that, compared with the thermal decomposition reactions, reactions (194) and (195) are negligible in atmospheric pressure laminar flames. Generally, as the size of the alkyl or other radical hydrocarbon species increases, the thermal stability of the radical decreases and its decomposition rate grows. Therefore the relative importance of the reaction with 0 2 molecules (or any other species) will decrease. The flame modeling

20

C.K. W~STaROOKand F. L. DRYER anisms and current propene oxidation mechanisms. Just as early ethylene mechanisms (e.g. [38, 183]) simplified the process by assuming that the primary products of radical attack on C2H4 were species such as CH20, HCO, and CH3, early models of C3H s combustion have either neglected C3H 6 consumption or greatly simplified it. McLain and Jachimowski T M assumed that the only reaction ofC3H 6 was C3H 6 + O = CH20 + C2H 4 (197)

results of Warnatz show that, in addition to the propyl and butyl radicals explicitly studied, larger alkyl radicals will be consumed in laminar flames almost entirely by thermal decomposition. However, at lower temperatures (T ~< 1100 K) the larger activation energies of the decomposition reactions sharply reduce the rates of these reactions, and the reactions between radicals and 0 2 which produce HO 2 again become competitive. Ratios of rates have been measured at room temperature for reactions between propyl radicals and O atoms168 nC3H 7 + O -~ C2HsCHO + H nCaH 7 + O = CH20 + C2H5 iCaH 7 + O : CH3COCH 3 + H iC3H 7 + O = C H 3 C H O + C H 3 but absolute rate measurements are not available, and high temperature rates and product distributions may be different from the room temperature values. Similar reactions between propyl radicals and H or OH radicals can also be considered, although the combination of low radical concentrations and rapid propyl decomposition rates in most combustion environments make such reactions usually of minor importance. Recombination paths between H atoms and propene are important in both propane and propene combustion. In particular, the sequence C 3 H 6 + H = nC3H 7 nC3H 7 = C2H 4 + CH 3 ( - 191) (190)

with an estimated rate expression. This path will result in relatively low CH 3 levels and low production rates of methane. Warnatz t36 included two reactions dealing with propene, producing acetaldehyde and CH3CO radicals C3H 6 + O = CH 3 + CH3CO C 3 H 6 + O H = CH3 +CH3CHO. (198) (199)

This path leads to high methyl radical concentrations, as noted earlier in the discussion of acetaldehyde mechanisms. Cathonnet e t al. 23s included a number of reactions in place of reactions (197-199), including C 3 H 6 + O = C2H5 + H C O C3H 6 + OH = C3H 5 -I-H 2 0 C 3 H 6 + O H = C2H 5 + C H 2 0 C3H 6 + H = C3H 5 + H 2 C3H 6 + C H 3 = C3H 5 + C H 4 C3H6+HO2 = C 3 H 6 0 + O H C3H6+C2H 5 = C3H5+C2H 6 (200) (201) (202) (203) (204) (205) (206)

has a strongly inhibiting influence on the overall rate of combustion, while the sequence iCaH 7 = C3H 6 + H Call 6 + H = nC3H7 (193) ( - 191)

is equivalent to an isomerization reaction between the two propyl radicals. Since nCzH7 does not produce C 3 H 6 + H in any significant amounts, the reverse isomerization from nCaH7 to iC3H 7 through C3H 6 + H is not important. Up to this point, the mechanism construction has been straightforward and relatively unambiguous. However, the subsequent oxidation of propene may be considerably more complex. Two initiation reactions for propene ignition have been considered extensively C3H 6 = C2H 3 + C H 3 C3H 6 = C3H 5 + H. (71) (196)

most of which are likely to result in low methyl radical production rates. Finally, Perry 246 ruled out H abstraction by O atoms for temperatures as high as 820K, but at shock tube temperatures such a path might occur. The fate of the C3H 5 radicals produced in many of these reactions is obscure. Although the model of Cathonnet e t al. predicted the production of considerable amounts of C3H 5 (more than 10 times the C2H5, H, or OH levels), no provision was made for its consumption. Similarly, Hautman e t a/. 227 did not include any reactions which would remove C3H 5 during propane pyrolysis, although concentrations of 100 ppm of C 3 H 4 were observed at 1150K. There is also no indication whether this Call 4 is propyne (methyl acetylene) or propadiene (allene). Burcat 229 measured both propyne and propadiene in shock tube studies of propene pyrolysis, but accompanying modeling analysis did not distinguish between formation routes for the two forms of C3H 4. The principal reactions of C3H 5 include C3H 5 = C3H 4 + H C3H5+O 2 = CaH4-4-HO 2 C3H 5 + C H 3 = 1C4H8 C3H 5 + H = C3H 6 C3Hs+H = C3H4+H 2 C3H 5 + C H 3 = C3H4+CH 4 (207) (208) (209) ( - 196) (210) (211)

Recent modeling analysis 19L243 indicates that reaction (71) is more rapid in shock tube conditions, although resonance stabilization of the allyl radical bring the rate of reaction (196) much closer to that of reaction (71) than might be expected from a consideration of C - C and C - H bond energies alone. There is a noteworthy parallel in the published literature between the early ethylene oxidation mech-

Modeling of hydrocarbon combustion reactions (210) and (211) were included '95 to help reproduce observed rates of C3H4 formation in the shock tube pyrolysis experiments of Burcat. 229 Reaction (208) was considered to simulate C a l l 5 consumption under oxidizing conditions, t95 although it is unlikely that this path represents an elementary reaction of real significance. It is more probable that reactions of the form C3H 5 + O = products C3H 5 + O H = products will actually prevail on an elementary level. These products should include both abstraction as well as addition channels, but no detailed information is currently available on such reactions at combustion temperatures. Because of the high activation energy of reaction (207) (70kcal/mo1239), it is unlikely that thermal decomposition of allyl radicals will be important. The unusual stability of the allyl radical makes it particularly likely that its reactions with other radical species will provide a major portion of its total consumption rate. C3H 4 is then destroyed by a combination of CaH4+O = CH20+C2H 2 C3H,~+OH = CH20+C2H 3 C3H4+O = HCO+C2H 3 C 3 H a + O H = H C O + C 2 H 4. (212) (213) (214) (215) C4H 10 + O H = sC4H 9 + H 2 0 C4H10 + H O 2 = PC4H 9 + H 2 0 2 C 4 H l o + H O 2 = sC4H 9 + H 2 0 2 C4H10 + C H 3 = PC4H 9 + C H 4 C4H10 + CH3 = sC4H 9 + C H 4
C 4 H l o + C 2 H 3 = pC,,H 9 + C2H 4 C4H1o + C 2 H 3 = sC4H 9 + C2H 4

21 (224) (225) (226) (227) (228) (229) (230) (231) (232) (233) (234)

C4Hlo + C2H 5 = pC4H9 + C2H6 C4Hlo + C2H 5 = sCaH 9 + c 2 a 6 C4H10 + C 3 H 5 = pC4H 9 + C 3 H 6 C4Hlo + C3H5 = sC4H 9 + C3H 6. The butyl radicals can then react pC4H 9 = C2H 5 + C2H 4 sC4H 9 = C3H 6 + C H 3 pC4H 9 = 1C4H s + H sC4H 9 = 2Calls + H pCaH 9 + sCaH 9 +
0 2 ~--- 1C4H8 0 2 =

(235) (236) (237) (238) (239) (240) (241)

+ HO 2

1C4H s + H O 2

sC4H 9 + 0 2 = 2C4Hs + H O 2

although reactions with 0 2 molecules are probably even less important than for propyl radicals, as discussed above.118.136 Reactions of butene include 1C4H 8 + H = C4H 7 + H 2 2C4H8 + H = C4H7 + H 2 1C4H8 + O = C a H 6 + C H 2 0 2 C 4 H 8 + O = nC3H7 + H C O IC4Hs + O H = nC3H 7 + C H 2 0 2C4Hs + O H = C2H 5 + C H 3 C H O 1C4H8 + C H 3 = C4H 7 + C H 4 2C4Ha + C H 3 = C4H 7 + C H 4 2C4H8 + O = C2H5 + C H 3 C O . The C 4 H 7 radicals are assumed to decompose C#H 7 = C4H 6 + H C4H 7 = C2H 4 + C2H 3 in addition to C4H7 + C4H 7 = 1C4H8 + C4.H 6 C4H7 + C4H 7 = 2C4H8 + C4H 6 C4H7 + C H 3 = C4H 6 + C H 4 and possibly (253) (254) (255) (251) (252) (242) (243) (244) (245) (246) (247) (248) (249) (250)

Like the reactions of ethylene and acetylene discussed earlier, the reactions suggested here consist of an addition followed by a rearrangement and decomposition. Future efforts are needed to determine the important primary oxidation paths for propane and its major intermediate products. At present the rates and products are known only for the principal reactions between C3H s and radical species including H, O, and OH. Very little reliable information is available for the reactions of C a l l 6 and c 3 n , , , as well as for the C 3 radical species. 3.1.14. C4Hlo A number of detailed modeling studies of n-butane combustion have appeared recently.1 ~s, ~36,235,24o,241, 247-249 The level of sophistication in the n-butane mechanism is comparable to that of propane, beginning with the initiation steps C4Hlo = C2H5 + C 2 H 5 C4Hlo = nCaH 7 + C H 3 C4Hlo +
0 2 =

(216) (217) (218)

sC4H 9 + H O 2

followed by the abstraction reactions C4Hlo + H = P C 4 H 9 + H 2 C4Hlo + H = sC4H 9 + H 2 C4H1 o + O = pC4H 9 + O H C4H10 + O = sC4H 9 + O H C4Hlo + O H = pC4H 9 + H 2 0 (219) (220) (221) (222) (223)

C4H7 + 0 2 = C4H 6 + H O 2. The description of the consumption of C 4 H 6 is in the same condition as that of C a l l 4. Little is known about its reactions, and there are virtually no data available at temperatures above 1000 K. Much of the n-butane reaction mechanism, as well as a considerable portion

22

C.K. WESTBROOK and F.L. DRYER position reaction products. In this way the model could account for the experimentally observed yields of the C2-C 7 olefins. However, like the above model for n-heptane ignition, this mechanism is limited by a lack of data on specific reaction rates and product distributions at high temperatures. For large hydrocarbon molecules, site-specific reaction rates are necessary for H-atom abstraction, since the subsequent reactions of each isomeric form of a given radical are so important. Although its temperature range (650 840K) is slightly lower than most of those being discussed here, the recent study of n-hexane pyrolysis by Ebert et al. 252 is an example of a model in which a great amount of site-specific detail is included explicitly. This mechanism, based on those of Edelson and co-workers, 242'253'254 considers isomerization of C 6 and C 5 alkyl radicals, as well as various aromatization reactions. Generally speaking, the relatively small number of modeling studies of large hydrocarbon combustion have been remarkably successful in interpreting experimental results. This success provides some assurance that the principles and rate expressions established in the development of mechanisms for smaller hydrocarbon molecules remain valid when the fuel molecule size increases. All of the alkane fuels we have discussed here have been straight-chain hydrocarbons. One important area in which experience with previous mechanisms may not apply is that of branched-chain molecules. Sundaram and Froment 241 found significantly different product distributions for the pyrolyses of nbutane and iso-butane, incorporating these differences into their detailed mechanism. A great deal of further work, both experimental and modeling, is needed to improve the mechanistic understanding of high temperature combustion of branched-chain hydrocarbons. Further extensions of these mechanisms to other large hydrocarbon fuel molecules are occurring gradually. Reliable, comprehensive detailed models are just now being established for C a hydrocarbons and in a few cases for C# species as well. For many practical applications detailed mechanisms for propane and butane should be sufficient. As we commented earlier, the combustion properties of methane and ethane are not really typical of most higher hydrocarbon fuels, but propane and butane exhibit many of the same properties as practical hydrocarbon fuels, including burning velocity, vapor pressure, quenching properties, ignition behavior, effects of pressure on combustion rates, and other important macroscopic properties of practical fuels. Therefore there is an immense body of experimental data and practical environments for which C3 and C 4 reaction mechanisms will be sufficient. For those specific problems where larger fuel molecules are essential, specialized mechanisms are being developed. Each such mechanism provides additional experience in modeling large sets of reactions, helping to establish the amount of detail needed to simulate combustion in a variety of practical combustion systems. For

of that for propane, is based on lower temperature experiments where different mechanisms apply, 4 so the application of these data to higher temperatures is somewhat speculative. Furthermore, identification of major product channels for reactions of C4H8, C4H6, C3H6, and C3H 4 is lacking. 3.1.15. Higher hydrocarbons For hydrocarbons heavier than butane, reaction mechanisms with the same degree of detail as those already discussed become very large. One attempt to model the oxidation of such a fuel has been that of Coats and Williams25 dealing with the shock tube ignition of n-heptane. The initial reaction steps are reasonably complete, including nCvH16 = C5Hll +C2H5
nCvH16 + H = CTH 15 + H2

nCvH16 + 0 = C7H15 + OH nCTH16+OH = CTH15+H20 nCvH16+HO 2 = CvH15+H202 nCTH16 + CH3 = C7H15 + CH4 nC7HI 6+C2H5 = CTH15+C2H6 followed by an abbreviated reaction scheme for the radicals C7H15 = CsHll +C2H4 CsHll = C 3 H 7 + C 2 H 4 C3H 7 = CH 3+C2H 4 and a more detailed C 2 oxidation mechanism. This type of approach simplifies those parts of the mechanism which are not known but includes all of the detail where it is available. In this sense it is similar to the quasi-global approach discussed later in this paper. There are difficulties which can arise from the particular way the unknown parts of the mechanism are chosen. In the above n-heptane mechanism, all of the radical decomposition reactions produce ethylene, so this overall mechanism will be very dependent on the C2H4 oxidation submechanism. Such a sequence is appropriate only for the case of abstraction of a primary H atom from the original n-heptane molecule. Decomposition steps subsequent to abstraction of H atoms from secondary sites will produce higher alkenes including propene and butene in addition to ethylene. The Coats and Williams mechanism will predict high soot precursor levels due to the high concentrations of C2 species. In fact, Coats and Williams used their mechanism to examine possible soot formation paths, but their choices for the product distributions for these reactions have ensured that their calculations will produce soot precursors. Recently, Doolan and Mackie T M presented a detailed mechanism for the pyrolysis of n-octane in shock tubes. In their treatment, isomerization reactions between various forms of each alkyl radical were included in an approximate manner, and each isomeric form then produced a unique set of decom-

Modeling of hydrocarbon combustion other applications in which higher hydrocarbon fuels are required, a good alternative at present lies in the quasi-global or other simplified mechanisms to be described later. 3.1.16. Aromatics Detailed mechanisms for the combustion of aromatic hydrocarbon fuel species have not yet been developed, although some preliminary modeling studies of benzene combustion in shock tubes have appeared. 255'256 Interest in the high temperature chemistry of aromatics has been stimulated by the fact that aromatics are becoming a major component of liquid fuels, and that this class of compounds has a propensity to form soot when burned. Several works have appeared recently 257,258 and discuss some of the principal elements to be considered in constructing complex models for the single ring aromatics. In the case of benzene, the principal initiation reactions include C6H 6 "4-M = C6H 5 + H + M C6H 6 + 0 2 = C6H 5 + HO 2 as well as abstraction of H atoms from the ring by various radicals c 6 n 6 + H = C6H 5 + H 2 C6H6+O = C6H5+OH C6H 6 + OH = C6H 5 + H 2 0 C6H6 + HO 2 = C6H 5 + H 2 0 2 At low temperatures 259 it appears that these reactions are followed by
C 6 H 5 - ] - 0 2 = C6H5OO

23

major intermediates detected in these flames ( C 6 H 6 0 and C5H6). These species would likely be derived by direct addition of oxygen atoms to benzene and expulsion of carbon monoxide, and suggest yet another route leading to ring destruction. Oxidation of alkylated single ring aromatics appear to proceed by initial oxidation of the side chain, with the subsequent reactions leading to ring destruction continuing to be the rate-controlling steps. It has been suggested that toluene has a mechanism which is different from those of alkylated aromatics with larger side chains (which probably proceed through the oxidation of phenyl radicals). Benzyl radicals formed in toluene oxidation apparently can neither form a double bonded analogue nor be oxidized by molecular oxygen. 262 It is not likely that detailed mechanisms for aromatic species oxidation will be available for some time. The important abstraction, addition, and ring destruction reactions remain to be more clearly identified and their rates determined before serious modeling efforts can proceed. Finally it is worth noting that at high temperatures, direct ring fragmentation may further complicate the mechanistic description. 263 3.1.17. Extensions to other fuels We have emphasized the combustion of hydrocarbon fuels in our discussions, in part to provide a convenient set of limits to this discussion. For a number of reasons, partly economic and partly historical in origin, hydrocarbons have always received the most practical attention. However, the basic formalism developed here is certainly not restricted to one family of fuels. We shall discuss below the coupling of other chemical species to hydrocarbon oxidation mechanisms, but many other fuel systems can be examined with the present methods. In general, the detailed approach to kinetics modeling should be applicable whenever the fuel molecule consists of a fairly small number of atoms. Very recently, significant advances have been made in modeling the pyrolysis and oxidation of ammonia (NH3) in flames, shock tubes, and flow systems. 264 267 In addition to these analyses of NH 3 combustion, the NH3_NO_O2 reaction268 277 is of great interest since this system represents an important kinetic approach to limiting pollutant emissions from practical combustors. Other non-hydrocarbon fuels to which detailed modeling approaches have been applied, with varying degrees of success, include C S 2 and C O S , 278-280 H2S, 281-287 and HCN and C 2 N 2 .288-295 Detailed modeling studies of the gas phase oxidation of nitromethane (CH3NO2) 296'29v and methyl nitrate (CH3ONO2) 29s have also been published recently. Another practical fuel which should also be suitable for the present approach is hydrazine (N2H, 0. Combustion of sulfur-bearing species and the relevant kinetic mechanisms have been reviewed, 299-32 but many reactions and their rates are still poorly understood. In addition to reaction mechanisms which

with ensuing steps leading to final products. At higher temperatures, thermochemical calculations 26 and recent flow reactor 15s and flame 261 results indicate that other routes must dominate. Intermediate products appearing in flow reactor experiments suggest that the reaction of phenyl radicals with molecular oxygen lead to phenoxy radicals and an oxygen atom C6H 5 + O 2 = C 6 H 5 0 + O with subsequent expulsion of carbon monoxide leading to the next lower hydrocarbon (cyclopentadienyl) radical C6H50 = CO + CsH 5. This sequence is followed by similar reactions leading to butadienyl radicals, its decomposition products (vinyl acetylene and vinyl radicals) and butadiene. In near-sooting premixed low pressure flames of benzene, Bittner and Howard 26~ have suggested that the decomposition and reaction of phenyl radicals with molecular oxygen and hydroxyl radicals are too slow to account for their experimental observations. They suggest that the only plausible alternative might be destruction of phenyl by oxygen atoms. This mechanism would continue to leave unexplained

24

C.K. WESTBROOK and F. L. DRYER often modeled by adding NH 3 or HCN to the fuel. There are a number of excellent reviews of current status of modeling pollutant species formation, 1 3 and we will not go into further detail here. From a modeling point of view, NO x and SOx formation can often be considered as trace species mechanisms. This means that the pollutant species mechanisms do not materially affect the radical species concentrations which are instead established and maintained by the fuel oxidation mechanism. Mathematically, this means that the pollutant species equations decouple from the other differential equations and can even be integrated independently. However, in other situations the trace species compete effectively for available radicals and the entire coupled kinetic problem must be considered.
4. KINETIC MODIFICATIONS TO MECHANISMS

describe the combustion of these species as fuels themselves, additional complications arise when cyanogens or sulfur compounds are present in a hydrocarbon oxidation environment. Such situations occur frequently in studies of pollutant emissions from combustors in the form of oxides of nitrogen (NOx) and sulfur (SOx), When the hydrocarbon oxidation and the pollutant species mechanisms are combined, additional reactions coupling the individual families of reactions can be important. That is, the hydrocarbon and N or S oxidations do not necessarily take place in parallel. Often, new species must be considered, such as HSO, HSO2, COS2, HNO, and NCO, as well as elementary reactions between species in the different mechanisms, such as CH3+NO 2 =CH30+NO H + S O 2 -- S O + O H N+SO = NO+S. We have tacitly assumed in these discussions that O z is the oxidizer. For modeling purposes it is often useful and interesting to use some other oxidizer instead, such as ozone or nitrous oxide. Detailed reaction mechanisms for fuel-N20 mixtures in shock tubes and in flames 33 307 have been developed. Other numerical models of hydrocarbon oxidation have appeared in which N 2 0 or NOz have been added to fuel-O 2 mixtures. From a kinetic basis, such studies are interesting because they emphasize different parts of the detailed reaction mechanism. If 02 is not present at all initially, then the usual chain branching reaction (1), H + O 2 = O + OH, is of much less importance in generating a radical pool. Instead, the initial source of radicals is N20+M = N2+O+M followed by H2+O = H+OH N20+O = NO+NO N20+O =N2+O 2
N20 +

In the preceding sections we have discussed mechanisms by which hydrocarbon fuels are consumed at intermediate and high temperatures. We now want to consider kinetic methods of altering the combustion of these fuels. Experimentally it is well known that the addition of certain chemical species to a fuel-oxidizer mixture can have a dramatic influence on its combustion properties, even when the additive is present in very small amounts. This can slow down or even halt the combustion, in which case the additive functions as an inhibitor, or it can accelerate the combustion process in which case the additive is a reaction promoter. Another term often used to describe the promotion of combustion is kinetic sensitization. In addition to the practical aspects associated with the reaction modification, analysis of these effects can often provide valuable insights into the fundamental processes occurring in many other environments as well. 4.1. Inhibition All of the reaction mechanisms we have seen depend on the existence of a radical supply or pool. Chain branching reactions accelerate the combustion process by increasing the size of the radical pool, and termination reactions slow the combustion by reducing the size of the pool. Many kinetic inhibitors act by enhancing radical removal rates, lowering the radical pool level. Perhaps the best known inhibitors are those containing halogen atoms F, CI, Br, and I. These species act primarily by catalyzing the recombination of H atoms. If we denote a halogen atom by X, inhibition by a halogen acid HX is given by the net result of three reactions: HX+H
= H2+X

H = N 2 + OH.

The last three reactions also play a part in the formation of oxides of nitrogen from atmospheric nitrogen. The most common NO~ reactions consist of the extended Zeldovich mechanism N+O 2 = NO+O N2+O = NO+N N+OH = NO+H. Because of their relatively high activation energies, these reactions generally produce NO at appreciable rates only at high temperatures. In contrast with the process of NO~ formation from atmospheric nitrogen, the mechanisms for conversion of trace amounts of nitrogenous species bound in the fuel to NOx are not as well understood. 38 In current programs, fuel nitrogen conversion is most

(i) (ii) (iii) (I)

X2+H = HX+X X+X+M=Xz+M which gives H+H


= H 2.

