+ All Categories
Home > Documents > 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W....

2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W....

Date post: 29-Sep-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
218
2 Combustion Chemistry of Nitrogen Anthony M. Dean 1 Joseph W. Bozzelli 1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions of nitrogen-containing species that are important in high temperature gas-phase systems so as to provide the best set of rate coefficients presently available for use in combustion chemistry modeling. Since the review of N/H/O rate coefficients by Hanson and Salimian (1984) there has been a major review of nitrogen chemistry under combustion conditions by Miller and Bowman (1989). Several compilations of evaluated rate coefficients have also appeared. (Cohen and Westberg 1991; Tsang and Herron 1991; Baulch et al. 1992, 1994; Tsang 1992) We update these discussions where appropriate and then analyze a number of chemically activated reactions that are relevant to understanding nitrogen chemistry. After an introductory overview of nitrogen chemistry and an outline of our approach to analyzing chemically-activated reactions and hydrogen abstraction reactions, we examine closely those elementary reactions considered by Hanson and Salimian for which later measurements have made it possible to define rate coefficient expressions more accurately. We then focus upon a number of chemically-activated reactions for which recent research has improved our knowledge of reaction rates and product distributions. Among these reactions are those between hydrocarbon radicals and nitrogen species,such as the ones involved in “prompt NO” formation. We demonstrate that for many of these reactions one can understand otherwise apparently conflicting information by recognizing that 1 Corporate Research Laboratory, Exxon Research and Engineering Company, P. O. Box 998, Annandale, NJ 08801 2 Department of Chemical Engineering and Chemistry, New Jersey Institute of Tech- nology, Newark, NJ 07102
Transcript
Page 1: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2

Combustion Chemistry of Nitrogen

Anthony M. Dean 1

Joseph W. Bozzelli 1,2

2.1 INTRODUCTION

The purpose of this chapter is to examine reactions of nitrogen-containing speciesthat are important in high temperature gas-phase systems so as to provide thebest set of rate coefficients presently available for use in combustion chemistrymodeling. Since the review of N/H/O rate coefficients by Hanson and Salimian(1984) there has been a major review of nitrogen chemistry under combustionconditions by Miller and Bowman (1989). Several compilations of evaluated ratecoefficients have also appeared. (Cohen and Westberg 1991; Tsang and Herron1991; Baulch et al. 1992, 1994; Tsang 1992) We update these discussions whereappropriate and then analyze a number of chemically activated reactions that arerelevant to understanding nitrogen chemistry.

After an introductory overview of nitrogen chemistry and an outline of ourapproach to analyzing chemically-activated reactions and hydrogen abstractionreactions, we examine closely those elementary reactions considered by Hansonand Salimian for which later measurements have made it possible to definerate coefficient expressions more accurately. We then focus upon a numberof chemically-activated reactions for which recent research has improved ourknowledge of reaction rates and product distributions. Among these reactions arethose between hydrocarbon radicals and nitrogen species,such as the ones involvedin “prompt NO” formation. We demonstrate that for many of these reactions onecan understand otherwise apparently conflicting information by recognizing that

1 Corporate Research Laboratory, Exxon Research and Engineering Company, P. O.Box 998, Annandale, NJ 08801

2 Department of Chemical Engineering and Chemistry, New Jersey Institute of Tech-nology, Newark, NJ 07102

Page 2: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

126 Anthony M. Dean, Joseph W. Bozzelli

chemically-activated adducts formed by making new chemical bonds in addition,radical recombination and insertion reactions often have multiple high temperaturereaction pathways. We then consider reactions for which the data base is sparsebut which must nonetheless be considered in modeling nitrogen chemistry; ourapproach is to assign rate coefficients that are consistent with the scant data andthermochemical and reaction-kinetics principles. Brief discussion is provided ofinteractions between nitrogen and chlorine chemistry that may arise in toxic wasteincineration. Finally we present a few modeling examples to illustrate predictionswith our rate coefficient set and point out areas that require additional study.

2.2 OVERVIEW OF NITROGEN CHEMISTRY

Interest in the combustion chemistry of nitrogen compounds derives from the roleof nitrogen oxides, collectively known as NOx, in our environment. Increasinglystringent environmental regulations require increasingly sophisticated approachesfor the control of NOx formation during combustion. (Bowman 1992) Detailedchemical modeling plays a major role in the implementation of such approaches.We set the stage for consideration of specific elementary reactions by discussingthe mechanisms for production and removal of NO, the primary NOx speciesproduced in combustion.

NO is known to be formed in a variety of ways: (1) “Thermal NO” (Zeldovich1946) is primarily a consequence of high flame temperatures; (2) “Prompt NO”(Fenimore 1976) is generated in fuel-rich parts of flames; (3) the “N2O mecha-nism” (Wolfrum 1972; Malte and Pratt 1974) can be important in high pressureflames; (4) “Fuel nitrogen” NO (Fenimore 1976) results from converting nitrogen-containing compounds in the fuel into NO; and the (5) NNH mechanism (Bozzelliand Dean 1995) is active in flame fronts where high atom concentrations appear.All but (4) convert atmospheric nitrogen into NO and relate to all fuels, whilepathway (4) pertains to flames of petroleum, coal, and biomass.

2.2.1 Thermal, or Zeldovich NO

The principal elementary reactions1 here are:

O+N2 ⇀↽ NO+N (1)

N+O2 ⇀↽ NO+O (26a)

N+OH ⇀↽ NO+H . (26c)

The first two2 reactions compose a chain sequence, so a small amount of atomicoxygen can produce large amounts of NO. The first one is rate-limiting because

1 Reaction numbers in this section correspond to those introduced serially in latersections of this chapter.2 Sometimes the term “Zeldovich mechanism” is used to refer only to reactions (1)and (26a); inclusion of reaction (26c) is then the “extended Zeldovich mechanism”.

Page 3: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 127

of its high activation energy, about 320 kJ/mol. The Zeldovich mechanism is verysensitive to temperature not only because of the high activation energy of reaction(1) but also because the concentration of oxygen atoms in flames increases rapidlywith increasing temperature. Since the rate coefficient for reaction (1) is well-established, computation of the Zeldovich NO production rate is straightforwardif the oxygen atom concentration is known. Away from the flame front regionone can indeed compute Zeldovich NO production with little concern for detailsof the chemical kinetics by assuming equilibrium O-atom concentrations. Thesituation is more complicated in the flame-front region, where super-equilibriumconcentrations of radicals are present. This is not because of any particular flame-front reactions but because flame-front chemistry is driven by fast bimolecularreactions while post-flame chemistry includes slow termolecular recombinationreactions. One must describe the kinetics of flame front chemistry properly inorder to predict Zeldovich NO production accurately.

2.2.2 Prompt, or Fenimore NO

This is aptly called Prompt NO, because it is manifested by prompt, i.e., rapid,production of NO in a flame front. The experimental observation is that anextrapolation of NO concentration to zero time in a plot of NO against residencetime, or distance from the flame front, gives a positive intercept. (It is invariablyfollowed by a slower rate of NO production attributable to Thermal NO.) In aflame front the concentration of hydrocarbon radicals is large, and NO productionis initiated by their reaction with N2 to break the N–N bond. As discussed in detaillater, the most important such reaction is thought to be

CH+ N2 ⇀↽ HCN+N . (24c)

The product N atom can form an NO molecule via reactions (26a) or (26c), whilethe HCN can lead to a second NO molecule through a series of reactions discussedby Miller et al. (1984). A description of NO production in this case requires, inaddition to accurate values of rate coefficients for reactions such as (24c), a reliablemechanism for fuel combustion, since one needs to know the concentrations of thehydrocarbonradicals, including CH. Although there has been appreciable progressin recent years in developing such mechanisms, there remains much doubt abouthydrocarbon radical profiles. This is still an active research area.