Modeling of hydrocarbon combustion For the halogen acids, the rates of reactions (i)-(ii) are fast compared to other important reactions of H atoms such as H + 02 or H + fuel. This competition for H atoms reduces the rate of chain branching from reaction (1). However, the real key to this type of inhibition is the regeneration of X2 (reaction (iii)) which permits the entire cycle to be catalytic.39'31 Otherwise the inhibitor is rapidly consumed and the overall inhibition is relatively inefficient. Most early modeling of inhibition used HBr, a very effective inhibitor, with H 2 as the fuel. Dixon-Lewis e t al. 3 1 0 - 3 t 2 simulated the inhibition of H2-air flames using a mechanism consisting of the H2-O 2 reactions and H + HBr = H 2 + Br H + Br 2 = HBr + Br Br+Br+M = Brz+M H+Br+M = HBr+M HBr + 0 2 = Br + HO 2. Lovachev e t al. 3 t 3 - 3 1 7 used the above mechanism and added HBr+OH = Br+H20 HBr+O = Br+OH HBr + HO2 = Br + H202 to model inhibition of ignition delays in Hz-air mixtures. Westbrook 39 then used the above HBr inhibition mechanism to model the inhibition of laminar flames in CH4-air and CH3OH-air mixtures. Further additions to the inhibition mechanism CH3Br+H = HBr+CH 3 CHaBr+Br = Br2+CH 3 make it possible to simulate the inhibition of flame propagation by both HBr and CH3Br. 318'319 When CH3Br is the inhibitor, the combined sequence H + CHaBr = HBr + CH 3 CH 3 + Br 2 --- CH3Br + Br H+HBr = H2+Br B r + B r + M = BrE+M again results in the net reaction H + H = n 2. When the halogen acid is the primary inhibiting species, reaction Cycle I predominates, while Cycle II is dominant when the halogenated hydrocarbon is the primary inhibitor. Further extension of the kinetic inhibition mechanism to include HC1, HI, CH3C1 , CH3I, C2H3C1, CEH3Br, C2H3I, C2H5C1, C2HsBr, and C2HsI greatly increases the number of inhibiting species which can be described) 2 In addition to the analysis of inhibition of laminar flame propagation, the complete inhibition mechanism can also be used to examine the effects of inhibitors on fuel oxidation in detonations.49'321 There is an extensive body of experimental data available on the inhibition of low pressure CH4-air (II)

25

flames322-327 and ignition328-33 by CF3Br. Reaction mechanisms have been developed for combustion inhibition by CF3Br in laminar flames and detonations 32'a31 and in shock tube experiments.328 There are two effects which contribute to the relatively large inhibition efficiency of CF3Br. In addition to the catalytic cycle for H atom recombination already observed for HBr, each of the F atoms eliminates an H atom, producing HF which is not further oxidized. The composite reaction mechanism for all the inhibitors mentioned here is included in Appendix II. Inhibition of hydrocarbon oxidation by halogenated species occurs primarily because the catalyzed recombination of H atoms competes with reaction (1). Hydrogen atoms which react with halogenated compounds and are converted to H2 are therefore unavailable for chain branching. There are other important mechanisms which provide alternate reaction paths for H atoms and therefore provide inhibition. For example, we have already observed (Table 1) that rates for reactions between H atoms and hydrocarbons are often much faster than reaction (1). When small amounts of these hydrocarbon species are added to H 2 - O 2 o r H2-air mixtures they inhibit the overall progress of the H 2 oxidation.87 As we discussed earher, the presence of the hydrocarbon provides an effective competitor to 02 for the available H atoms, reducing the chain branching rate and slowing the overall combustion rate. This is a key factor in determining the effect of fuel type on laminar flame speed. Another factor which affects the rate of reaction (1) is the pressure at which combustion occurs. We have already discussed the competition between reactions (1) and (9), in which reaction (9) gradually becomes dominant as the pressure is increased above one atm. As the pressure increases, the effectiveness of flame inhibition by halogenated species also grows, as observed experimentally325,332'333 and predicted by modeling analysis.39,32,331 Both reaction (9) and the normal inhibition cycles compete with reaction (1) at elevated pressures, reducing the chain branching rate even further than at atmospheric pressure. The above inhibition processes have emphasized the importance of H atom removal from the combustion system. However, reductions in the size of the radical pool by other means can also inhibit fuel oxidation. For example, the presence of sulfur dioxide can activate a catalytic cycle resulting in the recombination of O atoms S O 2 + O + M = SO 3 + M
S O 3 -[- O = S O 2 -[- 0 2 .

Modeling analysis of this system demonstrated TM a reduction in flame speed and other changes in the flame structure when small amounts of SO2 were added to C O - H 2 - O 2 - A r mixtures. Further radical recombination paths are provided by SO2+H+M = HSO2+M HSO2 + O H = SO2 + H 2 0

26 and by SO2+O+M = SO3+M SO3+H+M = HSO3+M

C.K. WESTBROOK and F. L. DRYER metric mixtures in air, 95 ~ CH4-5 ~ C2H 6 requires approximately one-tenth the amount of energy for detonation initiation as would be required for a fuel consisting only of methane. 4.3. Simplified Mechanisms We have been emphasizing detailed reaction mechanisms for the combustion of hydrocarbon fuels. However, there is a continuing need for reliable models for fuel oxidation which are very simple and yet still reproduce experimental data over extended ranges of operating conditions. Detailed mechanisms cannot currently be included in most multidimensional combustional models because the computer size, speed, and cost requirements of such a treatment would be too great. Furthermore, we have seen that detailed mechanisms have been developed and validated for only the simplest fuel molecules and are not available for most practical fuels. In addition, there are many occasions where the great amount of chemical information provided by a detailed mechanism is unnecessary and a much simpler model will suffice. The simplest overall reaction representing the oxidation of a typical hydrocarbon fuel is Fuel + n l O 2 --~ H2CO 2 -~-n 3 H 2 0 where the stoichiometric coefficients {hi} are determined by the choice of fuel. This global reaction is often a convenient way of approximating the effects of the many elementary reactions which actually occur. Its rate must therefore represent an appropriate average of all the individual reaction rates involved. The single rate expression is usually expressed
kov = A T " exp ( - E a / R T ) [Fuel]al-Oxidizer]b. (eq. 2)

HSO3 + H = $02 + H 2 0 . These two catalytic cycles are equivalent to H + OH = H 2 0 and O + H + H = H20, respectively. 4.2. Promotion or Sensitization It has been observed that the addition of small amounts of hydrogen or higher hydrocarbons such as ethane or propane will result in a substantial reduction in the ignition delay period for methane oxidation in shock tubes. 335 34.0 Other studies 130'341-344 showed the same type of behavior when the added species was N 2 0 or NO2. Methane is notorious 158 for having an unusually long ignition delay time, much longer than any of the other alkane fuels. This behavior is directly attributable to the difficulty of oxidizing methyl radicals and generating a sufficient radical pool to support the rapid consumption of fuel. Although the earliest studies of kinetic sensitization339 interpreted this process as a thermal mechanism in which the additive was oxidized first, increasing the temperature and reaction rate of the methane mixture, later studies 335 established that the mechanism is primarily kinetic. The hydrocarbon additives produce radical species (predominantly H atoms) much more effectively than does the methane itself, and even a trace amount of additive is sufficient to provide ample quantities of these radicals. When NO 2 or N 2 0 are added, the principal radical species produced initially are O atoms from the dissociation of NO2 or N20. Additional coupling between the additive and the methane oxidation mechanism has also been proposed, 13'341 including the reactions CH3+NO 2 =CH30+NO HO 2 + N O = NO 2 + OH CH4 + NO2 = CH3 + HNO2 H+NO 2 = OH+NO. The sensitization of methane ignition has great practical implications because natural gas consists of methane together with significant amounts (i.e. 5-10~o) of other hydrocarbon species, primarily ethane and propane. Experimental studies of the ignition of natural gas and of methane ethane or methane-propane mixtures 336 338 have shown that the relative concentration of the higher hydrocarbon is a critical factor in determining the ignition delay time. Further studies have shown 48'49 that the ignition delay time can be related directly to the amount of energy required to establish a gaseous detonation wave in a given fuel-oxidizer mixture (see below for further discussion). Therefore the question of susceptibility of natural gas to detonation hazards can be addressed by examining ignition delay times. Modeling studies 128'345'346 have shown that for stoichio-

Although we will discuss only hydrocarbon fuels, with 0 2 as the oxidizer, the same methods are equally applicable to other fuels and/or oxidizers. Comparisons between computed and experimental combustion rates are required to evaluate the constants in eq. (2), including A, n, Ea, a, and b. Note that, in general, a and b are not related to the stoichiometric coefficients in the global reaction, but are empirical. constants determined either from experiment or from detailed kinetics calculations. As we have seen earlier, different elementary reactions are important in different experimental conditions, so it can be expected that the constants in eq. (2) will vary with fuel type, pressure, equivalence ratio, and other parameters as well. In spite of these difficulties, this type of rate expression is probably the most commonly used combustion rate in the literature. In a great majority of cases, it is assumed that the overall reaction is first order with respect to both fuel and oxidizer, so that a = b = 1. However, recent modeling of laminar flame propagation using simplified rate expressions has shown a47 that this choice of reaction order can lead to serious errors. If it is assumed that n = 0, Ea = 30 kcal/mol, and a and b are assigned values, then the pre-exponential

Modeling of hydrocarbon combustion 70


I I [ I

27

kov = 1.15 X 10 TM exp( 30/RT)[C8H18] 1"0102 ] 1.0

50 --

/ f ~

" " " - " ""'----~

--

30-- u .

E
20

/////

~kov=4.eXlO
~

11 exp(-30/RT)[CaH 18 ] 0.25 [O211 5


. .

0//
o

/ /

~'L

J
1

I
2 Equivalence ratio -

I
3

I
4

~h
5

FIG. 7. Variation of laminar burning velocity with equivalence ratio for n-octane in air, computed using the single-step reaction rates indicated. Experimental values for the lean and rich flammability limits (~bL= 0.5 and ~bR= 4.3), for the observed burning velocity at q~ = 1 (open circle), and for the maximum burning velocity (open square) are also indicated. term A can be chosen so that the computed flame speed for a given fuel-air mixture correctly reproduces the known experimental value. F o r the case of noctane in air, the laminar flame speed is approximately 40 cm/sec for a stoichiometric mixture at atmospheric pressure. When A is evaluated, the resulting rate expression is flammability limit t~L and rich limit (~R are not correctly predicted. F o r example, tkR "" 4.3 but eq. (3) predicts a rich hmit at ~b = 10. These errors could be quite serious for simulations in which the local equivalence ratio varies with space or time. Parametric variations for the constants in eq. (2) led 347 to a set of values which correctly reproduces the maximum flame speed and the equivalence ratio at which that maximum occurs, as well as the proper lean and rich flammability limits, together with the correct variation of flame speed with pressure. F o r noctane, these results are obtained with kov = 4.6 x 1011 exp(_30000/RT)[CsHls]O.25[O211.5 (eq. 4) Predicted results using eq. (4) are also indicated in Fig. 7. Similar rate expressions for other c o m m o n hydrocarbon fuels are summarized in Table 4. The single-step mechanism predicts flame speeds

kov = 1.15 x 1014exp(- 3 0 0 0 0 / R T ) [ C s H l s ] I ' [ O 2 ] 1"


(eq. 3) which correctly computes a flame speed of 40 cm/sec. However, when the equivalence ratio is then varied, computed flame speeds do not agree with experimentally observed results. This is illustrated in Fig. 7, showing the results computed using eq. (3) as the dashed curve. The predicted maximum flame speed is about 55 cm/sec and occurs at an equivalence ratio ~b -~ 2. Experimentally, the maximum flame speed is about 42cm/sec at ~b -~ 1.15. 348 In addition, the lean

TABLE 4.

Fuel
CH 4

A 1.3 109 8.3 105 1.1 x 1012 8.6 x 1011 7.4 1011 6.4 x 1011 5.7 1011 5.1 1011 4.6 x 1011 7.2 1012 4.2 1011 3.8 x 1011 3.2 1012 1.5 1012 2.0 x 1011 1.6 x 1011

Ea 48.4 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 40.0 30.0 30.0 30.0 30.0 30.0 30.0

a -0.3 -0.3 0.1 0.1 0.15 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.15 -0.1 -0.1

b 1.3 1.3 1.65 1.65 1.6 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.6 1.85 1.85

q~'L 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5

q~'e 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5 0.5

~bL 1.6 1.6 2.7 2.8 3.3 3.6 4.0 4.5 4.3 4.3 4.3 4.2 4.1 3.4 3.4 3.2

~bR 1.6 1.6 3.1 3.2 3.4 3.7 4.1 4.5 4.5 4.5 4.5 4.5 4.0 3.6 3.6 3.5

CH 4
C2H 6

C3H s
C4Hlo

C5H12
C6H14

C7H16 CaHx8 CsHIa


C9H20

CloH22 CHaOH C2HsOH


C6H 6

C7H 8

28

C. K. WESTBROOK and F. U DRYER

TABLE5. Detailed mechanism Tad 0.8 1.0 1.2 1990 2220 2140 [CO]/[CO2] [H2]/[H20] 0.03 0.11 0.69 0.005 0.02 0.15 One-step mechanism
Tad

Two-step mechanism
Tad [CO]/[CO2]

2017 2320 2260

1975 2250 2200

0.08 0.14 0.43

reliably over considerable ranges of conditions, but it has several flaws which can be important in certain applications. By assuming that the reaction products are C O z and H 2 0 , the total extent of reaction will be overlaredicted. At adiabatic flame temperatures typical of hydrocarbon fuels ( ~ 2000 K), substantial amounts of C O and HE exist in equilibrium with COz and H / O in the combustion products. The same is true to a lesser extent with other species such as H, O, and OH. This equilibrium lowers the total heat of reaction and the adiabatic flame temperature below the values predicted by the single global reaction. In Table 5 this effect is illustrated for the case of C H 4 - a i r mixtures. Predicted adiabatic flame temperatures from a detailed mechanism are compared with those computed from a single-step model. Also shown are the relative amounts of C O and H z from the detailed model. The overestimate of Tad from the single-step model grows with increasing equivalence ratio and is directly related to the increasing amounts of C O and H 2 in the reaction products. This deficiency can be circumvented by using the one-step model to define the rate of reaction, but using equilibrium thermochemistry to establish the final product distribution and temperature. In addition to the effects of equilibrium composition in the product gases, single-step mechanisms also neglect the fact that typical hydrocarbon fuels burn in a somewhat sequential manner. That is, the fuel is partially oxidized to C O and Hz, which are not appreciably consumed until all of the hydrocarbon

species have disappeared. 349 Dryer and Glassman 37 used this observation to construct a two-reaction model for methane oxidation in a turbulent flow reactor
CH4+302 = CO+2H20

(eq. 5) (eq. 6)

CO +10 2 = CO 2

with empirically determined rates for both steps. The same two-reaction approach has been u s e d 347"35'351 for other hydrocarbon fuels. The general formula for a hydrocarbon fuel is C,H.Or + n+~m_2 ~ O2=nCO+ H20 (eq. 7)

together with eq. (6). A reverse rate for the C O oxidation step can be introduced to account for the variation of extent of reaction and heat release with equivalence ratio and the effects of pressure variation. The resulting rate expressions for c o m m o n hydrocarbon fuels are summarized in Table 6. In the case of methane-air, the improvements in predicted adiabatic flame temperature and the computed C O levels are indicated in Table 5. Another reaction can be added to account for the H 2 - H 2 0 equilibrium in the product gases. F o r each reaction added, the equilibrium concentration of another product species can be estimated. A logical extension of this process is the quasi-global reaction mechanismZSZ-3S7 which combines a single reaction of fuel and oxygen to form C O and H2, together with a detailed mechanism for the C O - H 2 - O 2 system.

TABLE 6.

Two-step mechanism Fuel CH 4 CH 4 C2H 6 Call 8 C4H1o


C5H12 C6H14 C7H16

Quasi-global mechanism b 1.3 1.3 1.65 1.65 1.6 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.6 1.85 1.85 A 1.1 X 109 6.9 x 105 9.2 x 1011 7.2 x 1011 6.2 x 1011 5.4 x 1011 4.8 x 1011 4.3 x 1011 3.9 x 1011 6.0 x 1012 3.5 x 1011 3.2 x 1011 2.7 x 1012 1.3 x 1012 1.7 x 1011 1.3 x 1011 E,
48.4

A 2.8 X 109 1.5 x 107 1.3 x 1012 1.0 X 1012 8.8 x 1011 7.8 x 1011
7.0x 1011

Ea

a
-0.3

b
1.3

Call18 CsHls
C9H2o

C10H22 CHaOH C2HsOH


C6H 6 CTH 8

6.3 x 1011 5.7 x 1011 9.6 x 1012 5.2 x 1011 4.7 x 101 x 3.7 x 1012 1.8 x 1012 2.4x 1011 1.9 x 1011

48.4 30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 40.0 30.0 30.0 30.0 30.0 30.0 30.0

-0.3 -0.3 0.1 0.1 0.15 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.15 -0.1 -0.1

30.0 30.0 30.0 30.0 30.0 30.0 30.0 30.0 40.0 30.0 30.0 30.0 30.0 30.0 30.0

-0.3 0.1 0.1 0.15 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.25 0.15 -0.1 -0.1

1.3 1.65 1.65 1.6 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.6 1.85 1.85

Modeling of hydrocarbon combustion Another extension is the partial oxidation of fuel to intermediates which are somewhat larger than CO and H2, which can be further oxidized by a detailed mechanism. The mechanism discussed earlier for nheptane oxidation 2so adopted a form of this approach, simplifying the reaction sequence until C2H4, CO, and other Ca and C2 hydrocarbons are formed, then dealing with their combustion with a detailed model. Another variation of this approach is to simulate the oxidation of a complex fuel by a sequence of semi-global steps, 35t,3ss without including any detailed mechanism. For example, such a sequence for ethane359 consists of
C2H6 + 0 2 -* C2H , + H 2 0

29

when transport terms are not included and coupled PDE's when spatial transport is included. Wbile relatively fast and inexpensive software packages are available to solve stiff ODE's, solution methods for the PDE's are much more difficult to develop and their computational costs are substantially greater. 5.1. Kinetic Models without Transport Several important combustion systems can be considered to be spatially homogeneous, so that transport effects may be neglected. Others, including plug flow reactors, interchange spatial coordinates with time but are still described by ODE's. Historically, much of the development of detailed kinetic mechanisms has taken place in these systems. 5.1.1. Shock tubes The earliest and most common research area for the development and application of kinetic modeling has been in shock tube experiments. There is an extensive literature dealing with this subject and we cannot hope to summarize it here. Much of this work is cited according to the types of fuel considered in the first half of this review. However, some general observations can be made with respect to the ways in which these experiments and their models interact. Models of combustion in shock tubes assume that the reactive region behind the shock wave is not affected by diffusional transport of mass or energy. For incident and non-ideal reflected shocks, the shocked gas has a net translational flow rate which must be taken into account when comparing experimental and computed results, but the dominant factor involved is the elapsed time subsequent to the shock arrival. Experimental features such as observational slit width, boundary layer growth, shock curvature and other factors can introduce uncertainties into both observed and computed results. 36 However, within the limits set by these considerations, it is possible to carry out very careful modeling studies of fuel pyrolysis and oxidation in shock tubes. Although there are many exceptions, it is possible to outline some of the general characteristics of the majority of shock tube experiments and the models describing them. Most often, the fuel or fuel-oxidizer mixtures are substantially diluted by argon, helium, or nitrogen. This dilution reduces the temperature change brought about by the heat of reaction; it also can simplify the kinetic mechanism needed to describe the reaction. A good example of both features is provided by the study of ethylene pyrolysis by Just et alfl TM discussed earlier (Fig. 6) in which the fuel was present in very low relative concentrations, between 20 and 800 ppm. The temperature of each experiment was essentially constant, and only three reactions were found to be important, C2H a + M = C2H 2 + H 2 + M
C2H4+M = C2H3+H+M

CEH 4 + 202 --' 2CO + 2H20


CO-~- 21-O2 ~ C O 2.

For each species which is added to a simplified mechanism, the computational costs increase. In multidimensional models where these costs are a limiting feature, the mechanism should be as small as possible. F o r such problems the two-step mechanism probably presents the greatest advantages and the fewest drawbacks. The true quasi-global mechanisms which include detailed treatments for CO, H 2 or even C 2 species oxidation, can be nearly as large and costly as a full detailed mechanism. The real advantage provided by quasi-global models is their ability to treat arbitrary fuels. Regardless of the complexity of the fuel molecule, an overall reaction can always be written, so the oxidation of that fuel can be simulated. As the above discussion shows, this simulation can be very acceptable if care is taken in the selection and validation of the simplified rate expression to be used. However, there are several limitations of the quasiglobal modeling approach which must be kept in mind. Quasi-global models developed for shock tube ignition delays are not generally suitable to model flame speed or reaction rates in plug flow or stirred reactors. Also, quasi-global models generally do not provide a good prediction of radical species and hence may not be useful in modeling NO x formation or inhibition, where radical effects are very important. Although it has not often been carried out, 351'357 detailed kinetics calculations can be used to develop quasi-global models for fuels where detailed mechanisms are available. The detailed models can then be used to assess quantitatively the errors resulting from the quasi-global model. Even when the quasi-global mechanism is derived from the detailed mechanism, however, the parameter ranges over which the simplified mechanism can accurately be applied are rather narrow. 5. APPLICATIONS Models in which detailed chemical kinetic reaction mechanisms are used can be subdivided on the basis of whether or not spatial transport effects are considered, which determines the type of differential equations to be solved. These consist of coupled ODE's

(74)
(75)

C2H3 + M = C2H 2 + H + M.

(77)

30

C.K. WESTBROOKand F. L. DRYER

C2H 4 + H = C2H 3 + H 2 1013

[72]

[24]

[3 [121

~7 [42]
[27] [28] [87]

A -~

[881
[89]

[90]
This study

%
o~ 1012

io ~

7.5 104/T

i.

10

12.5

FIG. 8. Rate expressions for C 2 H 4 + H = C2Ha+H2, using experimental values from the references cited by Westbrook et al. 45

The temperature ranges encountered in shock tube experiments are quite wide, from a low of 1100-1200 K to a maximum of 3000 K or higher. The great majority of these lie between 1400 K and about 2100 K. As we will see below, this temperature regime is considerably higher than that in which fuel consumption occurs in flames or plug flow reactors. Therefore shock tube simulations provide valuable information on high temperature rates of important reactions. When this is combined with other data at lower temperatures, many reactions display substantial non-Arrhenius temperature dependence, as illustrated in Fig. 8 for the reaction
C z H 4 + H = C z H 3 + H 2.

(79)

The high temperature data (T > 1300K) were all obtained in shock tube modeling studies and the rates are considerably faster than would have been obtained by extrapolation of the lower temperature (T < 1300 K) data. Overall correlations of shock tube ignition delay times are often derived from experimental data, and such relations are a form of global rate expression for the reaction of that particular fuel. However, the overall reaction order in shock tube experiments can vary with fuel concentration, with fuel-oxidizer equivalence ratio, and with temperature, making it difficult to extrapolate these correlations to parameter regimes not specifically examined by the experiments. Several characteristic features can be identified in the computed time history of a typical shock tube experiment. At the very beginning the reactants, which

have been prepared and kept at low pressure and temperature, can be assumed to consist entirely of the initial species. Depending on the details of the numerical solution being used, initial concentrations of product and intermediate species can be exactly zero or some small (e.g. 10-2mol fraction) vahle. Because of the nearly instantaneous change in thermodynamic state which the shock wave provides, the post-shock gases are also initially unreacted. Usually this post-shock state is chosen as the initial condition for the calculations. Boundary conditions for the computed results depend on the type of shock tube experiment being examined. The shock wave fluid mechanics conservation equations (Hugoniot relations) can be solved, coupled with the chemical kinetic rate equations. Simplifying assumptions concerning the fluid mechanics can also be made. F o r incident shock waves the assumption of constant pressure in the post shock gas is reasonably accurate during the induction period, while constant volume conditions are appropriate for reflected shocks. These conditions apply strictly only until the gas reactions begin to release energy and pressure gradients begin to appear. Thermal decomposition reactions and, in the case of oxidizing environments, reactions between fuel and 0 2 molecules are the initiation reactions in the shocked gas. Reactions in shock waves constitute the only combustion environment in which these initiation reactions are very important, so that simulation of such experiments provides a valuable means of isolating and studying these reactions. Heffington et

Modeling of hydrocarbon combustion TABLE7. Initial conditions for shock tube calculations Mixture 1 2 3 4 5 6 7 8 CH3OH mole fraction 0.02 0.01 0.0075 0.01 0.01 0.02 0.04 0.0075 Equivalence ratio 0.75 0.75 0.75 1.50 0.375 3.00 6.00 0.75
I I

31

10-8

al. 147 reviewed shock tube studies of methane oxidation, emphasizing the role of the thermal decomposition reaction (40). Gardiner and co-workers have examined the initiation reactions for several hydrocarbon fuels 161'184-186 by means of model simulations, and Dean et al. 11o,1 ~ evaluated the rate of the formaldehyde decomposition reaction (30), We can illustrate the process for the case of reaction (141) involving CH3OH. In a modeling study of methanol oxidationff 3 reflected wave shock tube data of Bowman 17 were used to validate the reaction mechanism. These experiments considered C H 3 O H - O 2 - A r mixtures at initial temperatures between 1545 K and 2180K, with fuel-O 2 equivalence ratios between 0.375 and 3.0. The observed induction times could be correlated with initial temperature and reactant concentrations using the expression

[]

d
[] [] C~ A Mixture Mixture Mixture Mixture Mixture Mixture 1 2 3 4 5 6

10-9 0.5 10 3 K 0.6 0.7

z[O2]o'5[CH3OH] 0"1 = 2.1 x 10-13 exp ( 1 8 2 1 4 / T ) s e c ( m o l / c m 3 ) '6. These experiments were simulated by the model, using the detailed reaction mechanism. The initial conditions for each set of mixtures are summarized in Table 7, and the computed results are shown in Fig. 9. The experimental correlation function is indicated as the solid line. The model results are very consistent with the correlation function, using a rate expression for reaction (141) k14 l = 3 x 101 a exp ( - 80000/RT) cm3/mol sec. The activation energy of 80 kcal/mol lies between the value of 67.7 kcal/mol determined by Bowman from a less detailed modeling analysis of his own experiments, and a value of 91 kcal/mol from flow reactor pyrolysis modeling by Aronowitz et a/. 26 If k~4~ is replaced by alternative expressions having these other values for E~41, and with all of the other rates unchanged, the model calculates induction times which are significantly different from those observed experimentally. It is clear that the overall activation energy for the induction period (~36kcal/mol from the experimental results) predicted by the model depends directly on the activation energy of the initiation reaction. This determination is, however, a function of the rest of the reaction mechanism. As we have already observed, Bowman derived a value for E14 l of 67.7 kcal/mol using a different mechanism. In addition, the rates of reaction (141) and other decomposi-

FIG. 9. Shock tube ignition delay time correlations, showing computed values of the correlation function for each mixture (see Table 7). The straight line is the best fit to the experimental results,27

tion reactions can be affected by pressure falloff which can also change their apparent activation energy. During the initiation period, the total amount of fuel consumed is quite small and the temperature remains effectively constant or even declines slightly due to the endothermicity of most of the initiation reactions. The most important event during this period is the establishment of a radical pool. In many cases, the initiation reaction produces one small radical such as H or OH, and one large radical such as CH3, CH2OH, or C2H5 (e.g. reactions (40) and (141)). In other cases two large radicals are produced (e.g. reactions (70) and (216)). The large radical often decomposes further, resulting in another small radical and a net production rate of two radicals for each fuel decomposition step. Methyl radicals, however, do not decompose directly to any significant degree, recombining to produce C2H6 or reacting instead with other radicals and with 0 2 as discussed earlier. Therefore, in the case of methane decomposition, fewer small radicals are produced for each fuel molecule that decomposes than when other fuels such as ethane or methanol are used. This smaller initial source rate for radical species in methane initiation persists throughout the subsequent stages of the induction period and is a key factor in explaining why methane has an anomalously long induction period, compared with other alkane fuels.

32

C.K. WESTBROOK and F. L. DRYER tion is effectively inhibited until most of these hydrocarbons have been consumed. For the same reason, O, H, and OH concentrations remain low until hydrocarbon species are gone, whereupon they rise rapidly to levels which are characteristic of CO oxidation. The same four general reaction phases can be identified in most combustion systems. In some environments one or more of these phases can be dominant while others may effectively disappear. In laminar flames and other systems in which spatial transport is important, these phases may be observable at the same time in different regions of the flame. Different experimental regimes are appropriate for studying different oxidation phases. We observed earlier that initiation reactions were negligible except in reacting shock waves. However, because of the speed of the overall reaction in the shock tube which results from the high pressures and temperatures, available diagnostic techniques often cannot resolve many of the details of the combustion process. In most cases only the total ignition delay time has been reliably measured. Single pulse shock tube experiments are sometimes carried out where an expansion wave quenches the reaction and makes some concentration measurements possible during the ignition period, and other techniques can detect some major species concentrations during the reaction, but relatively little species information is available from most shock tube experiments. Because in most cases only induction times are available for comparison with model predictions, shock tube simulations are not a particularly demanding test of a given reaction mechanism, although they can indicate cases in which a given mechanism is inadequate. 5.1.2. D e t o n a t i o n s Gaseous detonation waves have been studied experimentally for many years. Detonation limits, propagation rates, and initiation properties have been examined for many fuel-oxidizer mixtures. Theoretical descriptions of detonations have appeared recently, showing how hydrodynamic and kinetic processes interact in detonation waves. The weakest parts of existing detonation models have been their submodels for chemical kinetics. Since induction times play a prominent role in all of these models, the lack of reliable kinetics models has been a serious problem. The first applications of detailed models to compute induction times in detonation conditions and apply the results to predict detonation phenomena involved H 2 - O 2 systems. 7-73 These studies were concerned primarily with understanding the interactions between kinetic processes and the cellular structure of planar detonations in linear tubes. More recently, Atkinson e t al. 68 used a detailed hydrogen oxidation mechanism to relate kinetic parameters to the problem of initiation of unconfined spherical detonations in H2-air mixtures by high explosive charges. Oran e t a/. 69'74-76 also used a detailed H 2 - O 2 mechanism to examine the kinetic factors in reactive shock waves. In all of these cases, the relative simplicity of the H 2 - O 2

Following the brief initiation step, there is a rather long period during which most of the radical species concentrations increase exponentially. Some are produced by reactions between fuel and other radical species, but the exponential growth is accomplished by chain branching steps, particularly reaction (1) H+O 2 = O+OH. (1)

At the same time the concentrations of intermediate species including HE and CO are growing. These species cannot be further oxidized appreciably as long as the fuel and other hydrocarbon species are present because the OH concentration is suppressed by high rates of reaction with the hydrocarbons (see Table 2). As the radical species concentrations continue to grow, the characteristic fuel consumption time scales become shorter until the fuel finally disappears. This releases the constraint on the OH level and reactions (3) and (28) H2+OH = H20+H CO + OH = CO 2 + H (3) (28)

rapidly oxidize CO and Hz. The heat released by these reactions accelerates the process, as does the chain branching provided by the product H atoms with 02 (reaction (1)). The C O - H 2 oxidation phase can therefore be very rapid. Once the majority of the CO and H2 are oxidized, the final phase occurs, a much longer period characterized by the approach of the entire system to chemical equilibrium. The duration of this period is determined by the rates of the important termolecular radical recombination reactions, including H+H+M O+O+M H+OH+M = H2+M = 02+M = H20+M (5) (6) (8) (26)

CO + O + M = CO 2 + M.

Since these reactions depend on the cube of the density this phase is much slower than the oxidation reactions. In summary, for hydrocarbon oxidation in shock tubes the entire reaction is divided roughly into four phases. These consist of initiation, fuel consumption, CO oxidation, and equilibration. The first two phases may overlap, depending on the fuel considered and the initial temperature and pressure. However, the CO oxidation phase is usually quite distinct. Much of the heat release of the overall reaction is associated with the CO oxidation. Thus the rapid temperature and pressure transients which are often used to define the end of the ignition delay period also indicate the onset of CO oxidation. The duration of the overall reaction consists primarily of the time required for initiation and fuel consumption, with CO oxidation usually taking less than 10~o of the total ignition delay time. Since the rates of reactions between radicals and fuel or intermediate hydrocarbons are much higher than the rates of reactions with CO, CO oxida-

Modeling of hydrocarbon combustion mechanism and the availability of reliable rate data for the major reactions were responsible for the early use of modeling analysis for these problems. The slightly more complex C O - H 2 - O 2 mechanism has been used 36a'362 to examine in great detail the reignition processes within a single detonation cell. In these very important studies, it was found that because the primary shock wave decays as it travels through a cell, the induction process in the last third of the cell is far too slow to accomplish the required re-ignition. However, additional shock heating and compression provided by multiple Mach stems from waves reflected from the detonation tube walls effectively shorten the ignition delay time enough to achieve re-ignition at the end of the cell. In the above modeling studies, the principal goal was an improved understanding of fundamental detonation phenomena. The fuels were selected because they were known to support stable detonations under safe laboratory conditions over a range of fueloxidizer equivalence ratios, and the reaction mechanisms were reliable. F o r some time, modeling analysis of detonation processes with hydrocarbon fuels was limited by the lack of reliable kinetics submodels. 193'363'364 Recently, however, detailed kinetic mechanisms have been applied to the analysis of detonation waves in hydrocarbon-oxidizer mixturesff 8'49'195'321 With a single model, available experimental data for the influence of fuel-oxidizer equivalence ratio, initial temperature and pressure, and oxidizer type (i.e. 02 or air) on observed detonation parameters can be predicted with reasonable accuracy. The general model used is the Zeldovich-von Neumann-Doring (ZND) model 365-367 in which, locally, a detonation consists of a shock wave traveling at the Chapman-Jouguet (C J) velocity, followed by a reaction zone. The conditions behind this leading shock wave, called the von Neumann spike, are considerably different from the final CJ conditions. In particular, the pressure and density are much higher than the CJ values. The shock wave compresses and heats the fuel-oxidizer mixture which then begins to react. In most mixtures, the overall fuel oxidation consists of a relatively long induction period during which the temperature and pressure of the gas mixture remain roughly constant, followed by a rapid release of chemical energy and temperature increase. In actual detonations, this temperature increase is accompanied by a substantial decrease in the gas density and pressure due to the expansion of the product gases, leading eventually to the final CJ state. For each combustible mixture, a calculation is first made of the relevant CJ conditions, which depend on the initial pressure, temperature, and composition of the reactant mixture. F r o m the resulting value for the detonation velocity Dcj, the conditions of the von Neumann spike, including the temperature T1, pressure Pa and particle velocity vl of the post-shock, unreacted gases can be calculated from the shock conservation relations and then used as initial conditions
JPECS IO:I-C

33

for the chemical kinetics model. It is well established that the local shock velocity varies within a single detonation cell from a maximum initial value of from 1.2 to 1.6 times Dcj to a minimum of 0.6 to 0.85 times Dcj , so the CJ conditions represent average values, and the computed induction times will also be averages. This simplified model does not attempt to follow the entire history of the detonation wave, which would require a coupled multidimensional fluid mechanicschemical kinetics treatment. Instead, the reactive mixture is assumed to remain at a constant volume over its reaction time, and the induction time is defined in terms of its temperature history. Typical mixtures experience a large temperature increase of 12001800 K, and the induction time is defined as the time of maximum rate of temperature increase. In most cases this coincides approximately with the time at which the temperature has completed about half of its total increase. This is not, strictly speaking, a true induction period, often defined as the time required for a small (i.e. 1-5 ~ ) temperature or pressure increase, but here the release of macroscopic amounts of energy is of primary interest. Recently, Lee eta/. 368 have extended this approach, including the effects of time-varying density and pressure in the post-shock unreacted gases. In addition to the induction time T it is useful to define the induction length A = r ( D - V l ) where ( D - v l ) represents the particle flow relative to the shock. Atkinson et al. found that the variations in Dcj and vl with equivalence ratio q~ were sufficiently large that A and A 3 correlated better with experimental data than did z and z3. As a result of the simplifications outlined here, the computed induction times and induction lengths must be considered as characteristic time and space scales, not as precise descriptions of the state of gas elements through the detonation front. The evolution of the reacted gas subsequent to the induction period is dominated by the fluid mechanics of the post-reaction expansion of the reaction products. The expansion reduces both the pressure and density of these products and therefore alters the kinetic equilibrium state. Since virtually all of the reactants have been consumed by this time, the kinetics of this phase are controlled by relatively slow radical recombination processes and not by oxidation reactions. This model of the detonation kinetics is a simple one and neglects several potentially significant effects arising from hydrodynamic-kinetic interactions. Variations of density, temperature, and particle velocity in the post-shock unreacted mixture are not generally considered and can affect the induction period in a number of ways. Multiple shock wave reflections, rarefactions, interactions with confining walls, cellular structure, and other related effects are also not treated by the simplified model. A really complete detonation simulation model will have to include such interactions and at least two and probably even three space dimensions, but such a model is beyond the scope of current modeling capabilities. However,

34

C.K. WESTBROOK and F. L. DRYER

the progress to date shows that current kinetics mechanisms have reached the point where chemical kinetics is no longer the weakest part in the theoretical descriptions of detonation phenomena. Once the induction lengths are computed, they can be used to correlate experimental data from a variety of sources. Strehlow and co-workers 369 371 found that the ratio of induction length to transverse wave spacing was roughly constant for a given fuel and oxidizer as the equivalence ratio and/or pressure were varied. It has been observed that the transverse wave spacing increases as the lean and rich limits of detonation propagation are approached. At either limit, single-spin detonation occurs with a transverse wave spacing equal to twice the tube diameter. These facts can be combined to predict detonation limits in a variety of fuel-O 2 and hydrocarbon-O 2 mixtures.48 50.89.197,321 Another practical detonation problem concerns the initiation of an unconfined spherical detonation in a given fuel-oxidizer mixture. 372 One experimental technique which has been used extensively to study this problem is the use of a planar detonation emerging from the end of a linear tube. 373 -376 This work has shown that if the tube has a diameter less than a critical value de then a spherical detonation wave will not be established. For tube diameters larger than d~ a stable spherical detonation wave will result. Lee and co_workers377 380 have shown that dc is proportional to the induction length A for many fuel-oxidizer mixtures, and Westbrook e t al. 48 50,89.197 showed that this proportionality holds for a number of hydrocarbon fuels and for H2, using oxidizers ranging from 0 2 (fl ~ N 2 / O 2 : 0 ) to air (fl = 3.76), as long as a reliable comprehensive hydrocarbon oxidation mechanism is used. These results are summarized in Fig. 10, showing the variation of induction length A with amount of N 2 dilution in given fuel-O 2 mixtures. Also shown as symbols are available experimental data relating d~ to N 2 dilution for the same mixtures. The two scales are related by the equation dc = 380A. (eq. 8)

|~7 1 I-BI |FI

/"

CH4I(CO2) CH4 (CO) C2H6 (CO2) C2H6 (CO) C2H4 (CO2)

IZ&C2H4 (Ico) I 102 ~, C2H4 (CO2) C2H2 (CO2) | O C2H2 (CO) X C2H2 (CO2)/~ 10

= 10-

.~

1ff4"~ - A~/

"-t...x " ' X ~ O / , , ~ -

lO-5~ f / 0 J//

Model (CO2)

1
0

f
1

6
2

~
3
Air

10-3

FIG. 10. Induction length for fuel-O2-N2 mixtures, for (from top to bottom) CH,, C2H6, CzH , and C2H 2. Also shown as individual symbols are experimental data for critical tube diameter for initiation of unconfined spherical detonation. Data from Moen et al.378 are indicated by + symbols, from Zeldovich et al. T M by symbols and all other data from Matsui and Lee.377

lengths for H 2 air mixtures, using a detailed kinetic model. Bull e t a/. 363 extended the same approach to C2H6-air mixtures but used a global induction time correlation for computing the induction times. Westbrook 48 then used a detailed kinetic model to correlate the experimental data of Bull e t al. as shown in Fig. 11. The agreement between computed values of A3 and critical H.E. charge mass is excellent, with Mc = 2.5 x 105A3 kgTetryl. Similar agreement was demonstrated for other hydrocarbon fuels, including CH4, C2H4, C2H2, and H 2.

Although not shown in Fig. 10, the same equation also correlates data for H2-O 2 N 2 mixtures equally well. 89 The other widely explored technique for initiation of unconfined detonation involves the use of high explosive charges within the fuel-oxidizer mixture.193 For a given fuel-oxidizer mixture there is a critical energy Eo (or equivalently a critical high explosive mass M~) below which a detonation cannot be established directly. Zeldovich e t al. 381 showed that this critical energy could be related to the induction length A by E~ = koAj (eq. 9)

[
~E i 10-6
[nation]

/D L

I -1

--I~=

[ Deto~ ~(~

No detonation

} _
~ ~c 10-7 0 1
I

- 1o-,
C2 H6 - Air Bull et al (1979) O Detonation No detonation 2
I

.C "~ o
10-2

wherej --- 1, 2, or 3 for planar, cylindrical, or spherical configurations respectively, and ko is a constant of proportionality. Atkinson e t al. 68 showed that this equation accurately related computed induction

Equivalence ratio

FIG. ll. Cube of induction length for ethane-air mixtures, with data from Bull e t a / . 363 for critical energy for initiation of spherical detonation.

Modeling of hydrocarbon combustion Predictions were also made for ignition of C H 3 O H oxidizer mixtures, although experimental data are not available for this fuel. Detonation properties of CH3OH appear to lie roughly between those of C2H4 and C2H 6. Westbrook and Urtiew 49'89'321 then showed that the variations of the critical initiation energy, the critical tube diameter and composition limits to planar detonation in tubes with initial pressure and temperature could also be predicted reliably by the above kinetic modeling approach. It was also shown that halogenated species could inhibit detonations by increasing the energy required for initiation of detonation and by narrowing the composition limits. The application of this type of analysis to the question of detonability of natural gas was discussed earlier. With the growing use of liquified natural gas (LNG), there is an increasing chance of a major accident involving spills which would result in a large atmospheric cloud of natural gas and air which might be flammable and/or detonable. Modeling studies 12s'345'346'382 agree that although unconfined, spherical methane-air mixtures are difficult or virtually impossible to detonate, detonations in mixtures of methane and ethane characteristic of natural gas can be initiated by sufficiently large high explosive changes in controlled experiments. Susceptibility of natural gas-air mixtures to detonation in actual spill scenarios is a much more difficult problem which depends on many factors, some of which cannot easily be quantified, but information provided by modeling studies and controlled experiments has been an important part of hazard assessment studies. 383'384 5.1.3. Plug flow reactor An important class of combustion systems are collectively called plug flow reactors. These combustors are characterized by high linear flow rates with negligible recirculation flow. Each element of gas reacts as it moves, with the characteristic time scale for heat and mass transfer by diffusion being much longer than that for convective motion. Models of these systems consider a control volume as it evolves in time, converting the time co-ordinate to a space coordinate for comparison with experimentally measured concentration and temperature data. Many recent modeling studies have simulated combustion in plug
flow reactors. 38'43'~5'195'26'226'227'235'358'385 -387

35

Combustion takes place at constant pressure, most often near atmospheric pressure. The flow is assumed to be linear, and all transport normal to the flow axis is neglected. This included heat losses to the walls of the flow duct which are usually heated to the same temperature as the inlet flow. Care is taken to achieve complete and rapid gas mixing at the inlet or source end of the reactor, and this mixing process is assumed to persist throughout the reactor. Combined with the neglect of radial transport and recirculation, steady plug flow is achieved. Further details are available elsewhere. 3 7,3 8 8 The length of the typical reactor is of the order of one meter, and the inlet flow velocity is adjusted so that the reaction zone of interest is spread over this distance. The residence time in the reactor is typically about 0.1 sec. At temperatures around 1100 K, characteristic initiation time scales are of the order of a millisecond. This means that regardless of the purity of the initial reactants, many other species are quickly created at the very beginning of the flow. Most important, the appropriate radical pool is formed within the first few centimeters in the reactor duct. The important strength of the flow reactor as a research tool is the access it provides to the primary fuel reaction and CO oxidation zones. By spreading these regimes over the space of a meter or more, and by operating at steady state, experimental access by either optical or sampling probe methods is very convenient. For illustration, we will again draw on our earlier study of methanol oxidation, 43 using experimental data obtained by Aronowitz et al. 222 in the Princeton University turbulent plug flow reactor. Other experiments using the same apparatus and modeling analysis were described earlier. Experimental results for concentrations of CH3OH, CO, CO2, and H2, as well as temperature are plotted as open symbols in Fig. 12. These data are shown as functions of axial distance along the cylindrical reactor duct. In addition, small amounts ( ~ 1000ppm) of C H 2 0 and trace amounts ( ~ 5 0 p p m ) of CH4, C2H2, C2H4, and C2H 6 were also detected. Attempts to simulate the results with previously existing methanol oxidation mechanisms were unsuccessful, leading to the development of a new mechanism as described earlier. For this particular set of data (~b = 0.05, TO = 1021 K), the fuel consumption was dominated by reaction (145) CH3OH + OH = CH2OH + H20. (145)

In typical flow reactors, dilute fuel or fuel-oxidizer mixtures are considered in order to keep the total temperature variation quite small. The most common temperature range is between 900K and 1300K. An important feature of the flow reactor is that it occupies a regime roughly midway between static reactor and shock tube temperatures. As we shall see later, most fuel and intermediate hydrocarbon consumption in flames takes place in the same temperature range as that encountered in flow reactor experiments. Therefore these results are of great value in obtaining information required for the analysis of most practical combustors.

Requiring the model prediction of the CH3OH profile to match the observations led to a determination of a rate for reaction (145). Reaction (144) CH3OH+H = CH2OH+H 2 (144)

was responsible for most of the predicted H 2 levels. In an accompanying fuel-rich case (~b = 1.5) reaction (144) was found to account for most of the methanol consumption as well. The combination of the lean and rich simulations permitted a determination of k144. In addition, reaction (150)

36
8 i

C.K. WESTBROOK and F. L. DRYER


i

1140

o x

6
o

1110
I

g
~= 4 -6 E

1080

g2
o 8
I I I

1050

1020 4

I
I

I
I

I
I

[]

6 g
~ E 4

o
x

~
-6 E

d
I 20 1 40
Position -- cm

z~
I 60

0
80

20

40 Position - cm

60

80

FIG. 12. Lean flow reactor experimental data (open symbols)for CH3OH, CO, COs, H 2 mole fractions and temperature, all shown as functions of position. Curves represent computed results using the reaction mechanism from Ref. 43.

CH2OH+O 2 = CH20+HO 2

(150)

accounted for about 90% of the consumption of hydroxymethyl radicals in the lean case and its rate could be determined. In the rich case reaction (150) was far less important than reaction (149) CH/OH+M = CH/O+H+M. (149)

achieve spatial homogeneity, while it also operates normally at steady state. Reactants are injected at a uniform rate into a cavity through a number of inlets. Within the PSR, reactions take place, heat is released,

The computed simulations provide more detailed information than is available from the experimental data. For example, in Fig. 13 the computed concentration profiles of some of the intermediate species in the lean methanol oxidation case are shown. Of particular interest is the way the O and OH levels rise dramatically near the point where the methanol and intermediate hydrocarbon species concentrations effectively disappear, a phenomenon that has already been discussed in detail earlier. These flow reactor simulations were essential to the development of the methanol oxidation mechanism. The key feature which the flow reactor provides is its ability to spread out the fuel consumption region in space. The reaction zone is also stationary, so there is ample time for diagnostics to measure concentrations of stable intermediate and even some radical species. 5.1.4. Stirred reactor The shock tube is, to some degree, homogeneous in space but varying with time, while the plug flow reactor is steady in time but spatially varying. A third commonly used experimental system, the perfectly stirred reactor (PSR), combines the simpler features of these two. 389 This combustor utilizes rapid mixing to

I cH2o"

II

a [-

- A : ' x ',

II

\\

o
0 20 40
P o s i t i o n -- cm

60

80

FIG. 13. Calculated intermediate species concentrations in the lean methanol flow reactor model. Concentrations of species indicated by superscript "a" have been reduced by a factor of 10.