2.2.3 The N2O pathway

Reaction (1) above is one pathway for reactions of oxygen atoms with molecularnitrogen. Another is intermediate formation of N2O by a “recombination reaction”

O+N2+M ⇀↽ N2O+M , (−3)

Page 4: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

128 Anthony M. Dean, Joseph W. Bozzelli

where the “collision partner” M collectively represents all molecules present.1

(Cf. Section 2.3)The N2O formed in reaction (−3) can then react to form NO by

O+N2O ⇀↽ 2 NO (4b)

and

H+ N2O ⇀↽ NO+NH . (11c)

The termolecular character of reaction (–3) implies that it becomes more impor-tant at higher pressures. NO predictions for this pathway are somewhat morecomplicated than for Thermal NO, but not as difficult as for Prompt NO. A majorissue is that there are other pathways besides (4b) and (11c) for the reactions of Oand H with N2O, and one must properly account for the branching ratios in thesereactions. Moreover, as with Thermal NO, super-equilibrium concentrations ofoxygen atoms in the flame front region must be taken into account.

2.2.4 Fuel nitrogen

It is frequently assumed that most fuel nitrogen is quickly converted to HCN inflames. (Morley 1980) The subsequent kinetics of HCN are generally similarto those that apply to HCN in forming Prompt NO. We note, however, there isevidence suggesting that the assumption of rapid conversion of fuel nitrogen toHCN is not always valid. (Mackie et al. 1990) It may eventually turn out thatcorrect accounting for NO formation from fuel nitrogen also requires includingdetails of other N-containing species besides HCN.

2.2.5 The NNH mechanism

Under combustion conditions where the concentrations of atoms are high, i.e., inflame fronts, the reaction

O+NNH −→ NO+NH (−11g)

contributes significantly to NO production. This mechanism requires participationof H atoms to form NNH from N2 and O atoms to react with NNH.

1 Reaction numbers preceded by a minus sign refer to the “reverse reaction”, i.e.,the reaction with that number understood to be occurring in the opposite directionthan indicated by the arrow of the “forward reaction” denoted by the numberwithout the minus sign.

Page 5: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 129

2.2.6 Effects of temperature and pressure

The relative importance of the first three channels for NO production in methane–air flames has been addressed in a series of model calculations reported by Drakeet al. (1990; 1991a; 1991b), and these can be also taken as representative of thesituation for other fuels. For a 0.1 atm stoichiometric laminar premixed flame,their calculations indicated that the Fenimore mechanism was dominant. At theserelatively low pressures, much of the Zeldovich NO production could be tracedto super-equilibrium oxygen atom concentrations. In a 20 atm stoichiometricmethane–air flame, the major contributor is equilibrium Zeldovich NO, but thehigher pressure has increased the contribution of the N2O pathway to be compa-rable to that of Fenimore NO. In laminar diffusion flames at 1 atm, these workersconcluded that Fenimore NO was the dominant pathway, contributing more than2/3 of the total NO, with the Zeldovich and N2O paths being comparable minorcomponents.

These results illustrate the complexity of NO production chemistry. Given thevery high sensitivity of the Zeldovich route to temperature, use of different fuels(with different adiabatic flame temperatures) or considering a flame configurationwith a different amount of heat loss dramatically affects the predicted amount ofZeldovich NO. Similarly, different fuels change the amount of CH produced inthe flame front, thus affecting the amount of Fenimore NO formed.

2.2.7 NO reduction

There are several reactions by which NO can be converted to N2 or to species thatmay subsequently be converted to N2 during the combustion or post-combustionprocess. As discussed in detail later, these include

NO+ CH3 ⇀↽ H2CN+OH (19d)

+NH2 ⇀↽ N2+H2O (18a)

+NH2 ⇀↽ NNH+OH (18b)

+NH ⇀↽ N2+OH (11e)

+NH ⇀↽ N2O+H . (−11c)

The first can be important in “reburning”, where additional fuel is added afterthe primary combustion stage (Bowman 1992). The second and third playcentral roles in post-combustion NO removal—Selective Non-Catalytic Reduction(SNCR)—processes such as Thermal DeNOx, wherein ammonia is added in thepost-combustion zone to reduce NOx emission levels in the exhaust gas (Lyon1975). The last two illustrate the complexity of the chemistry again: N2 may bea direct product (11e), or the initial product N2O can subsequently react either toreform NO or to be converted to N2.

Page 6: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

130 Anthony M. Dean, Joseph W. Bozzelli

2.3 UNIMOLECULAR AND CHEMICALLYACTIVATED BIMOLECULAR REACTIONS

2.3.1 Unimolecular reactions

Combustion scientists specializing in chemical kinetics have known for some timethat it is necessary to consider the pressure dependence of the rate coefficients ofdissociation and recombination reactions. The pressure effect can be qualitativelyunderstood just by recognizing that these reactions are really not one-step pro-cesses. Considering the rate coefficient krec for a recombination reaction, forexample,

A+ B −→ AB ,

one has to take into account that formation of the new chemical bond initiallyproduces an energy-rich adduct that can either redissociate to A and B or bestabilized in an energy-transferring collision with a “bath gas” molecule M toform products:

A+ B ⇀↽ AB* (a)

AB*+M −→ AB+M . (b)

Assuming that the overall process can be described by assigning rate coefficientska and kb to these two processes and using the steady-state approximation—thatthe time derivative of the concentration of the energy-rich adduct AB* is smallcompared to the rate of product formation—one readily derives

d[AB]

dt= krec[A][B] = kakb[M]

k−a + kb[M][A][B]

from which it is seen that krec is actually a function of [M]

krec = kakb[M]

k−a + kb[M].

Inspection of these equations shows that the reaction rate is predicted to scalelinearly with pressure at sufficiently low pressure, i.e., sufficiently low [M]. Whenk−a >> kb[M] the “low pressure rate coefficient” becomes

krec(P→0) = ka

k−akb[M] = Kakb[M] = k◦ [M] .

The rate coefficient is independent of pressure at high pressures, where k−a <<

kb[M], for which we derive the “high pressure rate coefficient” for recombinationto be

krec(P→∞) = kakb[M]

kb[M]= ka = k∞ .

Page 7: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 131

The rate coefficient of the A + B−→AB recombination reaction can be expressedas a function of the high and low pressure limiting values k∞ and k◦[M] in theform

krec = k∞(

Pr

1+ Pr

)where the dimensionless reduced pressure Pr is given by

Pr = k◦[M]/k∞ .

Such scaling with [M], i.e., with pressure, has been observed experimentally formany reactions of this kind.

For the reverse reaction of dissociation

AB −→ A+ B ,

one can follow an analogous line of reasoning to find the pressure dependence ofthe rate coefficient for dissociation kdiss in the first-order decay rate law

d[AB]

dt= − kdiss[AB] .