Modeling of hydrocarbon combustion and product gases are removed through exhaust openings. The characteristic time scale is the average residence time in the reactor. The overall pattern of fuel consumption and product formation in the PSR is simpler than in the shock tube or flow reactor. Because radical species are present in appreciable quantities, there is no fuel pyrolysis or initiation phase at all. To maintain stability and prevent blowout, most of the heat of reaction must be released within the reactor. Therefore, the fuel consumption and the CO and H 2 oxidation phases occur simultaneously. We discussed earlier that the most important oxidation reactions of CO and H 2 are with OH radicals, both of which are substantially slower than reactions between hydrocarbon species and OH (see Table 2). If radical species levels are high enough to oxidize CO and H 2 on a time scale which is comparable with the residence time, then hydrocarbon consumption must take place on a much shorter time scale. As a result, hydrocarbon species are normally detectable in the PSR only near blowout conditions. 39 Under fuel-rich conditions, the fuel does not react as rapidly or completely, and fuel consumption reactions play a larger role than for lean or stoichiometric mixtures. Therefore, rich mixtures will also reach blowout limits more easily (i.e. at longer residence times or at higher operating temperatures). Of course, real stirred reactors are not homogeneous and these inhomogeneities can complicate comparison between computed model results and experimental data. Since radical levels and heat release in the PSR are not sensitive to the details of the fuel consumption, and because the consumption of hydrocarbon species is so rapid, the PSR is an ideal combustion environment for the quasi-global kinetics model of Edelman and Fortune, 352 discussed earlier. In this type of mechanism, the fuel reacts directly, generally in a single global reaction step, to produce CO and H 2. The subsequent oxidation of CO and H 2 is then described by a detailed C O - H 2 - O 2 mechanism. As we have seen, this describes very well the general oxidation pattern in the PSR, with a very rapid global step. The quasi-global approach has been successfully applied to the problem of N O x production in the PSR. 355'391'392 The rate of the global step usually depends on fuel and oxidizer concentrations and on pressure and temperature. However, since there is no explicit dependence on radical species concentrations, this global step plays no role in limiting radical levels. Predicted H, O, and OH concentrations are those which are characteristic of C O - H 2 oxidation, resulting in good predictions of NOx production via the Zeldovich mechanism. F o r fuel-rich conditions, reactions of fuel and hydrocarbon intermediates play a much larger role in controlling the radical pool, and other submechanisms are responsible for NOx production in addition to the Zeldovich reactions. As a result, the assumptions of the quasi-global mechanism are not as well satisfied, and the approach is less successful for describing rich condition in the PSR. 5.2. Kinetic Models with Transport

37

When transport effects are included in a combustion model new difficulties arise. Mathematically, the conservation equations become partial differential equations involving derivatives with respect to both space and time. The software packages which are available to solve ODE's for stiff kinetic equations must be combined with numerical methods for dealing with the spatial transport. The computational costs of these combined models are significantly greater than for the purely kinetic problems already discussed for two reasons. First, a separate kinetic matrix is needed for every spatial point, and there can be dozens or hundreds of such points in a given problem. Second, the kinetic problems at each point are coupled together through the effects of spatial transport. On present computers such as a CRAY-1 or CDC 7600, a typical kinetics problem can require anywhere from a few seconds to a fraction of a minute of CPU time while a one-dimensional laminar flame model can require several minutes or even hours, depending on the problem parameters, the complexity of the reaction mechanism, and the details of the overall solution algorithm. In addition to an increased computational cost, when spatial variations are included additional physical data are required. For the simplest cases this includes thermal and species transport coefficients which depend on the dominant mode of transport in the problem. Often the transport coefficients are not well defined because the fluid mechanical treatment is incomplete. For example, in turbulent flames both heat and mass transfer are controlled by the turbulence, but current turbulence models cannot specify these transport properties exactly. In laminar regimes, the transport mechanisms are established in principle, but computations of these properties can be very complex and costly. As a result of heat and mass transport, basic kinetic phenomena can be affected. For example, in laminar hydrocarbon flames, H atom diffusion is very rapid due to its low atomic weight. Significant percentages of the H atoms produced in the high temperature part of the flame are not consumed there. Instead the H atoms diffuse into the colder unreacted region of the flame and react with fuel and Oz molecules. These reactions therefore are more important at lower temperatures in flame environments than in flow reactors or shock tubes. Other examples of fundamental kinetic changes when spatial transport is included in the combustion model will be discussed below. 5.2.1. Laminar flames The majority of combustion problems in which detailed chemical kinetics and spatial transport have been combined are models of laminar flames. In one sense the laminar flame presents fewer difficulties than most other spatially varying combustion problems because the relevant transport parameters are well defined. Heat transport takes place primarily by

38

C.K. WESTBROOK and F. L. DRYER Warnatz 44 showed how pressure and temperature variations, with their influence on the relative rates of reactions (1) and (9), affected laminar flame structure and burning velocity of H 2 O 2 - N z laminar flames. For premixed gases, a variety of solution techniques for the laminar flame equations have been developed, each with features to recommend it in one or more types of applications. Some deal with stationary premixed flames46'4 1,402 while others treat freely propagating premixed flames. 45'46 Most models are for adiabatic flames, but some models 136'41 simulate heat loss to a burner. Models for one dimensional opposed-jet diffusion flames ~39 and two dimensional laminar axisymmetric diffusion flames64 have also been developed. Most of these formulations assume that the pressure is uniform but some 45'46 include the effects of pressure gradients in the momentum conservation equation. In order to address transient phenomena such as flame acceleration or flame quenching, a model needs to be time dependent and should include these pressure gradients. Generally speaking, comparison between computed and experimental laminar burning velocity alone is not an adequate test of the validity or accuracy of a given reaction mechanism. It is often possible for an inadequate or incomplete mechanism to provide an accurate prediction of the laminar burning velocity for a given fuel-oxidizer mixture. Burning velocity predictions are a necessary but not sufficient test of a reaction mechanism. A much more complete and demanding test is provided when both major and radical species concentration profiles through a flame are available from experimental data. Because atmospheric and higher pressure laminar flames are very thin, spatial resolution difficulties in conventional diagnostic systems have limited most careful flame structure measurements to much broader flames at very low pressure. For example, the methane oxygen laminar flames which have most often been used for detailed model comparisons have included results of Fristrom 4v at 76torr, of Biordi eta/. 327 at 32tort, and of Peeters and Mahnen ~4 at 40torr. Recent developments in the use of laser diagnostics for flame structure measurements have relieved some of the resolution limitations. For example, Cattolica 48 measured OH profiles in laminar methane-air flames at atmospheric pressure, using laser absorption spectroscopy, and then compared the measurements with theoretical predictions from several numerical modeling studies. However, much more information is needed in flames at atmospheric and higher pressures. Quite recently, Hahn and Wendt 139.409 presented a numerical model for a laminar opposed jet diffusion flame. Although the convectional flame in this system is two-dimensional, the spatial variations of temperature and density were shown to be essentially onedimensional. This type of model represents a very valuable development because it enables one to examine an entirely new range of experimental regimes which are not available in premixed systems. In a diffusion flame the composition ranges from pure fuel

thermal conduction, although a small contribution is also provided by the Dufour effect where a heat flux is caused by concentration gradients. Transport of chemical species is dominated by molecular diffusion, with another small contribution from the Soret effect where species diffusion can result from thermal gradients. However, in practice the computation of the structure of a laminar flame can be very difficult in multicomponent mixtures where steep time-varying temperature, density, and compositional gradients occur, coupled with the effects of chemical kinetics. The equations for laminar flame propagation in one dimension have also been u s e d 69'393 t o examine some features of detonations in Hz-Oz mixtures. Several authors have dealt specifically with methods for computing these transport effects. 1~,46,85,90,91,136, 394 4.02 These treatments cover a wide range in complexity, and in most cases the most sophisticated and exact prescriptions are also the most difficult and time-consuming to implement. Some useful generalizations can be made which are independent of the exact transport model chosen. First, molecular species diffusion depends on the species molecular weight such that heavier species diffuse more slowly. This means that low mass species such as H, H2, O, and OH will diffuse most rapidly. Hydrogen atom diffusion is especially important. Combined with the dominant role of reaction (1) H+O 2 = O+OH (1)

in determining chain branching rates in much of hydrocarbon combustion, the high diffusion rate of H atoms is responsible for many important phenomena connected with laminar flames. In addition, transport coefficients generally increase with increasing temperature and decrease with increases in density. These overall trends have a considerable effect on the variation of laminar burning velocity with pressure and unburned gas temperature. 43 Dixon-Lewis61'62'91 and Warnatz 44 have devoted a great deal of analysis to H z 0 2 and HE-air laminar flames. This work emphasized the importance of transport coefficients in determining flame properties. As we have already observed, H atoms are the most active diffusive species, and their role in promoting chain branching is most prominent in H2-fueled flames where hydrocarbon species are not present to inhibit their reaction with 0 2 molecules. The work of Dixon-Lewis demonstrated the importance of including variations in third body efficiencies for termolecular radical recombination reactions. We noted earlier that reaction (9) H+O2+M = HO2+M (9)

has a strong influence on H 2 and hydrocarbon oxidation rates because it competes with the chain branching reaction (1). The rate of reaction (9), like most other recombination reactions, varies with different third bodies. For example, water molecules are very efficient third bodies for recombination while monatornic gases such as argon are very inefficient.

Modeling of hydrocarbon combustion


I ' I ' H20 0.2 2000 I ' I

39

02
g
CH3OH v o, 0.01

0.1

1000~

0.005

I J

I I -2

I ~

I I -1

I / , / I

,I ' .~~ - I ~ I ~ 1

I ,

Relative flame position (mm)

Fie. 14. Temperature and major species concentration profiles in an atmospheric pressure, stoichiometric CHaOH-air laminar flame, taken from Ref. 47.

-2

-1

Relative flame position (mm)

to pure oxidizer in a single problem. Oxidation rates, radical levels, pollutant formation rates, and other kinetic details depend strongly on the local fueloxidizer ratio and are significantly different in diffusion flames from the trends observed in premixed flames. Since diffusion flames are an important and often dominant feature of many practical combustion systems ranging from diesel engines and gas turbines to fumaces, detailed kinetic models of diffusion flames promise to add a new dimension to present capabilities for analysis of real combustors. To provide a framework for subsequent discussion, we will first describe in some detail computed results for a typical laminar flame. For this reference model, a laminar flame in a premixed stoichiometric methanol-air mixture at atmospheric pressure and an unburned gas temperature of 300 K has been taken from our earlier work. 47 The specific numerical treatment and the detailed reaction mechanism used are described in the reference cited, but the points which will be emphasized do not depend on the details of the model. In Figs. 14 and 15 the spatial variations in the temperature, major species, and selected other species concentrations are plotted as functions of distance from an arbitrary flame location. In these computations the flame position is defined as the point at which the temperature is equal to 1500 K. Other convenient definitions, such as the point in the flame at which the local rate of heat release due to chemical reactions is a maximum, or the point at which the spatial temperature gradient is a maximum, give flame positions which are very nearly the same as that used for the present example. There are several major regions into which the flame can be subdivided, and in each region the principal kinetic behavior is different. Well ahead of the flame (Region I, not shown in Fig. 14), the tem-

FIG. 15. Major intermediate and radical concentration profiles for the CH3OH-air laminar flame of Fig. 14.

perature is low and the unreacted fuel and air are relatively inert. This is followed by an induction region (Region II) which extends from 1.0mm to about 3.0 mm ahead of the flame and is characterized by a gradual growth in the radical pool. Heat conduction from the hot flame zone provides a slow increase in the gas temperature, but very little heating from chemical reactions occurs. Because the temperature is still very low, chain branching does not occur (recall E 1 = 16.8 kcal/mol), and the principal source of radical species is by diffusion from the much hotter flame zone where the radical species levels are very high. Unlike the induction regime in shock tube models, fuel decomposition reactions do not occur due to the low temperature. However, reactions between fuel molecules and radical species, particularly H atoms, including CH3OH + H = CH 3 + H20 CH3OH+H = CH2OH+H 2 CH3OH+OH = CH2OH+H20 C H 3 O H + O = CH2OH + O H (143) (144) (145) (146)

begin the process of fuel consumption. Hydrogen atoms also recombine with 02 molecules to form HO2, leading to fairly large amounts of H 2 0 2 H+O2+M = HO2+M
HO 2 + HO 2 = H202 + 02

(9) (15) (148)

CH3OH+HO 2 = CH2OH+H202.

Recombination is favored over branching for H + 02 in this part of the flame because E 9 = - 1.0 kcal/mol, so the low temperature does not affect reaction (9).

40

C.K. WESTBROOKand F. L. DRYER the above example. Quantitative differences will be observed between various fuels. With acetylene, for example, some CO and H z oxidation will begin prior to the disappearance of the fuel because the reactions of C2H2 with H atoms and OH radicals do not compete as effectively with reactions (3) and (28) as the analogous reactions for other hydrocarbon fuels (see Table 1). Therefore Regions III and IV will overlap to some extent for C2H2-air flames, but the overall structure of the laminar flame will still be very similar to that above for CH3OH-air. Detailed kinetic flame models have been instrumental in explaining, the behavior of many combustion problems which had been observed experimentally. For example, it has long been known 41'411 that the laminar burning velocity in hydrocarbon-air mixtures decreased with increasing pressure. However, as shown earlier for several fuelair mixtures in Fig. 4, there appear to be two distinct regimes, divided at about 2atm. pressure. Below this range, variations in pressure have a relatively small effect on Su (SuctP - ' l z for methane-air) but a more pronounced effect at elevated pressure (SuctP-5). 85'157 The same type of two-regime behavior is observed for other hydrocarbon-air mixtures as well,86A96 although the respective pressure exponents are not the same as those for CH4-air. The explanation for this phenomenon is provided by laminar flame modeling analysis and detailed kinetic mechanisms provided an essential part of the answer. The ratio of the rates of reactions (1) and (9) is pressure dependent, so the relative importance of reaction (9) grows quite rapidly with increasing pressure. We have already seen how diffusion of H atoms ahead of the flame region is important, providing radical species from reaction (1) and, along with the O and OH radicals generated, beginning the process of fuel molecule consumption. Because the activation energy of reaction (1) is much greater than that of reaction (9), there is always a point ahead of the flame beyond which the temperature is so low that k 9 exceeds k~. As the pressure of the flame system is increased, this transition point moves closer to the principal flame zone and interferes with the process of fuel consumption. The competition between reactions (1) and (9) becomes important in the main fuel consumption zone ( ~ 700-1100 K) for pressures above one atmosphere, the same pressure regime in which the variation of Su with pressure begins to change character. Finally, when k 9 is artificially set equal to zero, there is no curved region for any of the fuel-air mixtures in Fig. 4. Clearly the high pressure variation of Su is caused almost entirely by reaction (9). This illustrates a useful feature of a computational tool, the ability to simulate problems with physically unrealizable parameters to isolate particular effects. Another combustion phenomenon which has been observed experimentally for many years is that of flame inhibition. We discussed earlier how kinetic modeling by many authors has shown that the dominant feature of inhibition by halogenated compounds

The hydroxymethyl radicals formed in Region II react primarily with 02 CH2OH+O / = CHzO+HO 2 (150)

rather than decomposing because E150 is much less than E149 for the decomposition (6.0 vs 29.0 kcal/mol). The concentration of the stable intermediate C H 2 0 begins to increase rather rapidly. In the next part of the flame (Region III), the majority of the fuel consumption occurs. The principal reactions are still reactions (143-146), but the higher temperature and radical species concentrations in Region III make the fuel consumption rate much faster than in Region II. The temperature is growing due primarily to thermal diffusion from the hot flame zone, which is also providing a steady source of radical species. Because the temperature is higher (400-1200K), the hydroxymethyl radicals can now decompose CH2OH + M = C H 2 0 + H + M (149)

providing H atoms which are much more reactive than the HO2 radicals from reaction (150). The fuel consumption is accompanied by a rapid growth in the levels of major intermediate species, including C H 2 0 , H2, and CO. These intermediates are not immediately consumed because the radical species concentrations (particularly OH) remain fairly small until the fuel has all disappeared. The inhibition of CO and H 2 oxidation by the hydrocarbon molecule has been emphasized in earlier discussions and is illustrated by the computed radical species mole fraction profiles shown in Fig. 15. Approximately half of the total temperature increase occurs in Region III, resulting in large part from the production of carbon monoxide. When the CHaOH fuel is effectively gone (at about - 0 . 2 m m in Figs. 14 and 15), the chain branching inhibition ends and the radical population grows very rapidly in Region IV, extending from about - 0 . 2 - - 1 . 0 mm. This region is characterized by the elementary reactions of the H 2 - C O - O 2 system and most of the remaining temperature increase occurs here. Finally, in the burned gas product Region V, a slow approach to final chemical equilibrium takes place, involving the termolecular recombination reactions. Although this example considers the fuel to be methanol, the same general regions occur for other hydrocarbon fuels as well. Much of the flame structure is a result of the diffusion of heat from the flame zone into the unburned reactant gases. Therefore the spatial extent of Regions I - I I I will depend on the characteristic length scale for thermal diffusion of energy. There is always an induction region which is about one thermal diffusion length ahead of the flame itself, where only reactions with low activation energies between fuel molecules and radicals will occur, together with reactions (9) and (15) as shown above. The other general regions will also be present, regardless of the specific hydrocarbon fuel, because the fundamental processes taking place are the same as in

Modeling of hydrocarbon combustion


(a) 1"0 Z - - ~ - - - ~ I F1/3 I
lOO

41

\
\k

[] o.1 atm 0 1 atm

--

~e a0
'
o

f /~\
/ /

zX10at.,
V 1%CH3Br

-~ 0.5 ~E

CF20 2~

.~

4 0

--

2o

r~b)
B~_r

2 Equivalence ratio

.0--

--

0.5 L~'--~
HBr / [Br2 10

FIG. 17. Computed variations of laminar burning velocity with equivalence ratio. Solid curves show results for ethylene air at indicated pressures, dashed curves are for atmospheric pressure ethylene-air with 1 ~ inhibitor added.

H I
"",

~ I
-

0
10

2111111
"'--"H C 4

\\~

/,"

jJ

.
1500
_,ooo =

CO

.//
0 1 Relative flame position (rnm)

500

-1

FIG. 16. Selected speciesand temperature profiles in stoichiometric, atmospheric pressure CH4-air flames. Solid curves represent results for mixtures with 1 ~ CF3Br added, dashed curves represent results without inhibitor. involves the catalyzed recombination of H atoms, lowering the chain branching rate and reducing the size of the radical pool. This process occurs over extended ranges of pressure and temperature, resulting in inhibition of flame propagation, ignition, and detonation. The interactions between kinetics and transport are particularly well defined for the case of a methane-air flame to which 1 ~ CFaBr has been added. 32'a31 In Fig. 16 computed species profiles through the flame are shown for the primary species, together with their respective values in the same flame without CF3Br. The additive produces a slight broadening of the flame but has a negligible effect on the final equilibrium temperature and composition. The inhibitor has a small effect on the main radical species levels in the principal flame region, but a much more pronounced effect in the pre-flame region ahead of the flame, where the inhibitor causes a large reduction in the radical pool. In Fig. 16 the principal Br-containing and F-containing species profiles are plotted. Comparison of these concentrations with the temperature curve indicates that the inhibitor con-

sumption takes place in the relatively cool pre-flame region (T ~- 1000 K). Levels of some of the important halogenated intermediate species Br2, CHaBr and others peak in the same region. Hydrogen atoms are produced in great quantities in the high temperature (T = 1400-1800 K) flame zone and diffuse rapidly upstream into the fuel induction region. At these cooler temperatures the reactions with the halogenated species proceed very rapidly, having lower activation energies than the reactions of H atoms with 02 or with fuel species. This means that the kinetic features of flame inhibition are most important at temperatures around 1000 K, rather than at the higher temperatures typical of the main flame zone or the burned products. Laminar flame modeling can provide theoretical estimates of flammability limits. For example, Fig. 17 shows the computed laminar burning velocity for a variety of ethylene-air mixtures, s6 The solid curves represent computed results at 0.1, 1.0 and 10.0atm. pressure, while the dashed curves show results for atmospheric pressure flames to which 1 ~ CH3Br and CFaBr have been added. The lean flammability limit ~bL for a given mixture and pressure is easily estimated by extrapolating the curve to zero burning velocity, and the lean limit is the intersection with the horizontal axis. Values of ~bL computed in this manner agree well with experimental d a t a . 46"47"86'136'196 Extension of this approach to computation of the rich limit JR is made difficult by two factors. First, as shown in Fig. 17, the computed curves fall much more slowly on the rich side of stoichiometric than on the lean side. Although it is not shown in the figure, the model computes a value of Su = 5 cm/sec at ~b = 4 for the uninhibited case at atmospheric pressure. Similar slow variations in burning velocity are calculated for other very rich fuel-air mixtures. 46n36,196 In fact, an adiabatic model will continue to predict a positive burning velocity for essentially any equivalence ratio. Extensive experimental and theoretical work by Hertzberg 4~2 has shown that, due to a variety of extinction mechanisms, actual flames will become unstable and cease propagation when the burning

42

C.K. WESTBROOKand F. L. DRYER

1.8 [

CH4-air
-e-o1 " 4 30 c ~ ~ ~ " ' ~ ...--

~ 8 12~--~
1,0 0.8 0,6 0.4

m/s~ 20cm/s

~%~0em/s

_
~ ~ -

_ ~

ammability limit

2 3 Molepercentinhibitor-- CH3Br

FIG. 18. Computed contours of constant laminar burning velocity in atmospheric pressure methane-air CH3Br mixtures. Dashed contour represents computed extinction limit.

velocity falls below approximately 5 8 cm/sec. When this range is used as a guideline, values of q~Rcomputed by current models agree fairly well with experimental data. The second limitation in the ability of numerical flame models to compute rich flammability limits lies in the fact that recombination and addition paths leading to species considerably larger than the initial fuel itself become quite important in such environments. Uncertainties in the rates and product distributions for some of these paths make accurate determinations of the rich limit difficult. Computations of the variation in laminar burning velocity with inhibitor concentration can provide a flammability limit at a given equivalence ratio. Under these conditions this particular flammability limit is more properly termed an extinction limit, an inhibitor concentration beyond which flame propagation will not occur. In Fig. 18 contours of constant laminar burning velocity are shown for methane-air flames with CH3Br as an inhibitor. 86 The amount of inhibitor which will extinguish the flame varies with equivalence ratio. The dashed line in Fig. 18 is a numerical estimate of the extinction-flammability limit, again defined as the mixture corresponding to a burning velocity of 5-Scm/sec, and the computed limits agree reasonably well with experimental data. 77 The entire subject of flammability limits, including the problems of defining a limit, understanding what physical and chemical factors influence these limits, and computing them from fundamental principles, is a large subject area in itself and has been reviewed a number of times recently.413,414 The previous examples show how detailed kinetic models, coupled to a description of the transport occurring in a laminar flame, can explain on a quite general basis processes which are important in many combustion environments. Although they represent greatly idealized theoretical models, each has substantial implications for very practical problems, since the same kinetic processes occur in practical combustors as in laminar flame environments. The same type of

modeling approach has begun to be applied directly to a variety of applied combustion problems. A coupled experimental and modeling study by Smith e t al. 415 examined flame propagation in lean premixed fuel-air mixtures which are subjected to a rapid volume expansion. These conditions were chosen to simulate the effects of piston motion in internal combustion engines in which extremely late ignition timings are used. With retarded ignition, combustion can continue well into the expansion stroke of the engine cycle. For fuel-lean operation, occasional flame propagation failures or "bulk quenchings" are observed 416 even when successful ignition has occurred. The model analysis indicated that the rapid expansion can lower radical concentrations to a point where chain propagation and branching rates are too low to sustain flame propagation. Perhaps the best example of this emerging class of combustion models concerns the analysis of flame quenching in cold boundary layers such as those which exist on the walls of internal combustion engines. For many years it was believed that this process was

T = Tadiabati c

Wall

~ t u r e T = 1500K~ - ~ I Burned / I .'T\ I reactionproducts Unburned . / J/l "'~. Chemicalheat ~ i ' ~ "'. relea~rate

T = Twall qtqr Distance


FIG. 19. Schematicviewofhead-onflamequenchingat acold wall, showing the layer of unburned fuel-air mixture between the stationary flame and the wall. The graphical definitions of the thermal flame thickness L~and flame positions qt and q, are also shown.

Modeling of hydrocarbon combustion responsible for the majority of the unburned hydrocarbon emissions from automobile engines. 417-419 However, the results of several modeling studies 214' 420-423 have indicated that this picture may not be correct. The configuration used 42 is shown in Fig. 19 and provides a good example of the way in which an actually three-dimensional problem can be addressed in an idealized one-dimensional form. In this schematic diagram, the flame approaches the wall at the left with the wall temperature held fixed at a prescribed value Tw, resulting in a narrow cold gas layer next to the wall. The progress of the flame towards the wall is eventually halted when radical recombination in the cold layer interferes with chain propagation and branching rates ahead of the flame. The flame stagnates at a distance from the wall which is in close agreement with experimental measurements, 417 leaving a quench layer of unburned hydrocarbon fuel between the wall and the flame. However, the model predicts that subsequent to flame stagnation, the fuel in the quench layer diffuses rapidly away from the wall and is oxidized in the nearly stationary flame zone. This consumption of fuel is illustrated in Fig. 20 for a stoichiometric methanol-air mixture quenching at 10 atm. pressure. The time scale is relative to tq, the time at which the progress of the flame towards the wall is arrested, and the remaining fuel and other intermediate hydrocarbon species are summed and expressed in parts per million unburned "C". The total unburned carbon falls to very low levels (less than 100ppm) on a time scale which is short compared with typical engine residence times. Similar numerical results have been obtained by several other studies and are consistent with recent experiments. 424-43 Thus, flame quenching in thermal boundary layers is probably not responsible for most of the unburned hydrocarbon emissions from internal combustion engines and other potential sources for these emissions must be examined. 43~ Similar analysis has been applied to the problem of the formation and evolution of a reactive boundary layer near a cold end wall of a shock tube. 432 A third distinct type of flame quenching occurs during stratified charge combustion and can be responsible for the emission of large quantities of unburned hydrocarbons in some cases. 433 ModelI I I I

43

ing analyses with detailed kinetics in one spatial dimension 434'435 and with global kinetics in two dimensions 436 were essential parts in the identification of the physical mechanism for this "volume quenching". The time-dependent simulations of all three flame quenching phenomena require a more complex and general treatment than is adequate for the previously described models of steady-state flame structure and propagation rates. When heat transfer to the wall is included, somewhat more complex spatial boundary conditions are also needed. 5.2.2. Sensitivity analysis As reaction mechanisms grow larger and more complex, the subject of sensitivity analysis becomes increasingly important. Uncertainties in reaction rate parameters, transport coefficients, thermochemical properties, initial and boundary conditions, and other kinetic model quantities may result in corresponding uncertainties in computed simulations of combustion problems. Sensitivity analysis provides a means of estimating these computational "error bars". In its most primitive form, "brute force" sensitivity analysis involves repeating a given calculation many times, varying one parameter at a time and then comparing the results with a reference or baseline case. If the solution changes a great deal as the parameter is varied, the sensitivity to that parameter is large. If transport effects can be neglected, then the observable parameters consist of the ( N - 1 ) species concentrations plus the temperature, while for spatially varying problems it would include (N + 3)L equations for concentrations, temperature, and components of momentum in each of L spatial zones. The variable parameters are the rate parameters (in principle the activation energy and pre-exponential of both forward and reverse rates could be considered independent parameters), thermochemical quantities, initial and boundary conditions, the transport coefficients in spatially varying problems, and any other system parameters in which uncertainties might influence the numerical solution. Most commonly the number of parameters M is considerably larger than the number of ODE's N. Most sensitivity analysis has emphasized zero-dimensional problems, but there is a considerable amount of current research into extending the approach to treat spatially-varying problems. The "brute force" method of sensitivity analysis becomes prohibitively expensive as the number of species, zones, and parameters grows. Alternative, less costly and more general techniques for evaluating the sensitivity coefficients have been developed in recent years. The Direct Method (DM) 437,438 considers the equation for the sensitivity coefficients together with the original ODE's, solving simultaneously for the time-dependent concentrations, temperature and sensitivity coefficients. The calculations are somewhat simplified because the Jacobians for the sensitivity equations are already available from the original kinetics problem. Since the computation of the Jacobian is a rather time-consuming task, this saving

0.14 0.12 E E
O.lO i

5000
I "~ E

4000 3000
2000 1000
Methanol-Air

o.o8 8
-- 0.06

P= 10.0atm

_ 0.04

@= 1.0 Tw = 3 0 0 K

E -- 0.02 ~_

I
-0.2 -0.1

I-""-~ 0 0'.1 t - tq (ms)

1 0.20

0.0

FIG. 20. Average fuel concentration and flame position, both as functions of time relative to quench time tq for the C H 3 OH-air flame indicated.

44

C.K. WESTBROOKand F. L. DRYER


REFERENCES

can be significant. However, this method still requires the solution of N ( M + 1) stiff equations, N in the original system and N M more for the sensitivity coefficients. In contrast, the Fourier amplitude sensitivity test (FAST) method 439-442 considers the system parameters to be periodic functions of a search variable s. Each parameter has by assumption a unique frequency response to variations in s. As s is varied and the O D E ' s are solved with modified values, of the input parameters, Fourier analysis of the computed results can identify those parameters with the greatest sensitivity by their frequency signatures. These sensitivity coefficients are actually averages over the variations in all of the other parameters being considered and therefore include some higher order sensitivity information. Both the D M and F A S T approaches require the solution of large numbers of equations. A third technique, the Green's Function Method ( G F M ) has been developed 66'443-445 to reduce the number of stiff differential equations which must be solved. Instead, the G F M requires the evaluation of integrals which are much less expensive to compute than the corresponding differential equations. In addition, higher order sensitivity data can be obtained inexpensively, and the G F M becomes increasingly advantageous, relative to the D M and F A S T approaches, as the size of the kinetic mechanism becomes larger. General, user-oriented computer software implementing the G F M is being developed 446 which will greatly improve the accessibility of sensitivity analysis techniques to non-specialists. Comparisons between sensitivity coefficients computed by these different techniques have shown generally consistent results. 67'444"447'448 For example, reactions (1) and (9) have the greatest sensitivities in H2 0 2 kinetics, 66'448 consistent with our earlier discussions. The large sensitivity of reaction (46b) in m e t h a n e - a i r mechanisms 445,447 was an important part of the re-evaluation of those mechanisms as noted earlier. The computational costs can vary widely depending on the method used and the details of the specific problem being analysed. Further extensions to spatially varying environments and to ever-larger reaction mechanisms are motivating continuing efforts at making sensitivity analysis programs faster and easier to implement.
Acknowledgements--This work has benefited from discussions and other assistance from colleagues too numerous to mention here. However, we wish to express our particular thanks to the reviewers from this journal and, additionally, to Professor I. Glassman. All were very conscientious and helpful to us in suggesting revisions of the original manuscript. This work was performed under the auspices of the U.S. Department of Energy by the Lawrence Livermore National Laboratory under contract number W-7405-ENG-48, and was supported at Princeton University by the U.S. Department of Energy, under contract No. DE-AC02-83ER13045 through the Division of Chemical Sciences, Office of Basic Energy Sciences, and by the Mobil Research and Development Corporation.

1. BOWMAN,C. T., Pro#. Energy Combust. Sci. 1, 33 (1975). 2. HEvwooo, J.B.,Prog. EnergyCombust.Sci. l,135(1976). 3. SgRovlM, A. F. and FLAGAN, R. C., Prog. Energy Combust. ScL 2, 1 (1976). 4. BENSON,S. W., Pro#. Energy Combust. Sci. 7, 125 (1981). 5. DIXON-LEWIS,G. and WILLIAMS,D. J., Comprehensive Chemical Kinetics, C. H. BAMEORDand C. F. H. TIPPER, eds., vol. 17, p. 1, Elsevier (1977). 6. MCKAY,G., Prog. Energy Combust. Sci. 3, 105 (1977). 7. BALDWIN,R. R., BENNETT,J. P. and WALKER,R. W., Sixteenth Symposium (International) on Combustion, p. 1041, The Combustion Institute, Pittsburgh (1977). 8. WILLIAMS,F, A., Combustion Theory, Addison-Wesley, Reading, Mass. (1965). 9. BRACCO,F. V., Society of Automotive Engineers paper SAE 741174 (1974). 10. MCDONALD,H., Prog. Energy Combust. Sci. 5, 97 (1979). 11. ORAN,E. S. and BORIS,J. P., Prog. Energy Combust. Sei. 7, 1 (1981). 12. GEAR, C. W., Numerical Initial Value Problems in Ordinary Differential Equations, Prentice-Hall, New York (1971). 13. CURTIS,C. F. and HIRSCHEELDER, J. O., Proc. Natl. Acad. Sci. 38, 235 (1952). 14. Symposium on Reaction Mechanisms, Models, and Computers, New Orleans ACS Meeting, 1977. J. phys. Chem. 81, 2309 (1977). 15. HINDMARSH, A. C. and BERNE, G. D., EPISODE, University of California Lawrence Livermore Laboratory report UCID-30112, Rev. 1 (1977). 16. KEE, R. J., MILLER, J. A. and JEFFERSON, T. H., CHEMKIN, Sandia National Laboratories report SAND80-8003 (1980). 17. DRYER,F. L., NAEGEL1,D. and GLASSMAN,[., Combust. Flame 17, 270 (1971). ZELLNER,R., J. phys. Chem. 83, 18 (1979). 18. BENSON,S. W., Thermochemical Kinetics, Wiley, New York (1976). 19. GOLDEN,D. M., J. phys. Chem. 83, 108 (1979). 20. Symposium on Current Status of Kinetics of Elementary Gas Reactions: Predictive Power of Theory and Accuracy of Measurement, J. phys. Chem. 83 (1979). 21. COHEN,N., Int. J. chem. Kinet. 14, 1339 (1982). 22. COHEN, N., Nineteenth Symposium (International) on Combustion, p. 31, The Combustion Institute, Pittsburgh (1983). 23. BAULCH,D. L., DRYSDALE,O. D., HORNE, D. G. and LLOYD, A. C., Evaluated Kinetic Data for High Temperature Reactions, vols. l and 2, Butterworths, London (1973). 24. BAULCH,D. L., DRYSDALE,D. D., DUXBURY,J. and GRANT, S. J., Evaluated Kinetic Data for High Temperature Reactions, vol. 3, Butterworths, London (1976). 25. HAMPSON,R. F., JR. and GARVIN,D., National Bureau of Standards Special Publication 513 (1978). GARVIN, D., BROWN,R. L., HAMPSON,R. F., KURYLO,M. J. and TSANG, W., National Bureau of Standards Special Publication 531 (1978). 26. HOYERMANN, K. and WAGNER, H. GG,, Oxidation Commun. 2, 259 (1982). 27. HERRON,J. T. and HUIE,R. E., J. phys. Chem. Ref Data 2, 467 (1973). 28. WESTLEY,F., National Bureau of Standards publication NBSIR 81-2254 (1981). 29. JENSON,D. E. and JONES,G. A., Combust. Flame 32, 1 (1978). 30. KONDRATIEV,V. N., National Bureau of Standards publication COM-72-10014 (1972). 31. TROTMAN-DICKENSON, A. F. and MILNE,G. S., National Bureau of Standards publication NSRDS-NBS 9 (1967). 32. KERR,J. A. and Moss, S. J., Handbook of Bimolecular and Termolecular Gas Reactions, vols. 1 and 2, CRC Press, Boca Raton, Florida (1981).

Modeling of hydrocarbon combustion 33. GARDINER, W. C., JR. and OLSON, D. B., A. Rev. phys. Chem. 31,377 (1980). 34. WARNATZ,J., Survey of Rate Coefficients in the C/H/O System, Sandia National Laboratories report SAND838606 (1983). 35. LLOYD,A. C., Int. J. chem. Kinet. 6, 169 (1974). 36. HAMPSON,R. F., JR. and GARVIN, D., J. phys. Chem. 81, 2317 (1977). 37. DRYER, F. L. and GLASSMAN,I., Fourteenth Symposium (International) on Combustion,. p. 987, The Combustion Institute, Pittsburgh (1973). 38. WESTaROOK,C. K., CRE1GHTON,J., LUNO, C. and DRYER, F. L., J. phys. Chem. 81, 2542 (1977). 39. TROE, J., Fifteenth Symposium (International) on Combustion, p. 667, The Combustion Institute, Pittsburgh (1975). 40. TROE,J., J. chem. Phys. 66, 4745 and 4758 (1977). 41. LUTHER, K. and TROE, J., Seventeenth Symposium (International) on Combustion, p. 535, The Combustion Institute, Pittsburgh (1979). 42. CROCE, A. and TROE, J., First Specialists Meeting (International) of The Combustion Institute, p. 321 (1981). 43. WESTBROOK, C. K. and DRYER, F. L., Combust. Sci. Technol. 20, 125 (1979). 44. WESTBROOK, C. K. and DRYER, F. L., Eighteenth Symposium (International) on Combustion, p. 749, The Combustion Institute, Pittsburgh (1981). 45. WESTBROOK,C. K., DRYER, F. L. and SCHUG, K. P., Nineteenth Symposium (International) on Combustion, p. 153, The Combustion Institute, Pittsburgh (1983). 46. MILLER, J. A., MITCHELL, R. E., SMOOKE, M. D. and KEE, R. J., Nineteenth Symposium (International) on Combustion, p. 18l, The Combustion Institute, Pittsburgh (1983). 47. WESTBROOK,C. K. and DRYER, F. L., Combust. Flame 37, 171 (1980). 48. WmTBROOK,C. K., Combust. Flame 46, 191 (1982). 49. WESTBROOK, C. K. and URTIEW, P. A., Nineteenth Symposium (International) on Combustion, p. 615, The Combustion Institute, Pittsburgh (1983). 50. WESTaROOK, C. K., Proceedings of the International Conference on Fuel-Air Explosions, J. H. LEE and C. M. GUIRAO, eds., p. 189, University of Waterloo Press (1982). 51. STULL,D. R. and PROPHET,H., JANAF Thermochemical Tables, NSRDS-NBS 37, National Bureau of Standards and supplements, Washington (1971). 52. BAHN,G. S., NASA report CR-2178 (1973). 53. VOEVODSKY,V. V. and SOLOUKrI1N, R. I., Tenth Symposium (International) on Combustion, p. 279, The Combustion Institute, Pittsburgh (1965). 54. BROKAW,R. S., Tenth Symposium (International) on Combustion, p. 269, The Combustion Institute, Pittsburgh (1965). 55. BELLES, F. E. and LAUVER, M. R., Tenth Symposium (International) on Combustion, p. 285, The Combustion Institute, Pittsburgh (1965). 56. ASABA,T., GARDINER, W. C., JR. and STUBBEMAN,R. F., Tenth Symposium (International) on Combustion, p. 295, The Combustion Institute, Pittsburgh (1965). 57. BELLES,F. E., Seventh Symposium (International) on Combustion, p. 745, Butterworths, London (1959). 58. JENKINS, D. R., YUMLU, V. S. and SPALDING, D. B., Eleventh Symposium (International) on Combustion, p. 779, The Combustion Institute, Pittsburgh (1967). 59. HAMILTON,C. W. and SCHOTT,G. L., Eleventh Symposium (International) on Combustion, p. 635, The Combustion Institute, Pittsburgh (1967). 60. GARDINER,W. C., JR., MCFARLAND, M., MORINAGA,K., TAKEYAMA, T. and WALKER, B. F., J. phys. Chem. 75, 1504 (1971). 61. DIXON-LEwis, G., Proc. R. Soc. London A298, 495 (1967).

45

62. DIXON-LEWIS, G., Proc. R. Soc. London A307, 111 (1968). 63. STEPHENSON,P. L. and TAYLOR, R. G., Combust. Flame 20, 231 (1973). 64. MILLER, J. A. and KEE, R. J., J. phys. Chem. 81, 2534 (1977). 65. SCHEFER,R. W., Combust. Flame 45, 171 (1982). 66. DOUGHERTY, E. P. and RAaITz, H., J. chem. Phys. 72, 6 5 71 (1980). 67. SLACK,M. W., Combust. Flame 28, 241 (1977). 68. ATKINSON, R., BULL, D. C. and SHUEE, P. J., Combust. Flame 39, 287 (1980). 69. ORAN, E., YOUNG, T. and BORIS, J., Seventeenth Symposium (International) on Combustion, p. 43, The Combustion Institute, Pittsburgh (1979). 70. DOVE,J. E. and TRIBBECK,T. D., Astronautiea Acta 15, 387 (1970). 71. TSUGE, S., FURUKAWA, H., MATSUKAWA, M. and NAKAGAWA,T., Astronautica Acta 15, 377 (1970). 72. STREHLOW,R. A. and RUBINS, P. M., AIAA J. 7, 1335 (1969). 73. STREHLOW,R. A., MAURER, R. E. and RAJAN, S., AIAA J. 7, 323 (1969). 74. KAILASANATH,K., ORAN, E. S. and BORIS, J., Combust. Flame 47, 173 (1982). 75. ORAN, E. S., YOUNG, T. R., BORIS, J. P. and COHEN, A., Combust. Flame 48, 135 (1982). 76. ORAN, E. S. and BORIS, J. P., Combust. Flame 48, 149 (1982). 77. LEWIS, B. and YON ELSE, G., Combustion, Flames and Explosions of Gases, Academic Press, New York (1961). 78. KONDRATIEV, V. S. and AZATYAN, V. V., Fourteenth Symposium (International) on Combustion, p. 37, The Combustion Institute, Pittsburgh (1973). 79. BALDWIN,R. R. and WALKER, R. W., Seventeenth Symposium (International) on Combustion, p. 525, The Combustion Institute, Pittsburgh (1979). 80. HACK, W., WAGNER,H. GG. and HOYERMANN,K., Ber. Bunsenges. phys. Chem. 82, 713 (1978). HACK, W., PREUSS, A. W. and WAGNER, H. GG., BeE. Bunsenges. phys. Chem. 82, 1167 (1978). HACK, W., PREUSS, A. W., WAGNER, H. GG. and HOYERMANN,K., BeE. Bunsenges. phys. Chem. 83, 212 (1979). HACK, W., PREUSS, A. W., TEMPS, F. and WAGNER, H. GG., BeE. Bunsenges. phys. Chem. 83, 1230 (1979). 81. SRIDHARAN, U. C., Qiu, L. X. and KAUEMAN, F., J. phys. Chem. 86, 4569 (1982). 82. KURYLO, M. J., KLAIS, O. and LAUFER, A. H., J. phys. Chem. 85, 3674 (1981). 83. KEYSER, L. F., J. phys. Chem. 86, 3439 (1982). SIMONAtTIS, R. and HEICKLEN, J., J. phys. Chem. 86, 3416 (1982). 84. HOWARD,C. J., J. chem. Phys. 71, 2352 (1979). 85. TSATSARONIS,G., Combust. Flame 33, 217 (1978). 86. WESTBROOK, C. K., DRYER, F. L, and SCHUG, K. P., Combust. Flame, in press (1983). 87. SNYDER, A. D. and SKINNER, G. B., Combust. Flame 7, 390 (1963). 88. SNYDER,A. O. and SKINNER, G. B., Combust. Flame 8, 164 (1964). 89. WESTBROOK,C. K., Combust. Sci. Technol. 29, 67 0982). 90. DIXON-LEWIS,G., GREENBERG,J. B. and GOLDSWORTHY, F. A., Fifteenth Symposium (International) on Combustion, p. 717, The Combustion Institute, Pittsburgh (1975). 91. DIXON-LEwlS, G., Proc. R. Soc. London A317, 235 (1970). 92. STRENG, A. G. and GROSSE, A. V., Sixth Symposium (International) on Combustion, p. 264, Reinhold, New York (1957). 93. CRAMAROSSA, F. and DIXON-LEWIS, G., Combust. Flame 18, 43 (1972). 94. BLEDJIAN,L., Combust. Flame 20, 5 (1973). 95. MARGOLIS,S. B., J. computational Phys. 27, 410 0978).

46

C.K. WESTBROOKand F. L. DRYER

96. WARNATZ, J., Ber. Bunsenges. phys. Chem. 82, 193 (1978). 97. HEIMERL,J. M. and COFFEE, T. P., Combust. Flame 39, 301 (1980). 98. WILDE,K. A., Combust. Flame 18, 43 (1972). 99. CAMPBELL,E. S., Chem. Engng. Sci. 20, 311 (1965). 100. HIRSCHEELDER, J. O., CURTISS, C. F. and CAMPBELL, D. E., J. phys. Chem. 57, 403 (1953). 101. REITZ, R. D., Eighteenth Symposium (International) on Combustion, p. 433, The Combustion Institute, Pittsburgh (1981). 102. GALANT, S., Eighteenth Symposium (International) on Combustion, p. 1451, The Combustion Institute, Pittsburgh (1981). 103. BROKAW, R. S., Eleventh Symposium (International) on Combustion, p. 1063, The Combustion Institute, Pittsburgh (1967). 104. ATRI, G. M., BALDWIN, R. R., JACKSON, D. and WALKER,R. W., Combust. Flame 30, 1 (1977). 105. HOWARD,C. J., J. chem. Phys. 71, 2352 (1979). 106. YETTER, R. A., DRYER, F. L. and RABITZ, H., Seventh International Symposium on Gas Kinetics, Gottingen (1982). 107. GETZINGER, R. W. and SCHOTT, G. L., J. chem. Phys. 43, 3237 (1965). 108. SMITH,I. W. M. and ZELLNER, R., Proc. Eur. Combust. Symp., 24 (1973). 109. BIERMANN, H. W., ZETZSCH, C. and STUHL, F., Ber. Bunsenges. phys. Chem. 82, 633 (1978). 110. DEAN, A. M., CRAIG, B. L., JOHNSON,R. L., SCHULTZ, M. C. and WANG, E. E., Seventeenth Symposium (International) on Combustion, p. 577, The Combustion Institute, Pittsburgh (1979). 111. DEAN, A. M., JOHNSON, R. L. and STEINER, D. C., Combust. Flame 37, 41 (1980). 112. VARDANYAN,1. A., SACHYAN,G. A., PHILIPOSYAN,A. G. and NALBANDYAN,A. B., Combust. Flame 22, 153 (1974). 113. SCHECKER, H. G. and JOST, W., Ber. Bunsenges. phys. Chem. 73, 521 (1969). 114. PELTERS, J. and MAHNEN, G., Fourteenth Symposium (International) on Combustion, p. 133, The Combustion Institute, Pittsburgh (1973). 115. TABAYASHI,K. and BAUER,S. H., Combust. Flame 34, 63 (1979). 116. JUST,TH. and TROE, J., d. phys. Chem. 84, 3068 (1980). 117. BENSON,S. W. and O'NEAL, H. E., National Bureau of Standards publication NSRDA-NBS 21 (1970). 118. WARNATZ, J., Combust. Sci. Technol., to be published (1983). 119. CHERIAN,M. A., RHODES, P., SIMPSON, R. J. and DIXONLEWIS, G., Eighteenth Symposium (International) on Combustion, p. 385, The Combustion Institute, Pittsburgh (1981). 120. SEERY, D. J. and BOWMAN,C. T., Combust. Flame 14, 37 (1970). 121. MARTENEY,P. J., Combust. Sci. Technol. 1,461 (1970). 122. BOWMAN,C. T., Fourteenth Symposium (International) on Combustion, p. 729, The Combustion Institute, Pittsburgh (1973). 123. BOWMAN, B. R, PRATT, D. T. and CROWE, C. T., Fourteenth Symposium (International) on Combustion, p. 819, The Combustion Institute, Pittsburgh (1973). 124. SKINNER,G. B., LIESH1TZ,A., SCHELLER,K. and BURCAT, A., J. chem. Phys. 56, 3853 (1972). t25. BRABBS,T. A. and BROKAW, R. S., Fifteenth Symposium (International) on Combustion, p. 893, The Combustion Institute, Pittsburgh (1975). 126. BHASKARAN, K. A., RAVIKUMAR, R., KARUPPANNAN, K. M. and NATARAJAN, K., First Specialists Meeting (International) of The Combustion Institute, p. 278 (1981). Also NATARAJAN, K. and BHASKARAN, K. A., Proceedings of the Thirteenth International Shock Tube Symposium, Niagara Falls (1981).

127. ENGLEMAN, V. S., EPA Report EPA-600/2-76-003 (1976). 128. WESTBROOK,C. K., Combust. Sci. Technol. 20, 5 (1979). 129. BOWMAN,C. T., Fifteenth Symposium (International) on Combustion, p. 869, The Combustion Institute, Pittsburgh (1975). 130. BURCAT,A., Combust. Flame 28, 319 (1977). 131. SMOOT, L. D., HECKER, W. C. and WILLIAMS, G. A., Combust. Flame 26, 323 (1976). 132. COOKE, D. F. and WILLIAMS, A., Thirteenth Symposium (International) on Combustion, p. 757, The Combustion Institute, Pittsburgh (1971 ). 133. D'SOuzA, M. V. and KARIM, G. A., Combust. Sci. Tecknol. 3, 83 (1971). 134. OLSON, D. B. and GARDINER, W. C., JR., Combust. Flame 32, 151 (1978). 135. HEAP, M. P., TYSON, T. J., CICHANOWICZ, J. E., GERSHMAN, R., KAN, C. J., MARTIN, G. B. and LANIER, W. S., Sixteenth Symposium (International) on Combustion, p. 535, The Combustion Institute, Pittsburgh (1977). 136. WARNATZ,J., Eighteenth Symposium (International) on Combustion, p. 369, The Combustion Institute, Pittsburgh (1981). 137. WARNATZ,J., Pro#. Astronaut. Aeronaut. 76, 501 (1981). 138. OLSON, D. B. and GARDINER, W. C., JR., J. phys. Chem. 81, 2514 (1977). 139. HAHN, W. A. and WENDT, J. O. L., Eiyhteenth Symposium (International) on Combustion, p. 121, The Combustion Institute, Pittsburgh (1981 ). 140. MIYAUCHI,T., MORI, Y. and YAMAGUCHI, T., Eighteenth Symposium (International) on Combustion, p. 43, The Combustion Institute, Pittsburgh (1981). 141. BECHTEL, J. H., BLINT, R. J., DASCH, C. J. and WEINBERGER,D. A., Combust. Flame 42, 197 (1981). 142. DIXON-LEwis, G., First Specialists Meeting (International) of The Combustion Institute, p. 284 (1981). 143, KRISHNAN, K. S,, RAVIKUMAR,R. and BHASKARAN, K. A, Combust. Flame 49, 41 (1983). 144. DIXON-LEwIS, G. and ISLAM, S. M., Nineteenth Symposium (International) on Combustion, p. 283, The Combustion Institute, Pittsburgh, 1983. 145. ROTH, P. and JUST, T., Bet. Bunsenyes. phys. Chem. 79, 682 (1975). 146. HARTIG, R., TROE, J. and WAGNER, H. GG., Thirteenth Symposium (International) on Combustion, p. 147, The Combustion Institute, Pittsburgh (1971 ). 147. HEFFINGTON,W. M., PARKS,G. E., SULZMANN,K. G. P. and Penner, S. S., Sixteenth Symposium (International) on Combustion, p. 997, The Combustion Institute, Pittsburgh (1977). 148. SEMENOV,N. N., Some Problems oJ" Chemical Kinetics and Reactivity, Pergamon, London (1958). 149. CLARK, T. C. and DOVE, J. E., Can. J. Chem. 51, 2147 (1973). 150. ZELLNER,R. and STEINERT,W., Int. J. chem. Kinet. 8, 397 (1976). 151. ROTH, P. and JUST, T., Bet. Bunsenges. phys. Chem. 81, 572 (1977). 152. BALDWIN,A. C. and GOLDEN, D. M., Chem. Phys. Lett. 55, 350 (1978). 153. BHASKARAN,K. A., FRANCK, P. and JUST, TH., Proc. 12th Int. Syrup. Shock Tubes Waves. Jerusalem: Magnus, Hebrew Univ. Press, p. 503 (1980). 154. Hsu, D. S. Y., SHAUB,W. M., CREAMER,T., GUTMAN, D. and LIN, M. C., to be published. 155. ROTH, P., BARNER, U. and LOHR, R., Bet. Bunsenges. phys. Chem. 83, 929 (1979). 156. ANDREWS,G. E. and BRADLEY,D., Combust. Flame 19, 275 (1972). 157. GARFORTH,A. M. and RALLIS, C. J., Combust. Flame 31, 53 (1978). 158. BURCAT, A., SCHELLER, K. and LIFSHITZ, A., Combust. Flame 16, 29 (1971).