Because the recombination and dissociation reactions represent the forward andreverse directions of the same chemical process, their rate coefficients are relatedto one another by the equilibrium constant Keq for the reaction A + B ⇀↽ AB, i.e.,

krec

kdiss= Keq .

This equation is a fundamental relationship known as the principle of microscopicreversibility. It can readily be derived for the above two-step mechanism bysolving for the pressure dependence of kdiss using the same line of reasoningthat led us to the pressure dependence of krec. Because Keq is a thermochemicalquantity dependent on temperature only, the “falloff” pressure range over whichkrec and kdiss decrease from their “high pressure limit” to “low pressure limit”values is identical for both and similarly depends only on temperature. In termsof the kinetics of the two-step mechanism one can understand the sense of thetemperature dependence by observing that the dissociation rate coefficient k−a forthe energized complex AB* has to increase with temperature. The stabilizationrate coefficient kb has a much weaker temperature dependence, and so the valueof [M] (and thus the pressure) needed to maintain high-pressure limiting behaviormust be higher at higher temperature. This proves to be a rather large effect over awide temperature range, and explicit account must be taken of it when constructingcombustion models. (Cf. Warnatz (1983).

Neither the pressure falloff nor its temperature dependence, however, areaccurately accounted for just by writing down the pair of reactions (a) and (b)as we have done above. This is because considering only a single energized

Page 8: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

132 Anthony M. Dean, Joseph W. Bozzelli

intermediate AB* is an oversimplification of the real process. In order to accountfor the experimentally observed pressure falloff and its temperature dependenceone must recognize that a wide range of energization, and thus a wide range of kaand k−a values, really characterizes the observed reaction. The consequences ofthis situation have been thoroughly explored theoretically, primarily by means ofwhat is termed “RRKM Theory”, and it is readily possible to account for observedfalloff behavior in terms of plausible molecular models for recombination anddissociation reactions. (Forst 1972; Robinson and Holbrook 1972; Gilbert andSmith 1990)

Several better approximations have been introduced to allow more realisticdescriptions of falloff behavior to be incorporated into kinetic models. One intro-duced by Troe and co-workers (Gilbert et al. 1983) involves adding a “broadeningfactor” F to the pressure-dependent krec formula

krec = k∞(

Pr

1+ Pr

)F

where F is a smooth but complicated function of temperature given by the set ofequations

log F =[

1+(

log Pr + c

n − d(log Pr + c

)2]−1

c = −0.4− 0.67 log Fc

n = 0.75− 1.27 log Fc

d = 0.14

Fc = (1− a) exp(−T/T ∗∗∗)+ a exp(−T/T ∗)+ exp(−T ∗∗/T ) .

The idea is to fit Fc parameters a, T ∗∗∗, T ∗, and T ∗∗ to computed or experimentalfalloff data once and then use these equations to compute krec(P, T ) for dynamicmodeling. An alternative representation of F(T ) was introduced by Stewart et al.(1989)

F = [a exp(−b/T )+ exp(−T/c)]X

X = 1/(

1+ log2 Pr

)with the fitting parameters a, b and c. Either of these approximations can be usedwith the Chemkin modeling package (Kee et al. 1989).

An alternative representation of falloff behavior is to compute theoretical kdiss

or krec values for one fixed pressure at a number of temperatures, then fit these to anArrhenius-like temperature function. (E.g. Tsang 1991) The pressure-dependentrate coefficients discussed in later sections are described in this way for pressuresof 1/10, 1 and 10 atm. More compact and more accurate interpolation formulashave been discussed by various authors (Gardiner 1988; Oref 1989; Pawlowskaet al. 1993; Wang and Frenklach 1993; Kazakov et al. 1994; Prezhdo 1995;Venkatesh et al. 1997).

Page 9: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 133

2.3.2 Pressure-dependent bimolecular reactions

In recent years it has been recognized that similar falloff behavior is to be expectedfor any reaction that proceeds via a “chemically-activated” energized complex,including reactions that involve only transient adduct formation. A variety ofreaction types, including additions, recombinations and insertions, as well asreactions whose overall stoichiometry might suggest a simple bimolecular reactionmodel, have been analyzed in this way using RRKM theory. (E.g., by Bermanand Lin 1983; Larson et al. 1984; Golden and Larson 1985; Wagner et al. 1990;and Just 1994).

A special consideration that arises in analyzing reactions that proceed viaenergized complexes is the possibility of multiple dissociation channels. Radicaladdition, recombination and insertion reactions can manifest complex temperatureand pressure behavior as stabilization channels compete with multiple decompo-sition channels. Significant errors may result if experimental measurements ofrate coefficients for reactions of this type are extrapolated to different regions oftemperature and pressure without accounting for the competition between uni-molecular reactions of the energized complexes and their bimolecular collisionalstabilization, as described in the following section.

2.3.3 Quantum Rice-Ramsperger-Kassel Theory

A theoretical approach to chemical activation systems, intended for initial evalu-ations and modeling applications, was developed by Dean (1985). In its originalformulation it was based on a quantum version of Rice-Ramsperger-Kassel theory(QRRK) and used a geometric mean frequency to represent the internal dynamicsof the reacting molecule. Robinson and Holbrook (1972) suggest that the quantumversion is an improvement over the more commonly used classical version ofKassel theory because the quantum version considers the full set of internalvibrational modes of the reacting molecule, while in using the classical RRKversion one typically reduces their number by a factor of 1/2 to 1/3 in order tofit experimental falloff data. In essence, QRRK theory uses statistical mechanicsto calculate the probability that sufficient energy will be localized in a givenvibrational mode for reaction to occur. It requires Arrhenius A-factors, either fromthe literature or from “Transition State Theory” (hereafter TST; cf. Benson 1978and Zellner 1984) estimates. In combination with a modified “strong collision”model due to Troe (1979) to describe collisional stabilization it was applied to theanalysis of a variety of chemical activation reactions (Westmoreland et al. 1986;Dean and Westmoreland 1987; Bozzelli and Dean 1989, 1990, 1993) where therewere sufficient experimental results in the literature to evaluate its predictions.Its success in these applications shows that QRRK theory provides a simpleframework with which the effects of temperature and pressure can be readilyunderstood and evaluated.

QRRK theory is a complementary technique to RRKM theory. In applyingRRKM theory one calculates an Arrhenius A-factor for k∞ from computed

Page 10: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

134 Anthony M. Dean, Joseph W. Bozzelli

properties of the transition state and the reactant molecule, i.e., their vibrationalfrequencies and their internal rotor barriers, moments of inertia and symmetries, ifapplicable, and their moments of inertia for overall rotation. The value computedfor this A-factor is generally determined by the values used for the low frequencyvibrations of the transition state, which for complex molecules are sufficientlyuncertain that adjustments are often made to obtain agreement with experiment.Thus, although RRKM theory is in principle a more comprehensive descriptionof unimolecular reaction, uncertainties in the necessary input parameters severelyrestrict its predictive value even when significant computational effort is investedin computing these parameters. While use of RRKM theory may be appropriatein some contexts, it does not offer substantial advantages for the combustion mod-eler. Most of the unimolecular and chemical activation reactions of combustionchemistry are of such complexity that computing the molecular properties of thetransition state at the required level of accuracy is infeasible.