Modeling of hydrocarbon combustion 159. BOWMAN,C. T., Combust. Sci. Technol. 2, 161 (1970). 160. SORENSON, S. C., MYERS, P. S. and U~nARA, O. A., Thirteenth Symposium (International) on Combustion, p. 451, The Combustion Institute, Pittsburgh (1971). 161. OLSON, D. B., TANZAWA,T. and GARDINER, W. C., JR., Int. J. chem. Kinet. 11, 23 (1979). 162. NOTZOLD, D. and ALGERMISSEN,J., Combust. Flame 40, 293 (1981). 163. NOTZOLD, D. and ALGERM1SSEN,J., Combust. Flame 40, 315 (1981). 164. LEE, W.-M. and YEH, C.-T., J. phys. Chem. 83, 771 (1979). 165. TULLY, F. P., RAVISHANKARA, A. R. and CARR, K, submitted to Int. J. chem. Kinet. 166. WHITE, J. N. and GARD1NER, W. C., JR., J. phys. Chem. 83, 562 (1979). 167. BALDWIN, R. R., SIMMONS, R. F. and WALKER, R. W., Trans. Faraday Soc. 62, 2476 (1966). 168. HOYERMANN, g. and SIEVERT, R., Seventeenth Symposium (International) on Combustion, p. 517, The Combustion Institute, Pittsburgh (1979). 169. PEETERS, J. and SMETS, B., First Specialists Meeting (International) of The Combustion Institute, p. 307 (1981). 170. TANZAWA, T. and GARDINER, W. C., JR., Combust. Flame 39, 241 (1980). 171. KIEFER, J. H., KAPSAL1S, S. A., AL-ALAMI, M. Z. and BUDACH, K. A., Combust. Flame 51, 79 (1983). 172. BACK, M. H. and MARTIN, R., Int. J. chem. Kinet. II, 757 (1979). 173. AYRANCI,G. and BACK, M. H., Int. J. chem. Kinet. 13, 897 (1981). 174. JUST, TH., RUTH, P. and DAMM, R., Sixteenth Symposium (International) on Combustion, p. 961, The Combustion Institute, Pittsburgh (1977). 175. WESTENBERG, A. A. and DE HAAS, N., Twelfth Symposium (International) on Combustion, p. 289, The Combustion Institute, Pittsburgh (1969). 176. NIKI, H., DABY, E. E. and WEINSTOCK,B., Twelfth Symposium (International) on Combustion, p. 277, The Combustion Institute, Pittsburgh (1969). 177. DAVIS, D. D., HUIE, R. E., HERRON, J. T., KURYLO, M. J. and BRAUN, W., J. chem. Phys. 56, 4868 (1972). 178. KLEMM, R. B. and SKOLNIK, E. G., ACS/CSJ Chem. Congress, Honolulu, Hawaii, paper 369. 179. BALDWIN, R. R. and WALKER, R. W., Eighteenth Symposium (International) on Combustion, p. 819, The Combustion Institute, Pittsburgh (1981). 180. SCnUG, K. P., SANTORO, R. J., DRYER, F. L. and GLASSMAN, I., Western States Section meeting of The Combustion Institute (1978). 181. BRADLEY, J. N., HACK, W., HOYERMANN, K. and WAGNER, H. GG., J. chem. Soc. London, Faraday Trans. 1 69, 1889 (1973). 182. BENSON,S. W. and HAUGEN, G. R., J. phys. Chem. 71, 1735 (1967). 183. JACHIMOWSKI,C. J., Combust. Flame 29, 55 (1977). 184. GARDINER, W. C., JR. and WALKER, B. F., J. chem. Phys. 48, 5279 (1968). 185. TANZAWA, T. and GARDINER, W. C., JR., Seventeenth Symposium (International) on Combustion, p. 563, The Combustion Institute, Pittsburgh (1979). 186. TANZAWA,T. and GARDINER, W. C., JR., J. phys. Chem. 84, 236 (1980). 187. SHAUB, W. M. and BAUER, S. H., Combust. Flame 32, 35 (1978). 188. BROWNE, W. G., PORTER, R. P., VERLIN, J. D. and CLARK, A. H., Twelfth Symposium (International) on Combustion, p. 1035, The Combustion Institute, Pittsburgh (1969). 189, PELTERS, J. and VINCKIER, C., Fifteenth Symposium (International) on Combustion, p. 969, The Combustion Institute, Pittsburgh (1975).

47

190. WARNATZ,J., BUCKHORN,H., MOSER, A. and WENZ,


H. W., Nineteenth Symposium (International) on Combustion, p. 197, The Combustion Institute, Pittsburgh (1983). 191. LEVY,J. M., Combust. Flame 46, 7 (1983). 192. LEVY, J. M., TAYLOR, B. R., LONGWELL, J. P. and SAROFIM, A. F., Nineteenth Symposium (International) on Combustion, p. 167, The Combustion Institute, Pittsburgh (1983). 193. BULL,D. C., Trans. Inst. chem. Engng. 57, 219 (1979). 194. VANDOOREN, J. and VAN T1GGELEN, P. J., Sixteenth Symposium (International) on Combustion, p. 1133, The Combustion Institute, Pittsburgh (1977). 195. WESTaROOK,C. K. and PITZ, W. J., to be published. 196. WESTBROOK,C. K. and PITZ, W. J., Ninth International Colloquium on Dynamics of Explosions and Reactive Systems (1983). 197. WESTBROOK, C. K., PITZ, W. J. and URTIEW, P. A., Ninth International Colloquium on Dynamics of Explosions and Reactive Systems (1983). 198. WAGNER,H. GG. and ZABEL,F., Ber. Bunsenqes. phys. Chem. 75, 114 (1971). 199. JONES, I. T. N. and BAYES, K. D., Proc. R. Soc. London A335, 547 (1973). 200. JONES, I. T. N. and BAYES, K. D., Fourteenth Symposium (International) on Combustion, p. 277, The Combustion Institute, Pittsburgh (1973). 201. LAUFER,A. H., Rev. chem. Intermed. 4, 225 (1981). 202. VINCKIER, C. and DEBRUYN, W., Seventeenth Symposium (International) on Combustion, p. 623, The Combustion Institute, Pittsburgh (1979). 203. LAUEER,A. H., d. phys. Chem. 85, 3828 (1981). 204. RENLUND, A. M., SHOKOOHL, F., REISLER, H. and WITTIG, C., J. phys. Chem. 86, 4165 (1982). 205. OKABE, H. and D1BELER,V. H., J. chem. Phys. 59, 2430 (1973). 206. ARONOWITZ,D., NAEGELI, O. W. and GLASSMAN, 1., J. phys. Chem. 81, 2555 (1977). 207. BOWMAN,C. T., Combust. Flame 25, 343 (1975). 208. ADELMAN,H. G., BROWNING, L. H. and PEFLEY, R. K., AIAA J. 14, 793 (1976). 209. CARELLI,G. and D'ALESSlO, A., Riv. Combust. 32, 125 (1978). 210. AKRICH, R., VOVELLE,C. and DELBOURGO,R., Combust. Flame 32, 171 (1978). 211. WESTBROOK,C. K. and DRYER, F. L., Proceedings of the Alcohol Fuels Technology Third International Symposium, Asilomar, California, May (1979). 212. PAUWELS,J.-F., CARLIER, M. and SOCHET, L.-R., First Specialists' Meetin 9 (International) of The Combustion Institute, 162 (1981). 213. NATARAJAN,K. and BHASKARAN,K. A., Combust. Flame 43, 35 (1981). 214. HOCKS, W., PETERS, N. and ADOME1T, G., Combust. Flame 41,157 (1981). 215. VANDOOREN, J. and VAN TIGGELEN, P. J., Eighteenth Symposium (International) on Combustion, p. 473, The Combustion Institute, Pittsburgh (1981 ). 216. TSUBOI,T. and HASHIMOTO,K., Combust. Flame 42, 61 (1981). 217. OLSSON, J. and ANDERSSON, L., Ninth International Colloquium on Dynamics of Explosions and Reactive Systems (1983). 218. ADERS, W. K., Combustion Institute European Symposium 19 (1973). 219. HENDERSON, H. T. and HILL, G. R., J. phys. Chem. 60, 874 (1956). 220. DEWILDE,E. and VAN TIGGELEN,A., Bull. Soc. Chim. Belges 77, 67 0968). 221. GAYDON, A. G. and WOLFHARD, H. G., Third Symposium on Combustion and Flame and Explosion Phenomena, p. 504, Williams and Wilkins, Baltimore (1949). 222. ARONOWITZ, D., SANTORO,R. J., DRYER, F. L. and GLASSMAN,I., Seventeenth Symposium (International) on

48

C.K. WESTBROOKand F. L. DRYER


Combustion, p. 633, The Combustion Institute, Pittsburgh (1979). HARVEY,R. and MACCOLL,A., Seventeenth Symposium (International) on Combustion, p. 857, The Combustion Institute, Pittsburgh (1979). HAYASHI,K. and DRYER,F. L., to be published. BEELEY,P., GRIFFITrfS,J. F., HUNT, B. A. and WILLIAMS, A., Sixteenth Symposium (International) on Combustion, p. 1013, The Combustion Institute, Pittsburgh (1977). COLKET,M. D., III, NAEGELI,D. W. and GLASSMAN,I., Sixteenth Symposium (International) on Combustion, p. 1023, The Combustion Institute, Pittsburgh (1977). HAUTMAN,D. J., SANTORO, R. J., DRYER, F. L. and GLASSMAN,I., Int. J. chem. Kinet. 13, 149 (1981). LIFSmTZ,A. and FRENKLACH,M., J. phys. Chem. 79, 686 (1975). BURC&T,A., Fuel 54, 87 (1975). CHINIa'z, W. and BAURER, T., Pyrodynamics 4, 119 (1966). McLAIN, A. G. and JACmMOWSKI,C. J., Star 15, 19, N77-28251, 2505 (1977). LAYOKUN,S. K. and SEATER,D. H., Ind. Engng. chem. Process Des. Dev. 18, 232 (1979). LAYOKUN,S. K. and SLATER,D. H., Ind. Engng. chem. Process Des. Dev. 18, 236 (1979). LAYOKUN,S. K., Ind. Engng. chem. Process Des. Dev. 18, 241 (1979). CATHONNET, M., BOETTNER, J. C. and JAMES, H., Eighteenth Symposium (International) on Combustion, p. 903, The Combustion Institute, Pittsburgh (1981 ). BARRONET,F., HALSTEAD,M. P., PROTHERO, A. and QUINN,C. P, C. R. Acad. Sci. Paris C17, 275 (1972). JACHIMOWSK1,C. J., Combust. Flame, to be published (1983). Also NASA Technical Paper 1794 (1980). CHIANG,C.-C. and SKINNER, G. B., Eighteenth Symposium (International) on Combustion, p. 915, The Combustion Institute, Pittsburgh (1981 ). AMANO,A. and UCHIYAMA,M., d. phys. Chem. 68, 1133 (1964). EDELSON,D. and ALLARA,D. L., Int. J. chem. Kinet. 12, 605 (1980). SUNDARAM,K. M. and FROMENT,G. F., Ind. Engng. Chem. Fundam. 17, 174 (1978). ALLARA,D. L. and EDELSON,D., Int. J. chem. Kinet. 7, 479 (1975). AL-ALAMI,M. Z. and KIEFER,J. H., J. phys. Chem. 87, 499 (1983). TULLY, F. P., RAVISHANKARA,A. R. and CARR, K., submitted for publication (1983). DRYER, F. L. and GLASSMAN,I., Alternative Hydrocarbon Fuels: Combustion and Chemical Kinetics, C. T. BOWMAN and J. BIRKELAND,eds., vol. 62, Pro#. Astr. Aeronaut. (1978). PERRY,R. A., submitted for publication (1983). KOIKE, T. and MORINAGA,K., Bull. chem. Soc. Japan 54, 2439 (1981). FRITZ, K. and GRONIG,H., Eleventh International Symposium on Shock Tubes and Waves, Seattle, University of Washington Press (1977). POWERS, D. R. and CORCORAN, W. H., Ind. Engng. chem. Fundam. 13, 351 (1974). COATS,C. M. and WILLIAMS,A., Seventeenth Symposium (International) on Combustion, p. 611, The Combustion Institute, Pittsburgh (1979). DOOLAN,K. R. and MACKIE,J. C., Combust. Flame 50, 29 (1983). EBERT, K. H., EDERER, H. J. and ISBARN,G., Int. J. chem. Kinet. 15, 475 (1983). ALLARA,D. L. and SHAW,R., J. phys. Chem. Ref. Data 9, 523 (1980). EDELSON,D. and ALLARA,D. L., Int. J. chem. Kinet. 12, 605 (1980). ASABA, T. and FuJn, N., Thirteenth Symposium (International) on Combustion, p. 155, The Combustion Institute, Pittsburgh (1971). FuJII, N. and ASABA,T., Fourteenth Symposium (International) on Combustion, p. 433, The Combustion Institute, Pittsburgh (1973). SANTORO, R. J. and GLASSMAN, I., Combust. Sci. Technol. 19, 161 (1979). VENKAT,C., BREZINSKY,K. and GLASSMAN,I., Nineteenth Symposium (International) on Combustion, p. 143, The Combustion Institute, Pittsburgh (1983). NORRISH,R. G. W. and TAYLOR,G. W., Proc. R. Soc. London A234, 160 (1956). BENSON,S. W., J. Am. chem. Soc. 87, 972 (1965). BITTNER,J. D. and HOWARD, J. B., Eighteenth Symposium (International) on Combustion, p. 1105, The Combustion Institute, Pittsburgh (1981). LITZINGER, T., BREZINSKY,K. and GLASSMAN, 1., Eastern States Section meeting of The Combustion Institute, Paper 69, December (1982). KERN, R. D., SINGH, H. J., ESSLINGER, M. A. and WINKLER,P. W., Nineteenth Symposium (International) on Combustion, p. 1351, The Combustion Institute, Pittsburgh (1983). FuJII, N., MIYAMA, n., KOSHI, M. and ASABA, T., Eighteenth Symposium (International) on Combustion, p. 873, The Combustion Institute, Pittsburgh (1981). YUMURA, M. and ASABA, T., Eighteenth Symposium (International) on Combustion, p. 863, The Combustion Institute, Pittsburgh (1981 ). MILLER,J. A., SMOOKE,M. D., GREEN,R. M. and KEE, R. J., Combust. Sci. Technol., to be published (1983). YUMURA, M., ASABA, T., MATSUMOTO, Y. and MATSUkH., Int. J. chem. Kinet. 12, 439 (1980). BRANCH,M. C., KEE, R. J. and MILLER,J. A., Combust. Sei. Technol. 29, 147 (1982). LYON,R. K., Int. J. chem. Kinet. 8, 315 (1976). LYON, R. K. and BENN, D. J., Seventeenth Symposium (International) on Combustion, p. 601, The Combustion Institute, Pittsburgh (1979). MuzIo, L. J., ARAND, J. K. and TEIXEIRA, D. P., Sixteenth Symposium (International) on Combustion, p. 199, The Combustion Institute, Pittsburgh (1977). FENIMORE, C. P., Combust. Flame 37, 245 (1980). SALIMIAN, S. and HANSON,R. K., Combust. Sci. Technol. 23, 225 (1980). MILLER,J. A., BRANCh, M. C. and KEE, R. J., Combust. Flame 43, 81 (1981). AZUHATA,S., KAJI, R., AKIMOTO,H. and HISHINUMA, Y., Eighteenth Symposium (International) on Combustion, p. 845, The Combustion Institute, Pittsburgh (1981). ROOSE, T. R., HANSON, R. K. and KRUGER, C. H., Eighteenth Symposium (International) on Combustion, p. 853, The Combustion Institute, Pittsburgh (1981). DEAN,A. M., HARDY,J. E. and LYON,R. K., Nineteenth Symposium (International) on Combustion, p. 97, The Combustion Institute, Pittsburgh (1983). HOWGATE,D. W. and BARR, T. A., JR., J. chem. Phys. 59, 2815 (1973). HARDY, J. and GARDINER,W. C., JR., Sixteenth Symposium (International) on Combustion, p. 985, The Combustion Institute, Pittsburgh (1977). LIFSHITZ, A., FRENKLACH,M., SCHECHNER, P. and CARROLL,H. F., Int. J. chem. Kinet. 7, 753 (1975). BOWMAN,C. T. and DODGE,L. G., Sixteenth Symposium (International) on Combustion, p. 971, The Combustion Institute, Pittsburgh (1977). MERRYMAN, E. L. and LEVY,A., J. Air Poll. Cont. Assoc. 17, 800 (1967). MERRYMAN,E. L. and LEVY,A., J. phys. Chem. 76, 1925 (1971). BERNEZ-CAMBOT, J., VOVELLE,C. and DELBOURGO,R., Eighteenth Symposium (International) on Combustion, p. 777, The Combustion Institute, Pittsburgh (1981).

223. 224. 225.

256. 257.

258.
259. 260. 261. 262. 263.

226. 227. 228. 229. 230. 231. 232.

264. 265. 266. 267. 268. 269. 270. 271. 272. 273. 274. 275.

233.
234. 235. 236. 237. 238. 239. 240. 241. 242. 243. 244, 245.

276. 277. 278. 279.

246. 247. 248. 249. 250. 251. 252. 253. 254. 255.

280.
281. 282. 283. 284.

Modeling of hydrocarbon combustion

49

285. ROTH, P., LOHR, R. and BARNER,U., Combust. Flame 315. LOVACHEV,L. A. and LOVACHEV,L. N., Combust. Sci. 45, 273 (1982). Technol. 19, 195 (1979). 286. FRENKLACH,M., LEE,J. H., WHITE,J. N. and GARDINER, 316. LOVACHEV,L. A. and LOVACHEV,L. N., Combust. Sci. W. C., JR., Combust. Flame 41, 1 (1981). Technol. 22, 93 (1980). 287, WENDT, J. O. L., WOOTAN,E. C. and CORLEY,T. L., 317. LOVACnEV,L. A. and LOVACHEV,L. N., Combust. Sci. Combust. Flame 49, 261 (1983). Technol. 23, 181 (1980). 288, TAKAGI,T., TATSUMI,T. and OGASAWARA, M., Combust. 318. WESTaROOK, C. K., Paper presented at the Western Flame 35, 17 (1979). States Section Meeting of The Combustion Institute, Tempe, Arizona, October, 1981. Lawrence Livermore 289, MULVIHILL,J. N. and PHILLIPS, L. F., Fifteenth SymNational Laboratory report UCRL-86456 (1981). posium (International) on Combustion, p. 1113, The 319. WESTBROOK,C. K., BEASON,D. G. and ALVARES,N. J., Combustion Institute, Pittsburgh (1975). First Specialists Meetin 0 (International) of The Combus290, HAYNES,B. S., Combust. Flame 28, 113 (1977). 291. L]FSH]TZ,A. and FRENKLACH,M., Int. J. chem. Kinet. tion Institute, p. 196 (1981). 320. WESTBROOK, C. K., Nineteenth Symposium (Interna12, 159 (1980). 292. SONG,Y. H., BLAIR,D. W., SIMINSKI,V. J. and BARTOK, tional) on Combustion, p. 127, The Combustion Institute, W., Eighteenth Symposium (International) on ComPittsburgh (1983). bustion, p. 53, The Combustion Institute, Pittsburgh 321. WESTBROOK,C. K. and URTIEW, P. A., Fiz. Goren. i (1981). Vzry., in press (1983). 293. BEER, J. M., JACQUES, M. T., FARMAYAN, W. and 322. BIORDI,J. C., LAZZARA,C. P. and PAPa, J. F., J. phys. Chem. 80, 1042 (1976). TAYLOR, B. R., Eighteenth Symposium (International) on Combustion, p. 101, The Combustion Institute, 323. BIORDI,J. C., LAZZARA,C. P. and PAPa, J. F., J. phys. Chem. 81, 1139 (1977). Pittsburgh (1981 ). 294. PETERSON,R. C. and LAURENDEAU,N. M., paper pre- 324. BIORDI,J. C., LAZZARA,C. P. and PAPa, J. F., J. phys. sented at the Central States Section of the Combustion Chem. 82, 125 (1978). Institute, Columbus, Ohio, March (1982). 325. BIORDI,J. C., LAZZARA,C. P. and PAPa, J. F., Combust. 295. FIFES, R. A. and HOLMES, H. E., J. phys. Chem. 86, Flame 24, 401 (1975). 2935 (1982). 326. BXORDI, J. C., LAZZARA,C. P. and PAPa, J. F., Sixteenth 296. TRICOT,J. C., PERCHE, A. and LUCQU]N,M., Combust. Symposium (International) on Combustion, p. 1097, The Flame 40, 269 (1981). Combustion Institute, Pittsburgh (1977). 297. GUmGUIS, R., Hsu, D., BOGAN, D. and ORAN, E. S., 327. BIORDI,J. C., LAZZARA,C. P. and PAPa, J. F., Fifteenth Ninth International Colloquium on Dynamics of ExploSymposium (International) on Combustion, p. 917, The sions and Reactive Systems (1983). Combustion Institute, Pittsburgh (1975). 298. FIEER,R. A., Seventeenth Symposium (International) on 328. SKINNER,G. B., ACS Symposium Series, No. 16, HaloCombustion, p. 587, The Combustion Institute, Pittsgenated Fire Suppressants, R. G. GANN, ed. Am. Chem. burgh (1979). Soc., 295 (1975). 299. WENDT, J. O. L., MORCOMB,J. T. and CORLEY,T. L., 329. SKINNER,G. B. and RINGROSE,G. H., J. chem. Phys. 42, Seventeenth Symposium (International) on Combustion, 2190 (1965). p. 671, The Combustion Institute, Pittsburgh (1979). 330. SKINNER,G. B. and RINGROSE,G. H., J. chem. Phys. 43, 300. MULLER,C. H., III, SCHOFIELD,K., STEINBERG,M. and 4129 (1965). BROIDA, H. P., Seventeenth Symposium (International) 331. WESTBROOK,C. K, Combust. Sci. Technol., in press on Combustion, p. 867, The Combustion Institute, (1983). Pittsburgh (1979). 332. HOMANN,K. H. and Poss, R., Combust. Flame 18, 300 301. ATTAR,A., Fuel 57, 201 (1978). (1972). 302. CULL]S,C. F. and MULCAHY,M. F. R., Combust. Flame 333. BAaKIN,V. S. and VYUN, A. V., Fiz. Goren. i Vzry. 17, 15, 225 (1972). 8 (1981). 303. DEAN,A. M., STEINER,D. C. and WANG,E. E., Combust. 334. SMITH, O. I., WANG, S.-N., TSEREGOUNIS, S. and Flame 32, 73 (1978). WESTaROOK, C. K., Combust. Sci. Technol. 30, 241 304. BALAKHN]NE, V. P., VANDOOREN,J. and VAN T]GGELEN, (1983). P. J., Combust. Flame 28, 165 (1977). 335. CROSSLEY, R. W., DORKO, E. A., SCHELLER, K. and 305. DEAN, A. M. and JOHNSON,R. L., Combust. Flame 37, BURCAT,A., Combust. Flame 19, 373 (1972). 109 (1980). 336. BULL,D. C., ELSWORTH,J. E. and HOOPER,G., Combust. 306. MILKS, D. and MATULA,R. A., Fourteenth Symposium Flame 34, 327 (1979). (International) on Combustion, p. 83, The Combustion 337. VANDERMOLEN,R. and NICHOLLS,J. A., Combust. Sci. Institute, Pittsburgh (1973). Technol. 21, 75 (1979). 307. LomAr, H., CARALa, F. and DEST~aAU, M., First 338. EUBANK, C. S., RABINOWITZ,M. J., GARDINER,W. C., JR., Specialists Meetin# (International) of The Combustion and ZELLNER, R. E., Eighteenth Symposium (InternaInstitute, p. 267 (1981). tional) on Combustion, p. 1767, The Combustion 308. FEN]MORE,C. P., Seventeenth Symposium (International) Institute, Pittsburgh (1981). on Combustion, p. 661, and references cited therein, The 339. LIFSHITZ,A., SCULLER, K., BURCAT,A. and SKINNER, Combustion Institute, Pittsburgh (1979). G. B., Combust. Flame 16, 311 (1971). 309. WESrBROOK, C. K., Combust. Sci. Technol. 23, 191 340. EUBANK,C. S., GARDINER,W. C., JR. and SIMMIE,J. M., (1980). First Specialists Meeting (International) of The Com310. DAY, U. J., STAMP, D. V., THOMPSON,K. and DIXONbustion Institute, p. 336 (1981). LEWIS, G., Thirteenth Symposium (International) on 341. GRILLO, A. and SLACK, M. W., Combust. Flame 27, Combustion, p. 705, The Combustion Institute, Pitts377 (1976). burgh (1971). 342. DORKO, E. A., BASS, D. M., CROSSLEY, R. W. and 311. DIXON-LEWIS,G. and SIMPSON, R. J., Sixteenth SymSChiLLER, K., Combust. Flame 24, 173 (1975). posium (International) on Combustion, p. 1111, The 343. DABORA,E. K., Combust. Flame 24, 181 (1975). Combustion Institute, Pittsburgh (1977). 344. SLACK,M. W. and GR]LLO, A. R., Combust. Flame 40, 312. D]XON-LEWlS,G., Combust. Flame 36, 1 (1979). 155 (1981). 313. LOVACHEV,L. A., GONTKOVSKAYA, V. T. and OZERKOV- 345. WESTBROOK,C. K. and HASELMAN,L. C., Pro#. AstroSKAYA,N. I., Combust. Sci. Technol. 17, 143 (1977). naut. Aeronaut. 75, 193 (1981). 314. LOVACh~V,L. A. and LOVACHEV,L. N., Combust. Sci. 346. WESTBROOK, C. K. and PITZ, W. J., Combust. Sci. Technol. 18, 191 (1978). Technol., in press (1983). JPECSIO:I-D

50

C.K. WESTBROOKand F. L. DRYER

347. WESTBROOK,C. K. and DRYER, F. L., Combust. Sci. Technol. 27, 31 (1981). 348. DUGGER, G. L., SIMON, D. M. and GERSTEIN, M., Chapter 4 in NACA Report No. 01300 (1959). 349. DRYER, F. L. and WESTBROOK,C. K., NATO-AGARD Conference Proceedings No. 275 (1979). 350. MARTENEY, P. J. and KESTEN, A. S., Eighteenth Symposium (International) on Combustion, p. 1899, The Combustion Institute, Pittsburgh (1981). 351. ABDALLA,A. Y., BRADLEY,D., CHIN, S. B. and LAM, C., Second International Specialists Meeting of the Combustion Institute on Oxidation, Budapest, Hungary (1982). 352. EDELMAN, R. B. and FORTUNE, O. F., AIAA paper 69-86 (1969). 353. MELLOR,A. M., Emissions from Continuous Combustion Systems, W. CORNELIUSand W. G. AGNEW,eds., Plenum Press, New York (1972). 354. ROBERTS, R., ACETO, L. D., KOLLRACK, R., TEIXEIRA, D. P. and BONNELL,J. M., AIAA J. 10, 820 (1972). 355. EDELMAN, R. B. and HARSHA, P. T., Prog. Energy Combust. Sci. 4, 1 (1978). 356. DUTERQUE, J., BORGHI,R. and TICHTINSKY,H., Combust. Sci. Technol. 26, 1 (1981). 357. ABDALLA, A. Y., ALI, B. B., BRADLEY, D. and CHIN, S. B., Combust. Flame 43, 131 (1981). 358. HAUTMAN, D. L., DRYER, F. L., SCHUG, K. P. and GLASSMAN,I., Combust. Sci. Technol. 25, 219 (1981). 359. GLASSMAN,I., DRYER, F. L. and COHEN, R., presente.d at Joint Meeting of the Central and Western States Sections of The Combustion Institute, San Antonio, Texas (1975). 360. BOWMAN,C. T. and HANSON, R. K., J. phys. Chem., 83, 757 (1979). 361. LIBOUTON,J.-C., DORMAL, M. and VAN TIGGELEN,P. J., Prog. Astronaut. Aeronaut. 75, 358 (1981). 362. LIBOUTUN,J.-C., JACQUES,A. and VAN TIGGELEN, P. J., First Specialists" Meeting (International) of The Combustion Institute, p. 437 (1981 ). 363. BULL,D. C., ELSWORTH,J. E. and HOOPER, G., Combust. Flame 35, 27 (1979). 364. BULL,D. C., ELSWORTH,J. E., HOOPER, G. and QUINN, C. P., J. Phys. D. Appl. Phys. 9, 1991 (1976). 365. ZELDOVICH, Y. B., J. exp. theor. Phys. (USSR) 10, 542 (1940). 366. YON NEUMANN, J., Office of Scientific Research and Development report No. 549 (1942). 367. DORING,W., Ann. Physik 43, 421 (1943). 368. LEE, J. H., KNYSTAUTAS, R., GUmAO, C., BENEDICK, W. B. and SHEPHERD, J. E., Second International Workshop on the Impact of Hydrogen on Water Reactor Safety, Albuquerque, New Mexico (1982). 369. STREHLOW,R. A., MAURER, R. E. and RAJAN, S., AIAA J. 7, 323 (1969). 370. STREHLOW, R. A. and ENGEL, C. D., AIAA J. 7, 492 (1969). 371. BARTHEL,H. O., Phys. Fluids 17, 1547 (1974). 372. LEE, J. H. S. and RAMAMURTHhK., Combust. Flame 27, 331 (1976). 373. MITROFANOV,V. V. and SOLOUKHIN,R. I., Soviet Phys. Dokl. 9, 1055 (1964). 374. LEE,J. H. S., A. Re~. phys. Chem. 28, 75 (1977). 375. URTIEW, P. A. and TARTER, C. M., Prog. Astronaut. Aeronaut. 75, 370 (1981). 376. BULL,D. C., ELSWORTH,J. E. and SHUFF,P. J., Combust. Flame 45, 7 (1982). 377. MATSUI, H. and LEE, J. H., Seventeenth Symposium (International) on Combustion, p. 1269, The Combustion Institute, Pittsburgh (1979). 378. MOEN, I. O., DONATO, M., KNYSTAUTAS, R. and LEE, J. H., Eighteenth Symposium (International) on Combustion, p. 1615, The Combustion Institute, Pittsburgh (1981). 379. MOEN, I. O., MURRAY, S. B., BJERKETVEDT,O., RINNAN, A., KNYSTAUTAS, R. and LEE, J. H., Nineteenth Sym-

posium (International) on Combustion, p. 635, The Combustion Institute, Pittsburgh (1983). 380. KNYSTAUTAS,R., LEE,J. H. and GUIRAO,C. M., Combust. Flame 48, 63 (1982). 381. ZELDOVICH,Y. B., KOGARKO,S. M. and SEMENOV,N. N., Soviet Phys. tech. Phys. 1, 1689 (1956). 382. BONI, A. A., WILSON, C. W., CHAPMAN,M. and CooK, J. L., Acta astr. 5, 1153 (1978). 383. HOGAN, W. J., ERMAK, D. L. and KOOPMAN, R. P., Proceedings of the 1982 Hazardous Materials Spill Conference, Milwaukee, Wisconsin, April 1982. Lawrence Livermore National Laboratory report UCRL-86164 (1982). 384. KOOPMAN, R. P., CEDERWALL, R. T., ERMAK, D. L., GOLDWIRE, H. C., HOGAN, W. J., McCLURE, J. W., MCRAE, Z. G., MORGAN, D. L., RODEAN, H. C. and SI-IINN,J. H., J. Hazardous Mater. 6, 43 (1982). 385. HARDY, J. E. and LYON, R. K., Combust. Flame 39, 317 (1980). 386. SAROEIM,A. F. and BEER, J. M., Seventeenth Symposium (International) on Combustion, p. 189, The Combustion Institute, Pittsburgh (1979). 387. LYON, R. K. and HARDY, J. E., Combust. Flame 45, 209 (1982). 388. DRYER, F. L., Ph.D. thesis, Dept. of Aerospace and Mechanical Engineering, Princeton University (1972). 389. LONGWELL,J. P. and WEISS, M. A., Ind. Engng. Chem. 47, 1634 (1955). 390. HOTTEL, H. C., WILLIAMS,G. C., NERHEIM,N. M. and SCHNEIDER, G. R., Tenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh (1965). 391. ENGLEMAN, V. S., BARTOK, W., LONGWELL, J. P. and EDELMAN, R. B., Fourteenth Symposium (International) on Combustion, The Combustion Institute, Pittsburgh (1973). 392. EDELMAN, R. l., FORTUNE, O. and WEILERSTEIN, G., Emissions from Continuous Combustion Systems, W. CORNELIUSand W. G. AGNEW,eds., Plenum Press, New York (1972). 393. GASKELL, P. H., GOLDSWORTHY, F. A. and DIXONLEWIS, G., First Specialists' Meeting (International) of The Combustion Institute, p. 261 (1981). 394. HEIMERL, J. M. and COFFEE, T. P., Combust. Sci. Technol., in press (1983). 395. HIRSCHFELDER, J. O., CURTISS, C. F. and BIRD, R. B., Molecular Theory of Gases and Liquids, Wiley and Sons, New York (1967). 396. BIRD, R. B., STEWART, W. S. and LIGHTEOOT, E. N., Transport Phenomena, Wiley, New York (1960). 397. CHAPMAN, S. and COWLING~T. G., The Mathematical Theory of Non-Uniform Gases, Cambridge University Press, Cambridge (1970). 398. MONCHICK,L., MUNN, R. J. and MASON, E. A., J. chem. Phys. 45, 3051 (1966) and references cited therein. 399. SPALDING, D. B. and STEPHENSON, P. L., Proc. R. Soc. London A324, 315 (1971). 400. COFFEE, T. P. and HEIMERL, J. M., Combust. Flame 43, 273 (1981). 401. SMOOKE,M. D., MILLER,J. A. and KEE, R. J., Proceedings of the GAMM Conference on Premixed Laminar Flames, p. 112, N. PETERS and J. WARNATZ, eds. (1983). 402. SPALDING,D. B., STEPHENSON,P. L. and TAYLOR, R. G., Combust. Flame 17, 55 (1971). 403. Proceedings of the GAMM Conference on Premixed Laminar Flames, N. PETERS and J. WARNATZ, eds., Vieweg Verlag, Wiesbaden (1983). 404. WARNATZ, J., Combust. Sci. Technol. 26, 203 (1981). WARNATZ, J., Ber. Bunsenges. phys. Chem. 83, 950 (1979). 405. LUND, C. M., University of California Lawrence Livermore Laboratory report UCRL-52504 (1978). 406. KOOKER, D. E., Seventeenth Symposium (International)

Modeling of hydrocarbon combustion on Combustion, p. 1329, The Combustion Institute, Pittsburgh (1979). 407. FRISTROM, R. M., GRUNFELDER, C. and FAVIN, S., J. phys. Chem. 64, 1386 (1960). 408. CATTOLICA,R. J., Combust. Flame 44, 43 (1982). 409. HAHN, W. A., WENDT, J. O. L. and TYSON, T. J., Combust. Sci. Technol. 27, 1 (1981). 410. GILBERT, M., Sixth Symposium (International) on Combustion, p. 74, Reinhold, New York (1957). 411. SMITH,D. and At3NEW,J. T., Sixth Symposium (International) on Combustion, p. 83, Reinhold, New York (1957). 412. HERTZBERG, M., Proceedings of the International Conference on Fuel-Air Explosions, J. H. S. LEE and C. M. GUIRAO, eds., University of Waterloo Press (1982). HERTZBERG, M., U.S. Bureau of Mines reports 8127 (1976), 8469 (1980), and 8607 (1982). 413. MACEK,A., Combust. Sci. Technol. 21, 43 (1979). 414. LOVACHEV,L. A., Combust. Sci. Technol. 20, 209 (1979). LOVACHEV, L. A., BABKIN, V. S., BUNEV, V. A., V'VUN, A. V. and KRIVULIN, V. N., Combust. Flame 20, 259 (1973). 415. SMITH, O. I., WESTBROOK, C. K. and SAWYER, R. F., Seventeenth Symposium (International) on Combustion, p. 1305, The Combustion Institute, Pittsburgh (1979). 416. QUADER,A. A., Society of Automotive Engineers paper SAE 760760 (1976). 417. DANIEL, W. A., Sixth Symposium (International) on Combustion, p. 886, Reinhold, New York (1957). 418. DANIEL, W. A. and WENTWORTH, J. T., Society of Automotive Engineers Technical Progress Series, Vol. 6, Vehicle Emissions, New York (1964). 419. DANIEL,W. A., Society of Automotive Engineers paper SAE 700108 (1970). 420. WESTBROOK, C. K., ADAMCZYK, A. A. and LAVOIE, G. A., Combust. Flame 40, 81 (1980). 421. ADAMCZYK,A. A. and LAVO1E,G. A., SAE paper SAE 780969, SAE Trans. 87 (1978). 422. SLOANE,T. M. and SCHOENE, A. Y., Combust. Flame 49, 109 (1983). 423. BLINT, R. J. and BECHTEL,J. H., Combust. Sci. Teehnol. 27, 87 (1982). 424. BLINT, R. J. and BECHTEL,J. H., Society of Automotive Engineers publication SAE-820063 (1982). 425. BERGNER,P., EBERIUS,H. and POKORNY,H., Proceedings of the Alcohol Fuels Technology Third International Symposium, Asilomar, California (1979). 426. ADAMCZYK,A. A., KAISER, E. W., CAVOLOWSKY,J. A. and LAVOIE,G. A., Eighteenth Symposium (International) on Combustion, p. 1695, The Combustion Institute, Pittsburgh (1981). 427. LoRusso, J. A., LAVOIE, G. A. and KAISER, E. W., Eastern States Section of The Combustion Institute, Princeton, New Jersey (1979). 428. SELLNAU, M. C., SPRINGER, G. S. and KECK, J. C., Society of Automotive Engineers publication SAE810148 (1981). 429. WEISS, P. and KECK, J. C., Society of Automotive Engineers publication SAE-810149 (1981). 430. SLOANE, T. M. and RATCLIFFE, J. W., Combust. Flame 47, 83 (1982). 431. LAVOIE, G. A., LoRusso, J. A. and ADAMCZYK, A. A., General Motors Symposium on Combustion Modelling in Reciprocating Engines, Warren, Michigan (1978). 432. KEIPER, R. and SPURK, J. H., J. Fluid Mech. 113, 333 (1981). 433. LANCASTER, D. R., Proceedings of Conference on Stratified Charge Automotive Engines, I. Mech. E. Paper C397/80, I. Mech. E. Publication 1980-9, p. 33 (1980). 434. WESTBROOK,C. K., Acta astr. 5, 1185 (1978). 435. WESTBROOK, C. K., Society of Automotive Engineers paper SAE 790248 (1979). 436. DIWAKAR, R., Society of Automotive Engineers publi-

51

cations SAE-810225 (1981 ) and SAE-820039 (1982). 437. ATHERTON,R. W., SCHAINI~ER,R. B. and DUCOT, E. R., AIChE J. 21,441 (1975). 438. DICKENSON, R. P. and GELINAS, R. J., J. computational Phys. 21, 123 (1976). 439. CUKIER, R. 1., FORTUIN, C. M., SHrdLER, K. E., PETSCHEK, A. G. and SCHAmLY, J. H., J. chem. Phys. 59, 3873 (1973). 440. SCHAIBLY,J. H. and SHULER, K. E., J. chem. Phys. 59, 3879 (1973). 441. CUKIER, R. I., SCHAmLY, J. H. and SHULER, K. E., J. chem. Phys. 63, 1140 (1975). 442. CUKIER, R. I., LEVINE, H. B. and SHULER, K. E., J. computational Phys. 26, 1 (1978). 443. HWANG, J.-T., DOUGHERTY, E. P., RABITZ, S. and RAaITZ, H., J. chem. Phys. 69, 5180 (1978). 444. DOUGHERTY, E. P., HWANG, J.-T. and RAmTZ, H., J. chem. Phys. 71, 1794 (1979). 445. DOUGHERTY,E. P. and RABITZ, H., Int. J. chem. Kinet. II, 1237 (1979). 446. KRAMER, M. A., KEE, R. J. and RABITZ, H., Sandia National Laboratories report SAND82-8230 (1982); and KRAMER, M. A., CALO, J. M., RABITZ, H. and KEE, R. J., Sandia National Laboratories report SAND82-8231 (1982). 447. BONI, A. A. and PENNER, R. C., Combust. Sci. Technol. 15, 99 (1976). 448. TEETS,R. E. and BECHTEL,J. H., Eighteenth Symposium (International) on Combustion, p. 425, The Combustion Institute, Pittsburgh (1981). 449. MILLER,J. A., J. chem. Phys. 74, 5120 (198l). 450. MORETTI,G., AIAA J. 3, 223 (1965). 451. LLOYD,A. C., Int. J. chem. Kinet. 6, 169 (1974). 452. BAULCH, D. L., DRYSDALE, D. D. and LLOYD, A. C., High Temperature Reaction Rate Data, Report No. 3, Leeds University (1969). 453. BAHN,G. S., Pyrodynamics 2, 315 (1965). 454. BALDWIN, R. R., JACKSON, D., WALKER, R. W. and WEBSTER,S. J., Trans. Faraday Soc. 63, 1665 (1967). 455. HEIMERL,J. M. and COFFEE, T. P., Combust. Flame 35, 117 (1979). 456. HAMPSON,R, F., J. phys. Chem. Ref. Data 2, 267 (1973). 457. LEE, J. H., MICHAEL, J. V., PAYNE, W. A. and STIEF, L. J., J. chem. Phys. 69, 350 (1978). 458. RAVISHANKARA, A. R., WINE, P. H. and LANGFORD, A. 0., J. chem. Phys. 70, 984 (1979). 459. SIMONAmS, R. and HEICKLEN, J., J. chem. Phys. 56, 2004 (1972). 460. BAULCH,D. L. and DRYSDALE, D. D., Combust. Flame 23, 215 (1974). 461. ATRI,G. M., BALDWIN,R. R., JACKSON,D. and WALKER, R. W., Combust. Flame 30, 1 (1977). 462. ATKINSON, R. and P~TTS, J. N., J. chem. Phys. 68, 3581 (1978). 463. WESTENSERG,A. A. and DE HAAS, N., J. phys. Chem. 76, 2215 (1972). 464. SKINNER,G. B., LIFSIaITZ,A., SCHELLER,K. and BURCAT, A., d. chem. Phys. 56, 3853 (1972). 465. PACEY,P. D., Chem. Phys. Lett. 23, 394 (1973). 466. FENIMORE,C. P., Twelfth Symposium (International) on Combustion, p. 463, The Combustion Institute, Pittsburgh (1969). 467. BHAS~ARAN,K. A., FRANCK, P. and JUST, TH., Twelfth International Symposium on Shock Tubes and Waves, 503 (1979). 468. TENDER, R., MAYER, S., COOK, E. and SCHILLER, L., Aerospace Corporation report TR-1001(9210-02)-1 (AD813485). 469. HOYERMANN, K., LOFTFIELO, N. S., SIEVERT, R. and WAGNER, H. GG., Eighteenth Symposium (International) on Combustion, p. 831, The Combustion Institute, Pittsburgh (1981 ). 470. TAYLOR,J. E. and KULICH, D. M., Int..1. chem. Kinet. 5, 455 (1973).

52

C . K . WESTBROOKand F. L. DRYER 483. AYUR, A. L. and RoscoE, J. M., Can. J. Chem. 57, 1269 (1979). 484. KERR, J. A. and PARSONAGE, M. J., Evaluated Kinetic Data on Gas Phase Hydrogen Transfer Reactions of Methyl Radicals, Butterworths (1976). 485. KERR, J. A., Chem. Rev. 66, 469 (1966). 486. WATKINS, K. W. and WORD, W. W., Int. J. chem. Kinet. 6, 855 (1974). 487. WALKER,R. W., Reaction Kinetics, vol. l, S.P.R. Chem. Soc. (1975) and vol. 2, S.P.R. Chem. Soc. (1977). 488. JACKSON, W. M. and McNESBY, J. R., J. Am. chem. Soc. 83, 4891 (1961). 489. KERR, J. A. and TROTMAN-DICKENSON, A. F., Trans. Faraday Soc. 55, 921 (1959). 490. SZWARC, M., J. chem. Phys. 17, 284 (1949). 491. HUIE, R. E. and HERRON, J. T., Progress in Reaction Kinetics, K. R. JENNINGS and R. B. CUNDALL, eds., vol. 8, p. 1, P e r g a m o n Press (1978). 492. CVEXANOVlC,R. J. and IRWIN, R. S., J. chem. Phys. 46, 1694 (1967). 493. BAKER, R. R., BALDWIN, R. R., FULLER, A. R. and WALKER, R. W., J. Chem. Soc., Faraday Trans. I 71, 736 (1975). 494. RAVlSHANKARA,A. R., WAGNER, S., FISCHER, S., SMITH, G., SCHIFF, R., WATSON, R. T., TESI, G. and DAV~S, D. D., Int. J. chem. Kinet. 10, 783 (1978).

471. FRANCK, P., to be published. 472. HOYERMANN,K. H., Comments, Eighteenth Symposium (International) on Combustion, p. 382, The Combustion Institute, Pittsburgh (1981). 473. LOHR, R. and ROTH, P., Ber. Bunsenges. phys. Chem. 85, 153 (1981). 474. SMITH, I. W. M. and ZELLNER, R., J. Chem. Soc., Faraday Trans. II 69, 1617 (1973). 475. MAYER, S. W., SCHILLER, L. and JOHNSTON, H. S., Eleventh Symposium (International) on Combustion, p. 837, The Combustion Institute, Pittsburgh (1967). 476. BRAUN, W., BASS, A. M. and PILLING, M. J., J. chem. Phys. 52, 5131 (1'970). 477. LANGE, W. and WAGNER, H. GG., Ber. Bunsenges. phys. Chem. 79, 165 (1975). 478. LEFEVRE, H. F., MEAGHER, J. F. and TIMMONS, R. B., Int. J. chem. Kinet. 4, 103 (1972). 479. GRAY, P. and HEROD, A. A., Trans. Faraday Soc. 64, 2723 (1968). 480. NATARAJAN, K. and BHASKARAN,K. A., Thirteenth International Shock Tube Symposium, Niagara Falls, U.S.A. (1981). 481. Estimated by analogy with methanol reactions. 482. ADERS, W. K. and WAGNER, H. GG., Ber. Bunsenges. phys. Chem. 77, 712 (1973).

APPENDIX I. Hydrocarbon oxidation mechanism, reaction rates in cm 3 mol sec kcal units, k = ATnexp(-E~/RT) Forward rate Reaction 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. H + O 2 --* O + O H O + H 2 --* H + O H H 2 + O H --* H z O + H O +HzO ~ OH +OH H + H + M --. H 2 + M O + O + M --* O z + M O + H + M ---,O H + M H+OH+M ~H20+M H+O2+M ~ HO2+M H + H O 2 --* Hz + 0 2 H +HO2 ~ OH +OH H+HO 2 ~H20+O HO2 + O H ---*H 2 0 + O z H O 2 + O --* 0 2 + O H HOz + HO2 --* H2Oz + 0 2 H 2 0 2 + O H --, H 2 0 + H O 2 H 2 0 2 + H --* H 2 0 + O H H 2 0 2 + H --* H O 2 + H 2 H 2 O z + M --* O H + O H + M O + O H + M --, H O 2 + M H2+O z ~ OH+OH Os+M ~O2+O+M O a +O ~ O u+O u H + 0 3 --* O z + O H oa+o 3 ~ HO 2+0 2 C O + O + M --* C O 2 + M C O + 0 2 --* COz + O C O + O H --* CO2 + H CO + H O 2 --* C O 2 + O H C H 2 0 + M --* H C O + H + M CH20+OH ~ HCO+H20 C H 2 0 + H --* H C O + H z C H z O + O --* H C O + O H C H a O + H O z --* H C O + H z O 2 H C O + M --. H + C O + M H C O + O 2 ---,C O + H O u HCO+OH ~ CO+HzO H C O + H --, C O + H 2 H C O + O --, C O + O H log A 16.71 10.26 13.34 13.83 15.48 13.28 16.00 23.15 15.18 13.40 14.40 13.70 13.70 13.70 13.00 13.00 14.50 12.23 17.08 17.00 12.40 14.63 13.06 13.90 12.04 15.77 12.40 7.18 13.76 16.52 12.88 14.52 13.70 12.00 14.16 12.52 14.00 14.30 14.00 n -0.816 1.0 0.0 0.0 0.0 0.0 0.0 -2.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 1.3 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 Ea 16.51 8.90 5.15 18.36 0.00 -1.79 0.00 0.00 - 1.00 0.70 1.90 1.00 1.00 1.00 1.00 1.80 8.94 3.75 45.50 0.00 38.95 22.18 4.57 0.89 18.48 4.10 47.69 - 0.77 22.93 81.00 0.17 10.50 4.60 8.00 19.00 7.00 0.00 0.00 0.00 Reference 449 23 23 23 23 24 450 23 23 23 23 451 451 451 451 23 452 23 23 453 454 455 456 457 458 459 24 460 461 111 462 111 129 451 38 38 159 129 463

Modeling of hydrocarbon combustion APPENDIX I.