2.3.4 Implementation of QRRK theory

The information required for a QRRK analysis is available from the literature,fromthermodynamic parameter estimation codes (Ritter and Bozzelli 1991; Bozzelli etal. 1997), TST estimation codes (Bozzelli and Ritter 1993) and transport propertyestimation methods (Reid et al. 1987). The required input data include the Arrhe-nius A-factors for the high pressure limit rate coefficients k∞ for dissociation of theadduct. It can be derived from measured values or estimated from rate coefficientsof isoelectronic or structurally similar reactions with known kinetic parameters.Our preference is to use the equilibrium constant for the recombination reactionin conjunction with the recombination reaction rate coefficient and the equationkdiss = krec/Keq . More data is usually available for the reverse (bimolecular)reaction; alternatively, in the case of a theoretical estimation, it is more reliable toestimate recombination than dissociation rate coefficients. This approach avoidsthe use of adjustable parameters for the unimolecular dissociation process andguarantees thermochemical consistency. Typically we fit equilibrium constantvalues over the desired temperature range with appropriate algebraic expressionsthat account with good accuracy for changes in heat capacity. Equilibriumconstants needed for all QRRK calculations reported here can be computedusing the thermochemical data base summarized in Tables 1 and 2, which wereassembled from a variety of sources including the Sandia data base (Kee and Miller1987) and computations made with the group-additivity code THERM (Ritter andBozzelli 1991). The reader is referred to a review (Dean et al. 1991) for furtherdiscussion of the QRRK method and guidance in selection of input parameters.

A shortcoming of the original QRRK approach was the use of a single frequencyto represent the 3n − 6 vibration frequencies of a molecule. Weston (1986)compared two QRRK approaches (geometric and arithmetic mean), classical RRKtheory, and a quantized version of classical RRK by Schranz et al. (1982) withRRKM theory for a number of unimolecular dissociation reactions. He found that

Page 11: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 135

TABLE 2.1. Thermochemical data for elements and binary species.

Species1 f H ◦298

kJ/molS◦298

J mol−1 K−1C◦P(T )/J mol−1 K−1 at T/K =

300 400 600 800 1000 1500

Ar 0.0 154.7 20.8 20.8 20.8 20.8 20.8 20.8O 249.2 161.0 21.9 21.5 21.1 21.0 20.9 20.8O2 0.0 205.0 29.3 30.3 32.1 33.6 34.7 36.5H 218.0 114.6 20.8 20.8 20.8 20.8 20.8 20.8H2 0.0 130.6 28.9 29.0 29.3 29.8 30.4 32.1C 716.8 158.0 20.8 20.8 20.8 20.8 20.8 20.8N 472.7 153.2 20.8 20.8 20.8 20.8 20.8 20.8N2 0.0 191.5 28.7 29.3 30.4 31.5 32.6 34.6

OH 39.7 183.6 29.1 29.2 29.6 30.2 30.8 32.6HOO 15.9 229.0 35.3 37.6 41.6 44.8 47.6 51.8H2O –241.8 188.7 33.9 34.5 36.3 38.7 41.2 46.8HOOH –136.1 232.9 43.6 48.0 54.8 59.9 63.6 70.5

CO –110.5 197.6 28.7 29.4 30.8 31.9 33.0 35.0CO2 –393.5 213.7 37.2 41.2 47.3 51.5 54.3 58.3

CN 435.1 202.5 29.2 29.4 30.7 32.1 33.5 35.6NCCN 309.1 241.6 57.1 61.5 68.3 73.1 76.3 81.1

NO 90.3 210.7 29.3 30.1 31.5 32.8 33.8 35.6N2O 82.1 219.8 38.7 42.6 48.4 52.3 54.8 58.3NO2 33.1 240.0 37.3 40.7 45.7 49.3 51.6 54.7NO3 71.1 252.6 47.6 55.9 67.1 73.2 76.4 79.9

CH 594.1 182.9 29.2 29.2 29.8 31.0 32.5 36.63CH2 386.4 193.8 34.6 36.1 39.2 42.5 45.5 51.11CH2 424.5 184.7 34.8 36.5 40.1 43.4 46.4 51.9CH3 145.7 194.1 38.8 42.2 48.5 54.2 59.3 68.2CH4 –74.9 186.2 35.9 41.2 52.0 62.2 71.1 86.1C2H 562.6 215.5 42.0 43.6 45.9 48.0 50.0 54.1C2H2 225.4 200.9 44.2 50.1 57.5 62.5 66.6 74.1C2H3 299.7 236.9 41.9 50.1 63.1 72.5 79.7 91.4C2H4 52.4 219.5 42.7 53.2 71.1 84.3 94.3 110.4C2H5 119.7 250.5 48.6 59.9 80.2 95.7 107.9 127.6C2H6 –85.3 230.5 51.8 65.6 90.3 108.9 123.6 147.1HC≡CC

.H2 341.3 249.2 57.6 67.5 82.8 93.7 101.8 114.9

C4H2 439.6 250.1 73.8 84.1 96.3 104.3 110.4 120.7HC≡CC

.=CH2 491.9 282.3 73.6 87.5 107.5 119.4 128.3 142.5

3NH 358.8 181.1 29.2 29.1 29.5 30.3 31.2 33.61NH 505.8 174.4 29.2 29.3 30.4 31.8 33.1 35.3NH2 190.3 194.6 34.0 34.9 37.2 40.0 42.7 48.2NH3 –45.9 192.6 35.7 38.8 45.1 51.1 56.4 66.0NNH 249.2 224.4 34.8 36.9 41.3 45.4 48.2 52.0N2H2 209.2 218.5 36.6 41.2 49.6 56.5 62.1 70.5N2H3 201.2 228.6 43.8 51.5 63.6 72.5 79.2 90.3N2H4 95.4 238.6 51.1 61.7 77.5 88.4 96.2 110.1

See text for explanations of data sources. All values refer to the ideal gas state. Adjacentchemical symbols for the same element indicate atoms bonded to one another in the mannerof peroxides. Unpaired electrons are denoted by dot symbols only where needed to resolveambiguity.

Page 12: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

136 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.2. Thermochemical data for ternary and quaternary species.

Species1 f H ◦298

kJ/molS◦298

J mol−1 K−1C◦P(T )/J mol−1 K−1 at T/K =

300 400 600 800 1000 1500

HCO 43.5 224.5 34.6 36.7 40.9 44.9 48.2 52.5CH2O –108.8 218.9 35.4 39.6 48.1 55.8 62.2 70.91HCOH 98.3 239.3 45.3 49.8 56.9 62.0 65.9 71.8H2COH –15.1 247.0 49.2 55.5 64.7 72.0 77.9 87.2HOCO –207.9 251.6 44.6 50.6 60.2 67.0 71.6 77.0HCO2. –160.0 242.4 41.2 48.4 59.0 67.0 72.0 77.4CH3OH –200.8 239.7 43.9 51.6 67.0 79.7 89.5 105.1CH3O 16.6 233.6 39.8 46.2 59.1 69.7 77.7 89.7HC≡COH 85.5 245.6 55.3 61.9 74.2 80.2 85.0 96.0HC.

=C=O 173.1 253.1 51.3 56.9 65.5 69.1 72.7 77.2H2C=C=O –49.1 241.9 53.1 61.3 74.5 81.6 87.8 96.4

HNO 106.3 220.6 34.6 36.9 41.2 45.0 48.0 52.31HON 255.2 214.4 35.0 37.2 41.3 44.8 47.7 52.23HON 209.2 226.0 35.0 37.2 41.3 44.8 47.7 52.2HNNO 231.8 253.3 44.8 50.7 59.7 65.8 69.9 75.9NNOH 243.1 257.1 43.7 47.8 54.8 59.9 63.4 69.8HONO –77.8 254.0 45.1 51.1 59.9 65.4 68.9 74.91HNOO 235.5 249.8 42.9 49.8 60.3 67.3 71.8 77.7HNO2 –59.2 237.4 37.7 43.5 53.2 60.5 66.0 73.5HNOH 90.4 241.8 43.1 47.0 54.1 59.8 63.6 70.5NH2O 66.5 233.0 38.9 43.5 51.1 57.2 62.1 70.2NH2NO 74.9 252.3 50.9 59.9 74.6 85.3 92.5 100.5c-HNNOH 89.9 259.6 49.4 59.2 74.7 85.4 92.4 100.4t -HNNOH 80.7 258.2 49.1 61.1 77.8 87.7 93.3 100.5

HCN 135.2 201.7 36.0 39.0 43.7 47.5 50.3 55.1HNC 189.1 205.9 36.0 39.1 43.9 47.4 50.1 55.2HCNN 456.1 248.3 45.1 51.0 59.8 65.6 69.3 74.1HCN2 448.0 240.7 46.0 50.3 57.8 64.0 67.3 73.6H2CN 247.3 224.2 38.3 43.2 52.2 59.6 64.5 71.7HCNH 276.6 234.0 38.2 43.2 52.2 59.5 64.1 72.3CH2NN 286.4 242.5 49.1 58.0 71.3 80.2 86.5 96.03CH3N 329.3 231.3 41.9 49.2 62.4 72.6 80.4 92.31CH3N 470.6 225.5 41.9 49.2 62.4 72.6 80.4 92.3H2C=NNH2 190.4 264.0 59.8 72.0 82.8 105.4 121.8 135.3H2C=NH 91.4 234.1 38.2 44.7 57.9 69.5 77.7 89.5CH3N=NH 179.1 256.9 57.3 67.8 78.2 102.1 113.4 131.0CH3N=N 215.5 262.3 48.7 58.4 74.5 87.1 96.8 112.2CH3NH 181.0 249.7 46.6 55.0 70.2 81.2 89.7 105.3CH2NH2 151.7 242.4 54.5 62.5 76.4 86.8 94.7 108.8CH3NH2 –23.0 242.6 50.2 60.2 78.9 93.9 105.7 124.9

NCO 132.2 232.1 40.2 43.8 49.5 53.4 55.9 58.9

HCNO 176.3 242.7 46.6 53.0 61.3 66.9 71.1 77.7HOCN –9.5 242.1 46.1 50.5 57.3 62.4 66.3 72.7HNCO –101.7 240.7 46.6 51.2 59.0 64.5 68.1 73.7CH2NO 173.4 255.3 49.1 58.0 71.3 80.2 86.5 96.0.N=CHOH 58.0 263.2 52.3 59.6 71.8 80.2 85.6 93.7CH3NO 77.4 255.2 52.7 62.7 79.1 89.7 98.7 111.3NH2CH2O 9.2 274.5 54.8 66.1 76.1 99.6 110.0 127.7

See footnote to Table 1.

Page 13: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 137

use of a geometric mean QRRK approach tends to overestimate k(E), the energy-dependent dissociation rate coefficient, for typical molecules at energies within40 kJ/mol of the critical energy for dissociation. The overestimate is greaterwhen an arithmetic mean frequency is used. Comparisons of k/k∞ betweenRRKM and QRRK theories in Weston’s study showed that while the results werereaction specific, overall the agreement for the 8 molecular systems analyzed wasgood. Significant deviations of QRRK from RRKM theory did appear at lowertemperatures for several molecules. This is precisely the region where one wouldexpect to find such differences, since the effectively coarser energy graining with asingle frequency will have its most pronounced effect when it has to approximatethe relatively compressed low temperature thermal distribution function.

We addressed this issue by developing a three-frequency QRRK model. Threefrequencies and associated degeneracies are computed from fits to heat capacityestimates. (Bozzelli et al. 1997) Such 6-parameter sets, together with theappropriate statistical thermodynamics formulas, were shown by Ritter (1989) toreproduce molecular heat capacities and vibrational state densities accurately. Thisapproach avoids specification of the complete frequency distribution of the adduct,which is unavailable for most of the molecules of interest here. Comparisons ofk/k∞ for the 8 molecular dissociations considered by Weston show that thisthree-frequency QRRK treatment yields significant improvement over the singlefrequency version. We believe that it is suitable for analyzing both unimolecularfalloff and bimolecular chemical activation reactions.

In this chapter we report results of using QRRK theory to analyze a varietyof reactions of nitrogen-containing species. The QRRK calculations we reportwere done with a computational procedure that is a significant extension of theone used in our earlier work. (Dean et al. 1991) In addition to incorporatingthe three-frequency model, it was generalized to allow multiple intermediatespecies (potential energy wells) to be connected in many ways. The collisionalstabilization model has also been improved in two ways. First, the collisionalefficiency β is now calculated from the average energy α transferred in a downcollision following the proposal of Gilbert et al. (1983)

β =

( α

α + FE kBT

)2

∫ E0

0f (E)

[1− FE kBT

α + FE kBTexp

( E − E0

FE kBT

)]dE

.

As discussed by Gilbert et al., this approximation provides a more realisticdescription of weak collision effects at high temperatures than the suggestionof Troe (1979); in particular, β calculated in this manner does not decreaseto unrealistically small values at high temperatures. Second, the collisionaldeactivation rate coefficient is now computed from reduced collisional integrals�(2,2) (Reid et al. 1987) rather than set equal to the kinetic theory of gasesformula for the hard-sphere collision frequency; this change has the effect ofmodifying slightly the

√T temperature dependence of collision frequency given

by the kinetic theory formula.

Page 14: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

138 Anthony M. Dean, Joseph W. Bozzelli

2.4 ANALYSIS OF HYDROGEN ATOMABSTRACTION REACTIONS

Elementary reactions in which a hydrogen atom is removed from a stable speciesin reaction with an atom or radical, called “abstraction reactions”, play importantroles in the combustion chemistry of nitrogen. Experimental rate data are availablefor only a small fraction of them, which forces us to introduce a number ofestimated abstraction rate coefficients. In order to deal with them self-consistentlywe developed the following estimation procedure. Discussion is limited toabstractions by H, O, OH, NH2, and CH3, the radicals likely to be in highestconcentrations under combustion conditions. We also estimate rate coefficientsfor HO2 abstractions for a limited number of reactions.

Rate coefficients were estimated by relating the parameters for a specific reactiontype, e.g., abstractions by OH, to those of a well-characterized hydrocarbon analog,e.g., OH + C2H6. Consistent use of such prototypical reference reactions avoidspropagating whatever inconsistencies are present in compilations. Moreover,it takes into account the different temperature dependences found for differenttypes of hydrogen abstractions. It is well-recognized (Tsang 1986; Baulch etal. 1992) that many hydrogen abstraction reactions show upward curvaturein their Arrhenius plots, leading to appreciably higher rate coefficients at hightemperatures than one would estimate from linear Arrhenius extrapolation of low-temperature measurements. Such temperature dependence has to be includedin order to predict reliable rate coefficients under combustion conditions. Aconvenient way to include this dependence is to express the rate coefficient inthe form A T m exp(−E/RT ). Zellner (1984) shows that m in this equation isexpected to have maximum values of 1.5 for abstraction by atoms, 2 for abstractionby diatomic radicals, and 3 for abstraction by polyatomic radicals.