53

(continued)
Forward rate

Reaction 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. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. CH4+M ~ CH3+H+M C H 4 + O 2 --~ C H a + H O 2 CH 4 + H ~ CH 3 + H 2 C H 4 + O H -~ C H 3 + H 2 0 C H 4 + O -~ CH 3 + O H C H 4 + H O 2 -~ C H 3 + H 2 0 2 CH3+O 2 --~CH30+O C H a + C H 3 ---~C2H 6 CH3+CH 3 ~C2Hs+H CH3+CH 3 ~C2H4+H 2 CH 3 + O -o C H 2 0 + H CH 3 + O H ~ C H 2 0 + H 2 CH3+OH-o CH30+H CH3 + C H 2 0 ~ C H 4 + H C O CH 3 + H C O ~ CH 4 + C O C H 3 + H O 2 --, C H 3 0 + O H CH3+M--,CH2+H+M CHaO+M ~CH20+H+M C H 3 0 + O 2-o C H 2 0 + H O 2 C H 3 0 + H --* C H 2 0 + H 2 C2H 6 + 02 ~ C2H 5 + H O 2 C 2 H 6 + C H 3 ---~C2H5 + C H 4 C2H 6 + H ~ C2H 5 + H 2 c 2 a 6 + O ~ C2H 5 + OH C2H 6 + OH --* C2H s + H 2 0 C2Hs+M ~ C2H4+H+M C2H5+O 2 ~C2H4+HO 2 C2H 5 + O ~ C H 3 C H O + H C2H 5 + O ~ C H 2 0 + CH 3 C2H s + C 2 H s ~ C4H10 Call a ~C2Hs+CH 3 c 3 n 6 --* C 2 H 3 + C H 3 C2H 5 + C 2 H 3 ~ C4H a C2H 3 + C 2 H 3 --~ C4H 6 C2H 4 + M --, C2H 2 + H 2 + M C2H4+M ~ C2Ha+H+M C 2 H 4 + C 2 H 4 ~ C2H 3 + C 2 H 5 C2H 3 + M --* C2H 2 + H + M C 2 H 3 + O 2 --~ C 2 H 2 + H O 2 C 2 H 4 + H -~ C2Ha-I-H 2 C2H 4 + O H --, CzH 3 + H 2 0 C 2 H 4 + O --, CH3 + H C O C 2 H 4 + O --~ C H 2 0 + C H 2 C 2 H 4 + O H --, C H 3 + C H 2 0 C 2 H 3 + H -~ C 2 H 2 + H 2 C 2 H 3 + O -~ C H 2 C O + H C2H 3 + O H ~ C2H 2 + H 2 0 C2H3 + C 2 H 4 --~ C4H6 + H C2H2 + M --* C 2 H + H + M C 2 H 2 + C 2 H 2 --, C 4 H s + H C 2 H 2 + O 2 --r H C C O + O H C 2 H 2 + O 2 --, H C O + H C O CzH 2 + H ~ C2H + H 2 C2H 2 + O ~ C2H + O H C2H 2 + O ~ CH 2 + C O C2H2+O ~ HCCO+H C2H 2 + O H ~ C H 2 C O + H C2H 2 + O H ~ C2H + H 2 0 C2H 2 + O H ~ CH 3 + C O CzH 2 + O H ~ C2H2OH C2H2OH+H ~ CH2CO+H 2 C 2 H z O H + O ---*C H 2 C O + O H C2H2OH+OH ~ CH2CO+H20 C 2 H 2 O H + O 2 -o C H 2 C O + H O 2 C2H2OH+M ~ CH2CO+H+M C 2 H 2 + C 2 H --* C 4 H 2 + H C4H 2 + O H ~ C a l l 2 + H C O CH2CO+M ~CH2+CO+M

log A 17.30 13.90 4.35 3.54 7.07 13.30 13.68 13.00 14.90 16.00 14.11 12.60 16,30 10.00 11.48 13.30 16.29 13.70 12.00 13.30 13.00 --0.26 2.73 13.40 9.94 15.30 12.00 13.70 13.00 12.90 16.23 15.80 12.95 12.95 16.97 18.80 14.70 14.90 12.00 7.18 12.68 12.52 13.40 12.30 13.30 13.52 12.70 12.00 16.62 13.00 12.70 12.60 14.30 15.50 10.34 4.55 11.51 12.80 12.08 11.83 13.30 13.30 13.00 12.30 15.70 13.60 12.81 16.30

n 0.0 0.0 3.0 3.08 2.08 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.5 0.5 0.0 0.0 0.0 0.0 0.0 0.0 4.0 3.5 0.0 1.05 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 2.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 -0.6 1.0 2.7 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

Ea 88.00 56.00 8.75 2.00 7.63 18.00 29.00 0.00 26.52 32.00 2.00 0.00 27.41 6.00 0.00 0.00 91.60 21.00 6.00 0.00 51.00 8.28 5.20 6.36 1.81 30.00 5.00 0.00 0.00 0.00 84.84 85.80 0.00 0.00 77.20 108.72 64.70 31.50 10.00 6.00 1.23 1.13 5.00 0.96 2.50 0.00 0.00 7.30 107.00 45.00 23.50 28.00 19.00 15.00 2.58 1.39 0.20 7.00 0.50 0.23 4.00 4.00 2.00 10.00 28.00 0.00 1.00 60.00

Reference 146 464 149 150 151 464 125 465 136 136 114 466 467 468 468 226 155 125 127 469 470 149 149 27 165 161 45 136 168 136 195 195 195 195 471 45 182 182 45 45 45 27 45 45 170 472 46 182 186 185 46 184 188 188 473 473 194 188 474 46 46 46 46 46 46 477 190 46

54

C. K. WESTBROOKa n d F. L. DRYER

APPENDIX I.

(continued)
F o r w a r d rate

Reaction 108. 109. 110. 111. 112. 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. 169. 170. 171. 172. 173. 174. 175. CHzCO + OH ~ CHzO + HCO CH2CO + OH ~ HCCO + H20 CH2CO+O ~ HCCO+OH C H 2 C O + O ---, H C O + H C O CH2CO + H ~ HCCO + H z CH2CO +H ~ CH 3+CO H C C O + O 2 ---,C O + C O + O H HCCO+O ~ CO+CO+H HCCO + H ~ CH 2+ CO HCCO+OH ~ HCO+H+CO H C C O + C H z ~ C2H 3 + C O C H 2 + O 2 --, C O z + H 2 C H 2 + 0 2 --, C O 2 + H + H CH 2+ 0 2 ~ CO + H20 CH2+O a ~ CO+OH+H CH z + 0 2 ~ HCO + OH CH: + O ~ CH + OH CH2+O ~ CO+H+H C H z + O ---, C O + H 2 C H 2 + O H --* C H + H z O C H 2 + H ---, C H + H z C H z + C H 2 --, C 2 H 3 + H CH 2 + C H z ~ C2H 2 + H 2 CH 2+CzH 3 ~ CH a +CEH 2 C2H+O 2 ~ HCCO+O C2H+O 2 ~ HCO+CO C 2 H + O ---, C O + C H C 2 H + C 2 H a ---,C 2 H 2 + C2H2 CH + O 2 ~ HCO + O CH+O2 ~CO+OH C 4 H 3 q- M -~ C 4 H z + H + M C2H z + C 2 H ~ C4H: + H C4H 2 + M ~ C4H + H + M CHaOH + M ~ CH a + OH + M C H s O H + O 2 --* C H 2 O H + H O 2 CHaOH + H ~ CH 3+ H20 CHaOH+H ~ CHzOH+H z CHsOH + OH ~ CHzOH + H20 CHsOH + O ~ CHzOH + OH CHaOH + CH a ~ CH2OH + CH4 CHaOH + HO 2 ~ CH2OH + H20 2 CHzOH + M ~ CHzO + H + M CHzOH + O z ~ CHzO + HO z CHEOH + H ~ CHzO + H z C2HsOH + M ~ CH 3+ CH2OH + M CzHsOH + 0 2 ~ CHaCHOH + HO z C2HsOH+OH ~ CH3CHOH +H20 C 2 H s O H + H --* C H a C H O H + H 2 C2H5OH + O ~ CH3CHOH + OH C2HsOH+HO 2 ~CHaCHOH+H20 2 C 2 H s O H + C H a -~ C H a C H O H + C H 4 CHaCHOH+ M ~ CHaCHO+H +M CHaCHOH + 0 2 -, CHaCHO + HO 2 C 2 H s O H + H ~ C2H 5+ H 2 0 CH3CHO ~ CH 3+ HCO C H a C H O ~ C H A C O -k H CHaCHO + 0 2 ~ CHACO + HO 2 C H a C H O + H ~ CHACO + H 2 CHaCHO+OH ~ CHaCO+H20 CHaCHO+O ~ CHaCO+OH C H a C H O + C H 3 --~ C H a C O + C H 4 C H 3 C H O + H O 2 --* C H 3 C O + H 2 0 2 C H A C O ~ CH3 + C O C a l l a + 0 2 -~ iCaH v + H O 2 C a H s + 0 2 ~ n C a H ~+ H O 2 iCaH 7 + C 3 H a ~ n C a H 7 + C a l l 8 C 3 H 8 q- H ~ iCaH 7 q--H 2 CaHs+H ~ nCaHs+H 2

log A 13.45 13.00 13.00 13.00 13.00 13.04 11.80 12.08 12.70 12.30 13.48 12.21 12.57 11.00 11.30 14.00 11.30 12.70 12.70 11.40 11.40 12.70 13.50 13.48 12.52 13.00 13.70 13.48 13.00 11.13 16.00 13.60 17.54 18.48 10.60 12.72 13.48 12.60 12.23 11.26 12.80 1-3.40 12.00 12.48 18.48 10.60 13.48 12.64 12.83 12.80 12.60 13.70 13.00 12.72 15.85 14.70 13.30 13.60 13.00 12.70 12.23 12.23 13.48 13.60 13.60 10.50 6.94 7.75

n 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.68 0.0 0.0 0.67 0.67 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.67 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.5 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 2.0 2.0

Ea 0.00 0.00 0.00 2.41 0.00 3.40 2.00 0.00 0.00 0.00 0.00 1.00 1.50 0.00 0.00 3.70 25.00 0.00 0.00 25.70 25.70 0.00 0.00 0.00 0.00 7.00 0.00 0.00 0.00 25.70 60.00 0.00 80.00 80.00 50.91 5.34 7.00 2.00 2.29 9.80 19.36 29.00 6.00 0.00 75.47 50.00 5.96 4.57 1.51 15.00 9.69 21.85 5.56 5.00 81.76 87.86 42.20 4.20 0.00 1.79 8.43 10.70 17.24 47.50 47.50 12.90 5.00 7.70

Reference 194 46 46 136 46 136 46 199 46 46 46 46 46 46 46 182 475 46 46 189 475 46 476 46 46 188 188 46 46 189 185 477 185 43 222 43 43 43 478 479 222 43 222 469 480 481 480 482 483 481 484 480 480 481 226 485 226 136 136 136 484 226 486 487 487 227 195 195

Modeling of hydrocarbon combustion APPENDIX I.

55

(continued)
Forward rate

Reaction 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. 220. 221. 222. 223. 224. 225. 226. 227. 228. 229. 230. 231. 232. 233. 234. 235. 236. 237. 238. 239. 240. 241. 242. 243. C s H a + C H 3 ~ iC3H 7 + C H 4 C s H 8 + C H s --* nCsH 7 + C H 4 C3H 8 + C 2 H 3 ~ iC3H 7 + C 2 H a C a l l s + C 2 H 3 --->nCaH7 + C 2 H 4 C3H 8 + C 2 H 5 --->'iCaH7 + C 2 H 6 C3H 8 + C 2 H 5 ~ nC3H 7 + C 2 H 6 C3H 8 + C 3 H 5 ~ iC3H 7 + C 3 H 6 C 3 H s + C 3 H 5 --->n C 3 H v + C 3 H 6 C 3 H 8 + O -* i C a H T + O H C3Hs+O ~ nC3HT+OH C3Hs+OH--* iC3Hv+H20 C3Hs+OH ~ nC3H7+H20 C 3 H s + H O 2 --~ i C 3 H v + H 2 0 2 C 3 H s + H O 2 --~ n C a H 7 + H 2 0 2 nC3H ~ ~ C 2 H 4 + C H 3 nC3H v ~ C 3 H 6 + H iC3H 7 -* C 2 H 4 + C H 3 iC3H 7 -* C3H 6 + H n C 3 H 7 + O 2 -~ C 3 H 6 + H O 2 iCaH 7 + O 2 -'* C a H 6 + H O 2 C3H 6 ~ C 3 H s + H C 3 H 6 + O ---*C H 2 0 + C 2 H 4 C3H 6 + O ~ CH 3 + C H 3 C O C a l l 6 + O H ~ CH 3 + C H 3 C H O C3H 6 + O ---*C2H 5 + H C O C 3 H 6 + O H ~ C3H 5 + H 2 0 C3H 6 + OH ~ C2H 5 + C H 2 0 C3H6+H ~ C3Hs+H 2 C a l l 6 + C H 3 ~ C3H s + C H 4 C 3 H 6 + HO 2 --, C 3 H 6 0 + O H C a H 6 + C 2 H 5 -'~ C3Hs +C2H6 C3H 5 ~ C a H 4 + H C 3 H 5 + O 2 ---~C 3 H 4 + H O 2 C3H 5 + C H 3 --* CgH s C3H 5 + H ~ C 3 H 4 + H 2 C 3 H s + C H 3 "-* C 3 H 4 + C H 4 C 3 H 4 + O --r C H 2 0 + C 2 H 2 C 3 H 4 + O H --, C H 2 0 + C 2 H 3 C 3 H 4 + O --, H C O + C 2 H 3 C3H4+OH ~ HCO+C2H 4 C4Hlo ~ C 2 H s + C 2 H 5 C4H1o ~ n C 3 H v + C H 3 C4HIo+O 2 ~sC4Hg+HO 2 C4Hlo+H ~ pC4H9+H 2 C4Hlo + H --, sC4H 9 + H 2 C4H:o + O --, pC4H 9 + O H C4Hlo + O ~ sC4H 9 + O H C4Hlo + O H ~ pC4H 9 + H 2 0 C4H1o + O H ~ sC4H 9 + H 2 0 C a H l o + H O 2 --r p C 4 H 9 + H 2 0 2 C4H 1o + HO2 --~ sC4H9 + H2 02 C4HIo + C H 3 --* PC4H 9 + CH 4 C4HIo + C H 3 -* sCaH9 + CH4 C4HIo +C2 H3 -~ pC4H 9 + C 2 H 4 C4H1o + C 2 H 3 -~ sC4H 9 + C 2 H 4 C 4 H I o + C 2 H 5 -~ p C 4 H g + C 2 H 6 C 4 H l o + C 2 H 5 --, s C 4 H 9 + C z H 6 C4Hlo +C3H5 --' pC4H9 +C3H6 C4H~o + C3H 5 -~ sC4H 9 + C3H 6 nC4H 9 -~ C2H s + C 2 H 4 sC4H 9 -* C3H 6 + C H 3 nC4H 9 -~ , C 4 H s + H sC4H 9 ~ 2C4Hs + H n C 4 H g + O 2 --, IC4Hs + H O 2 sC4H 9 + 02 --'* IC4Ha + HO 2 $C4H 9 + 0 2 --~ 2C4H 8 + H O 2 1 C a H 8 + H --~ C 4 H 7 + H 2 2C4H 8 + H ~ C4H 7 + H 2

logA 15.04 15.04 11.30 11.30 11.30 11.30 11.90 11.90 6.70 6.70 8.68 8.76 12.70 12.70 13.98 14.10 10.30 13.80 12.00 12.00 13.00 13.77 12.70 12.85 12.55 12.60 12.90 12.70 10.95 7.30 11.00 13.60

n 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 2.0 2.0 1.4 1.4 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E~ 25.14 25.14 10.40 10.40 10.40 10.40 16.20 16.20 3.00 3.00 0.85 0.85 18.00 18.00 31.00 37.00 29.50 36.90 5.00 5.00 78.00 5.00 0.60 0.00 0.00 0.00 0.00 1.50 8.50 0.00 9.20 70.00 10.00 0.00 0,00 0.00 0.00 0.00 0.00 0.00 81.40 85.40 47.60 9.70 8.30 5.80 4.50 1.65 0.86 14.90 12.60 11.70 10.10 18.00 16.80 13.40 10.40 18.80 16.80 28.00 31.90 36.00 39.80 2.00 4.50 4.25 3.90 3.80

Reference 227 227 227 227 227 227 227 227 195 195 165 165 195 195 195 488 489 489 195 195 490 195 136 136 491 195 235 195 492 235 241 239 195 195 195 195 195 195 195 195 242 242 487 487 487 487 487 487 487 487 487 487 487 241 241 235 242 241 241 487 242 241 241 493 493 493 241 241

11.78
13.13 13.00 12.00 12.00 12.00 12.00 12.00 16.30 17.00 13.60 14.11 14.30 13.48 13.72 12.57 12.75 11.47 11.30 12.47 12.12 12.00 11.90 11.00 11.00 11.60 11.90 12.20 13.45 13.00 13.30 12.00 12.00 12.30 13.70 13.70

56

C. K. WESTBROOKand F. L. DRYER
APPENDIX I.

(continued)
Forward rate

Reaction 244. 245. 246. 247. 248. 249. 250. 251. 252. 253. 254. 255. 1C4H8 + O ~ C3H 6 + C H 2 0 2C4H8 + O ~ nC3H 7 + H C O a C 4 H a + O H ~ nCaH 7 + C H 2 0 2C4Ha + O H ~ C2H5 + C H 3 C H O 1C4H8 + C H 3 -~ C,,H7 + C H 4 2C4H8 + C H 3 ~ c 4 n 7 +CH,, 2C4Hs + O ~ C2H 5 + CH3CO C,,H 7 --* C4H 6 -I- H C4H 7 ~ C2H 4 -t- C2H 3 C4H v + C4H7 ~ 1C4H8 h- C4H 6 C4H 7 + G a i l y ~ 2C4H8 +C4.H 6 C4H 7 + CH 3 ---rC4H6 + CH 4

log A 12.70 12.78 13.26 13.41 11.00 11.00 12.70 14.08 11.00 12.20 12.20 12.90

n 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0

E~ 0.00 0.00 0.00 0.00 7.30 8.20 0.00 49.30 37.00 0.00 0.00 0.00

Reference 491 491 494 494 241 241 136 241 241 242 242 242

APPENDlX II. Inhibition mechanism, reaction rates in cm a mol sec kcal units, k = AT"exp(-EJRT) Forward rate Reaction 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 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. H+HBr ~ H2+Br H+Br 2 ~ HBr+Br B r + B r + M - , Br2 + M H+Br+M ~HBr+M H + H C I ~ H2+C1 H + C12 ~ HCI + C1 C I + C I + M ~ C12+M H+CI+M ~ HCI+M H+HI--* H2+I H+I2 ~HI+I I+I+M~I2+M H+I+M~HI+M H+HF~ H2+F H+Fz ~HF+F F+F+M~F2+M H + F + M --* H F + M H+CH3Br ~ HBr+CH 3 B r + C H 3 B r ---, B r 2 + C H 3 H+CH3CI ~ HCI+CH 3 CI+CH3C1 ~ C12 + C H 3 H+CH3I ~ HI+CH 3 I+CHaI ~I2+CH 3 H +C2H3Br ---, H B r + C 2 H 3 B r + C 2 H 3 B r ~ Br 2 + C 2 H 3 H+C2H3CI~ HCI+C2H 3 CI + C2H3C1 ~ C12 + C2H3 H+C2H3I ~ HI+C2H 3 I+C2H3I ~12+C2H3 H+C2HsBr ~ HBr+C2H 5 B r + C 2 H s B r --* Br 2 + C 2 H 5 H+C2HsCI~ HCI+C2H 5 C I + C 2 H s C 1 ~ C12 + C 2 H s H + C 2 H s I --* H I + C 2 H 5 I + C 2 H 5 I ~ 12+C2H5 B r + C H 4 ---, H B r + C H 3 Br+C2H 6 ~ HBr+C2H 5 Br+CH~OH ~ HBr+CH2OH CI+CH 4 ~ HCI+CH 3 CI+C2H ~ ~ HCI+C2H 5 I+CH4 ~ HI+CH3 I+C2H6 ~ HI+CzH 5 F + C H 4 --* H F + C H 3 F+C2H 6 ~ HF+C2H 5 F + C2H4 --* HF + C2H 3 H + C F 3 B r ~ CF3 + H B r log A 11.88 12.81 16.00 18.00 13.36 14.57 14.30 8.78 13.52 14.39 12.99 13.08 14.34 14.08 14.00 17.98 14.24 13.70 14.54 14.00 14.54 14.30 14.54 13.70 14.54 13.70 14.54 13.70 14.54 14.00 14.54 14.30 14.54 13.60 14.00 13.90 11.65 13.41 14.00 14.70 14.22 14.00 13.78 7.18 13.34 n 0.50 0.50 0.00 - 0.71 0.00 0.00 0.00 1.00 0.00 0.00 1.00 1.00 0.00 0.00 0.00 - 1.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 2.00 0.00 Ea 1.11 1.11 0.00 0.00 3.50 1.80 - 1.79 - 36.52 0.00 0.00 0.00 0.00 33.74 2.40 0.00 0.00 6.90 22.90 9.00 25.00 4.50 19.80 7.00 20.00 9.00 22.00 5.00 18.00 5.90 19.00 8.00 21.50 3.50 16.70 18.30 13.40 6.30 3.85 1.00 33.90 27.90 1.21 0.28 6.00 9.46

Modeling of hydrocarbon combustion APPENDIX II.

57

(continued)
Forward rate

Reaction 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. B r + C F 3 B r ~ Br 2 + C F 3 CH a + CF3Br ~ CH3Br + CF 3 CF3Br ~ C F 3 + Br CF 3 + H ~ CF 2 + H F CF3+O ~ CF20+F CF3 + O H ~ C F 2 0 + HF CF~ + C H 4 ~ C F 3 H + C H 3 C F 3 + C H 3 --, CH2CF2 + H F C F 3 O 2 --* C F 2 0 + F O CF3H+M ~ CF3+H+M CF3H ~ CF 2 + HF CF3H+H~CF 3+H 2 CF3H+O ~ CF3+OH CF3H+OH ~CF3+H20 CF3H + Br ~ C F 3 + HBr C H 2 C F 2 + O H --* C F 2 0 + C H 3 CH2CF 2 + O ~ C F 2 0 + CH 2 C H 2 C F 2 + O --, CF2 + CH2 O C F 2 0 + H ~ F C O + HF CF20+O ~ FCO+FO CF20 + OH ~ FCO + HOF CF 2 + H ~ H F + C F CF 2 + O --, F C O + F CF 2 + O H --, F C O + HF CF 2 + OH --* C F 2 0 + H CF2+CH 3 ~ CH2CF2+H FCO+M-* F+CO+M F C O + H --, H F + C O F C O + O ~ CO + F O F C O + O H -, CO + H O F F C O + O H ~ CO 2 + H F Br + HO 2 - , HBr + 0 2 HBr + OH ~ Br + H 2 0 HBr + O -, Br + O H B r + H 2 0 2 --* H B r + H O 2

log A 13.78 12.99 13.70 12.60 14.11 12.60 12.00 13.83 11.84 15.45 12.08 12.70 13.30 12.51 13.30 13.00 13.18 13.18 11.11 12.70 11.88 13.00 13.00 13.00 13.00 13.30 14.16 14.30 14.00 14.00 14.00 12.70 14.00 14.00 13.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00

E, 23.00 5.20 66.30 0.00 2.00 0.00 11.00 0.00 9.06 110.30 63.00 5.00 9.20 3.77 23.00 1.00 0.00 0.00 0.00 4.60 0.00 0.00 0.00 0.00 0.00 0.00 30.00 0.00 0.00 0.00 0.00 0.50 0.00 0.00 5.00

JPECS IO:I-D*

Anda mungkin juga menyukai