Our procedure is as follows:

(1) The values of A and m for all reactions in a homologous series are takenfrom a single reference reaction. For abstractions by H, O and OH, we usethe expressions recommended by Baulch et al. (1992) for the correspondingabstractions from C2H6 and assume that A scales in proportion to the numberof equivalent abstractable hydrogen atoms. The preexponential factors thustake the form AC2H6(nH/6), where nH is the number of equivalent hydrogenatoms that can be abstracted from the molecule of interest.

CH3, NH2 and HO2 have to be treated differently. The Baulch et al. ratecoefficient expression for CH3 + C2H6 −→ CH4 + C2H5 sets m to 6, muchlarger than the theoretical expectation. A substantial difference in m for theabstractions from C2H6 by OH, for which m = 2, and by CH3 is inconsistentwith the temperature dependence of the equilibrium constant for the reaction

Page 15: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 139

OH+ CH4 −→ CH3 + H2O.1 It is equally reasonable and more consistentto base the assignment of A and m for H-abstractions by CH3 on the basis ofthe rate coefficient for its reaction with H2O implied by the rate coefficientand equilibrium constant of OH + CH4 ⇀↽ H2O + CH3. The result is m =1.87, much more consistent with theory. (Zellner 1984) With A and m thusassigned for abstractions by CH3, a value of E for abstractions by CH3 isderived by fitting the Baulch expression for the rate coefficient of CH3 + C2H6.Although the resulting fit is imperfect, it is adequate for our purpose andensures thermodynamic consistency for the abstraction reactions of OH andCH3.

The m value for NH2 is similarly derived from the forward rate coefficientand the equilibrium constant of OH + NH3 ⇀↽ H2O + NH2, and an E value isderived from the reaction of NH2 with C2H6. For HO2, the reverse of OH +H2O2 was used. Values for A and m are listed in Table 3.

(2) The value of E is estimated using the “Evans-Polanyi relationship”, whichstates that the activation energy E for a set of reactions of similar type butdifferent enthalpy change of reaction is proportional to the enthalpy change ofreaction. This can be written as

E = Eref − f (1r H ◦ref,298 −1r H ◦298) ,

where the ref subscript denotes the reaction in the set chosen as a referenceand the proportionality constant f is known as the Evans-Polanyi factor. Forabstractions from hydrocarbons we estimated f using ethane as the referencemolecule. For each reaction class, we selected the f -values that gave theoverall best fits to representative data listed in the NIST compilation (Mallardet al. 1993) and other available sources. Values of f for abstractions formingalkyl radicals are included in Table 3.

For formation of nitrogen-centered radicals by hydrogen abstraction, ammo-nia is used as the reference. The scarcity of data for these reaction types forcedus to select f in a more arbitrary manner. For H, O, and OH, our assignmentsare based on the differences in reported E values for abstractions from NH3 andHNO. The data for abstraction from NH3are well-characterized experimentally(cf. Sections 5.6 through 5.8); transition state theory results were adopted forthe H-abstraction reactions from HNO by OH and H (cf. Section 7.11).Combining these data gives values for fH and fOH. Combining experimentaldata for O + NH3 with the assumption that O + HNO is comparable to H +HNO (cf. Section 7.11) yields fO. We obtained fNH2 from the rate coefficientsfor NH2 + C2H6 and NH2 + H2O. The data for CH3 + NH3 and CH3 + N2H4were combined to yield the fCH3 values listed listed in Table 3.

1 This reaction is a linear combination of the abstraction reactions of OH andCH3 from C2H6; for this reason the temperature dependence of its equilibriumconstant imposes a consistency requirement on the temperature dependence of therate coefficients for abstraction of H atoms from C2H6 by OH and CH3.

Page 16: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

140 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.3. Estimation parameters for hydrogen atom abstraction reactions.1, 2

Radical A m Eref/kJ 1r H ◦ref,298/kJ f

Formation of carbon-centered radicals; ethane reference.

H 2.4× 108 1.5 31.0 –13.0 0.65O 1.7× 108 1.5 24.3 –4.6 0.75OH 1.2× 106 2.0 3.8 –76.6 0.50NH2 9.2× 105 1.94 30.1 –31.4 0.23CH3 8.1× 105 1.87 44.4 –15.5 0.65HO2 1.4× 104 2.69 79.1 53.1 0.60

Formation of nitrogen-centered radicals; ammonia reference.H 2.4× 108 1.5 47.3 18.0 0.21O 1.7× 108 1.5 31.4 26.8 0.15OH 1.2× 106 2.0 3.3 –45.2 0.05CH3 8.1× 105 1.87 41.8 15.9 0.15

1 Only limited high-temperature rate data are available for the unusually exothermic abstrac-tion reactions that produce resonance-stabilized radicals. The rate coefficient expressionsreported for them are generally in the Arrhenius form, i.e., without a temperature-dependentpreexponential factor. Using such expressions together with three-parameter expressionsfor reactions that produce alkyl-type radicals would lead to significant underestimates inrate coefficient assignments at combustion temperatures for reactions involving resonance-stabilized radicals. Although these abstractions are typically about 50 kJ/mol more exother-mic than those that form normal radicals, the limited data available (Tsang 1991) suggest thattheir abstraction reactions are only slightly faster than reactions of normal exothermicity.Using the usual Evans-Polanyi factors would overestimate their rate coefficients. Onecan either employ different, i.e., smaller correction factors for these reactions, or one canassume that these abstractions will have an activation energy about 4 kJ/mol lower thantheir non-resonant analogs but with the same m-exponent.

Additional complications arise in estimating A-values for these reactions. Formationof the resonantly stabilized radical could restrict rotation about the radical center, loweringthe entropy of the transition state and correspondingly decreasing the A-factor. Hydrogenatom abstractions from aldehydes and from carbon atoms adjacent to carbonyl groups alsorequire special treatment.2 A complication that can arise when using the Evans-Polanyi relationship for reactions

with very large exothermicities is that the resulting activation energy can become negative,which is physically meaningless for an abstraction reaction. To avoid this, a lower limit of−300mR is given to E . The activation energy in the 2-parameter Arrhenius form mustthen be positive at 300 K.

A distinguishing feature of nitrogen chemistry is participation of long-livedradicals, e.g., NH2O. These molecules require substantially more energy thantheir hydrocarbon analogs to break the “β” bond (the one on the far side of the

Page 17: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 141

atom adjacent to a radical site) to form an unsaturated molecule and a smallerradical. The increased endothermicity for these “β-scission” reactions in nitrogenchemistry results in much lower rate coefficients for dissociation. As a result,radical-radical reactions are more important in nitrogen than in carbon chem-istry. Recombination reactions of radicals occur via chemically-activated adducts,whose pressure-dependent reactivity can be described as outlined in the precedingsection. However, estimates of the rate coefficients for “disproportionation”reactions, in which a hydrogen atom is transferred from one radical to anotherto form two stable products, are less reliable, because there is little informationavailable about this type of reaction for nitrogen species.

Rate coefficients for disproportionation reactions of hydrocarbon radicals havebeen estimated in the past by collision theory models. (E.g., by Benson 1985). Weprefer to estimate rate coefficients for these reactions as analogs of abstractionsfrom stable molecules. Since radical-radical hydrogen transfer reactions aregenerally more exothermic than those involving a radical and a stable molecule,E must frequently be assigned the lower limit value −300mR. (Cf. footnote toTable 3.) One would expect the transition state for hydrogen transfer between tworadicals to be more extended than for an abstraction from a stable molecule, sincebond strengths of hydrogens being transferred are usually weaker in radicalsthan in stable molecules. This means that our rate coefficient estimates fordisproportionation reactions are lower limits. Comparisons of rate coefficientsestimated with this method to values in the NIST data base (Mallard et al. 1993)for hydrocarbon reactions, confirm, however, that this method does provide closeestimates, especially at the higher temperatures of interest in combustion. Specificcomparisons for the few instances where information is available for reactionsinvolving nitrogen are discussed later in this chapter.

In later sections we refer to rate coefficients estimated by the approach describedhere as DHT (for “Direct Hydrogen Transfer”) estimates.

2.5 UPDATED RATE COEFFICIENTS FORTHE H/N/O SYSTEM

In this section we review reactions considered by Hanson and Salimian (1984)where there have been significant later measurements but where extensive theo-retical analysis is not necessary.

2.5.1 O + N2 −→ N + NO

There have been several new measurements, in quite remarkable agreement, asshown in Fig. 1. Thielen and Roth (1984) used atomic resonance absorptionspectroscopy to monitor N and O atoms produced from shock-wave heatedN2/N2O/Ar mixtures over the temperature range 2400 to 4100 K. Their ratecoefficient expression, 1.8 × 1014 exp(−38300/T) cm3mol−1s−1, agrees well

Page 18: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

142 Anthony M. Dean, Joseph W. Bozzelli

2

4

6

8

10

0.4 0.6 0.81000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Thielen-Roth (1984)

Hanson-Salimian (1984)

Davidson-Hanson (1990, reverse)

Michael-Lim (1992, reverse)

Zabielski-Seery (1989)

Koshi et al. (1990, reverse)

O + N2 → NO + N

FIGURE 2.1. Arrhenius plot for O + N2 −→ NO + N. The recommended expression isbased on the measurements by Davidson and Hanson (1990) for the reverse reaction.

with the recommendation of Hanson and Salimian. Davidson and Hanson (1990a)measured the reverse reaction rate by measuring NO disappearance behind shockwaves from 1400 to 3500 K. Their result and the well-established thermochemistryof this reaction led to the forward reaction rate coefficient expression 1.95 ×1014 exp(−38660/T) cm3mol−1s−1, in agreement with the other measurements.Koshi et al. (1990), using NO as a source for N–atoms, reported a slightly lowervalue for the reverse reaction rate over a narrower temperature range. Michaeland Lim (1992) also studied the reverse reaction and derived the forward reactionrate coefficient expression 1.0× 1014 exp(−37988/T ) cm3mol−1s−1. Zabielskiand Seery (1989) reported k = 1.99× 1014 exp(−38133/T) cm3mol−1s−1 frommeasurements of NO production in oxygen-enriched methane flames. Thus this

Page 19: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 143

rate coefficient for production of NO via the Zeldovich mechanism seems to bewell-characterized. We suggest adoption of the Davidson and Hanson expression

k1 = 1.95× 1014 exp(−38660/T ) cm3mol−1s−1

For the 1500 to 2500 K regime most relevant to combustion it appears that therecommended expression should be accurate to at least ±25%. Such a spreadencompasses most of the recent measurements with the exception of the highertemperature data of Michael and Lim and the data of Koshi et al., both of whichare lower than given by the recommended expression.

Michaud et al. (1992) suggest that this reaction proceeds by attack of singletoxygen atoms, O(11), rather than ground state O(3P) atoms. They report that thismechanism is consistent with the literature data if it is assumed that the 11 stateis equilibrated with the 3P ground state, but point out that it is not consistent withmeasurements of the reverse reaction rate by Felder and Young (1972), whereO(11) production accounted for less than 10% of the overall reaction. It wouldappear that although the rate of this reaction is well characterized, additional workwould be helpful to establish its mechanism.

2.5.2 NO + Ar −→ N + O + Ar

At the time of the Hanson-Salimian review the reported values showed widevariations and no recommendation was made. The subsequent measurements ofThielen and Roth (1984) are quite close to the earlier work of Wray and Teare(1962). (Fig. 2). While the higher values of Koshi et al. (1978) and the lowervalue of Myerson (1974) result in a substantial spread in measured values, we feelthat the measurements of Thielen and Roth are to be preferred, essentially becausethe same type of analysis yielded good agreement for reaction 1. Accordingly, werecommend

k2 = 9.6× 1014 exp(−74700/T) cm3mol−1s−1

over the temperature range 2400 to 6200 K. This recommendation is consistentwith that of Tsang and Herron (1991) .

2.5.3 N2O + Ar −→ N2 + O + Ar

There have been several new studies of this reaction in shock tubes (Roth and Just1984; Frank and Just 1985; Fujii et al. 1989; Zuev and Starikovskii 1991a; Michaeland Lim 1992) and in flow reactors (Johnsson et al. 1992). These are compared tothe Hanson and Salimian recommendation in Fig. 3. The remarkable agreementtestifies that accurate rate coefficient determinations can indeed be made in well-chosen systems. The extended temperature range now covered by experiments

Page 20: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

144 Anthony M. Dean, Joseph W. Bozzelli

3

5

7

9

11

0.2 0.3 0.4 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Koshi et al. (1978)

Thielen-Roth (1984)

Wray-Teare (1962)

Myerson (1974)

NO + M → N + O + M

FIGURE 2.2. Arrhenius plot for NO dissociation. The recommended expression is thatreported by Thielen and Roth (1984).

allows a recommendation to be made for a larger range. We recommend theexpression suggested by Johnsson et al.

k3 = 4.0× 1014 exp(−28230/T) cm3mol−1s−1

for M = Ar at temperatures from 1000 to 3000 K. Subsequent measurements byR�ohrig et al. (1996) for temperatures from 1700 to 3000 K and pressures from 0.3to 6 atm agree with this recommendation. These authors also report measurementsat pressures up to 450 atm which confirm that the reaction is near the low pressurelimit for pressures up to 10 atm at temperatures over 1500 K. Falloff parametersare given by Tsang and Herron (1991) for N2 and CO2 bath gases.

Page 21: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 145

2

4

6

8

10

12

0.3 0.5 0.7 0.9 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Fujii et al. (1989) Ar

Frank-Just (1985) Ar

Hanson-Salimian (1984) Ar, Kr

Roth-Just (1984) Ar

Michael-Lim (1992) Ar, Kr

Zuev-Starikovskii(1991) Ar

Johnsson et al. (1992) Ar

Johnsson et al. (1992) N2

Johnsson et al. (1992) Ar, combined

N2O + M → N2 + O + M

FIGURE 2.3. Arrhenius plot for N2O dissociation. The recommended expression is thecombined one reported by Johnsson et al. (1992).

2.5.4 O + N2O −→ Products

This reaction has two channels:

O+N2O −→ N2 +O2 (4a)

O+N2O −→ 2 NO . (4b)

The recent measurements are summarized in Figs. 4 and 5. Hidaka et al. (1985)reported rate coefficients for these reactions from a modeling analysis of N2O/Arand N2O/H2/Ar mixtures heated in shock waves. Their assessment is similarto that of Hanson and Salimian in that these reactions have rate coefficientsof comparable magnitude, but Hidaka et al. assigned larger values than thosesuggested earlier. Zuev and Starikovskii (1991b) reported values of k4a much

Page 22: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

146 Anthony M. Dean, Joseph W. Bozzelli

9

10

11

12

0.4 0.6 0.81000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Hikada (1985)

Hanson-Salimian (1984)

Dean-Steiner (1977)

Zuev-Starikovskii (1991)

Davidson et al. (1991)

O + N2O → N2 + O2

FIGURE 2.4. Arrhenius plot for O + N2O −→ N2 + O2. The recommended expressionis that reported by Davidson et al. (1991).

lower than recommended by Hanson and Salimian, but similar values of k4b.These reactions were reinvestigated by Davidson et al. (1991a) , who reportedk4b values similar to earlier ones but found that k4a is noticeably smaller athigher temperatures. Earlier work (Dean and Steiner 1977) had also reportedvalues somewhat lower than those recommended by Hanson and Salimian forboth channels. The Davidson et al. (1991a) experiments have the advantage ofdirect observation of products from each channel (O2 from 4a and NO from 4b)

Page 23: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 147

9

10

11

12

0.4 0.6 0.81000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Hikada (1985)

Hanson-Salimian (1984)

Dean-Steiner (1977)

Zuev-Starikovskii (1991)

Davidson et al. (1991)

O + N2O → 2 NO

FIGURE 2.5. Arrhenius plot for O + N2O −→ 2 NO. The recommended expression isthat reported by Davidson et al. (1991)

and for this reason we recommend their expressions

k4a = 1.4× 1012 exp(−5440/T ) cm3mol−1s−1

k4b = 2.9× 1013 exp(−11650/T) cm3mol−1s−1

realizing that k4a is still uncertain.This uncertainty is further amplified by the reports of R�ohrig et al. (1996),

Ross et al. (1997) and Meagher et al. (1997). R�ohrig et al. found that using the

Page 24: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

148 Anthony M. Dean, Joseph W. Bozzelli

Hanson and Salimian expression for k4a gave better fits to their N2O profiles thanthe Davidson et al. expression; Ross et al. report that the overall rate coefficientrequired to fit their O2 profiles near 1400 K is much lower than even the Hansonand Salimian expression; while the Meagher et al. direct measurements from1076 to 1276 K are significantly larger than given by the Hanson and Salimianexpression.

2.5.5 NH3 + Ar −→ NH2 + H + Ar

The measurements of Davidson et al. (1990b) agree with the expression recom-mended by Hanson and Salimian. Accordingly, that recommendation,

k5 = 2.5× 1016 exp(−47200/T) cm3mol−1s−1

for M = Ar over the temperature range 1740 to 3450 K, is confirmed.

2.5.6 NH3 + H −→ NH2 + H2

Hanson and Salimian (1984) noted the appreciable scatter in the high temperaturerate coefficients reported for this reaction. Since 1984 there have been severaldirect measurements of this reaction rate (Hack et al. 1986; Marshall and Fontijn1986; Michael et al. 1986; Sutherland and Klemm 1987; and Ko et al. 1990),all in agreement with one another and all substantially faster than the expressionrecommended by Hanson and Salimian.

These newer results and the Hanson and Salimian recommendation are shownin Fig. 6. We recommend that the combined expression developed by Ko et al.(1990),

k6 = 5.4× 105T 2.40 exp(−4990/T ) cm3mol−1s−1

be used over the temperature range 490 to 1780 K. This expression combines theirmeasurements from 490 to 960 K with those at higher temperatures by Michael etal. (1986) and Hack et al. (1986).

2.5.7 NH3 + OH −→ NH2 + H2O

Recent experiments have confirmed the marked non-Arrhenius behavior of thisrate coefficient. (Zabielski and Seery 1985; Jeffries and Smith 1986) Both groupscombined their results with earlier work and developed non-Arrhenius fits, asshown in Fig. 7. At high temperatures their expressions are in good agreement withthe expression recommended by Hanson and Salimian, but the newer expressionsare much higher at lower temperatures. The significantly lower high temperaturemeasurement by Fujii et al. (1986), also indicated in Fig. 7, was derived by fittinga computer model to data taken using shock-heated NH3–N2O–Ar mixtures and isthus more indirect. This reaction was critically evaluated by Cohen and Westberg

Page 25: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 149

8

9

10

11

12

0.5 1.0 1.5 2.0 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Michael et al. (1985) Ko et al. (1990, combined) Michael et al. (1986) Marshall-Fontijn (1986) Hack et al. (1986) Sutherland-Klemm (1987) Hanson-Salimian (1984) Ko et al. (1990)

H + NH3 → NH2 + H2

FIGURE 2.6. Arrhenius plot for H + NH3 −→ NH2 + H2. The recommended expressionis the combined one reported by Ko et al. (1990).

(1991). Their recommendation,

k7 = 5.0× 107T 1.6 exp(−480/T ) cm3mol−1s−1,

also displayed in Fig. 7, describes the data well over the temperature range from225 to 3000 K and we accordingly recommend it.

Page 26: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

150 Anthony M. Dean, Joseph W. Bozzelli

11

12

13

1 2 3 4 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Fujii et al. (1986) Diau et al. (1990) Zabielski-Seery (1985) Atkinson et al. (1989) Hanson-Salimian (1984) Jeffries-Smith (1986, combined) Zabielski-Seery (1985, combined) Cohen-Westberg (1991)

OH + NH3 → NH2 + H2O

FIGURE 2.7. Arrhenius plot for OH + NH3 −→ NH2 + H2O. The recommendedexpression is that of Cohen and Westberg (1991).

2.5.8 NH3 + O −→ NH2 + OH

Here again there has been significant progress since the Hanson-Salimian reviewof 1984. In addition to two new high temperature measurements (Fujii et al. 1986;Sutherland et al. 1990) and one low temperature study (Perry 1984), there havebeen an exhaustive review (Cohen 1987) and a critical evaluation (Baulch et al.1992).

This information is summarized in Fig. 8. Given the direct nature of both theSutherland et al. and the Perry experiments we suggest use of the Sutherland et al.

Page 27: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 151

9

10

11

12

1.0 2.0 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Fujii et al. (1986) Sutherland et al. (1990) Sutherland et al. (1990, combined) Hanson-Salimian (1984) Cohen (1987) Perry (1984) Baulch et al. (1992)

O + NH3 → NH2 + OH

FIGURE 2.8. Arrhenius plot for O + NH3 −→ NH2 + OH. The recommended expressionis the combined one reported by Sutherland et al. (1990).

expression

k8 = 9.4× 106T 1.94 exp(−3250/T ) cm3mol−1s−1,

which represents both data sets over the temperature range 450 to 1800 K. Weprefer this expression at high temperatures to the slightly lower one suggestedby Cohen since the latter is based upon a transition state theory extrapolationof Perry’s value rather than measurements. These data also suggest that therecommendation of Baulch et al. (1992) is not an adequate representation of themore recent experiments.

Page 28: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

152 Anthony M. Dean, Joseph W. Bozzelli

Summarizing the data in Figs. 6–8, it appears that the rate of NH3 consumptionin flames should be accurately described by the now thoroughly investigatedelementary reactions with the principal reactive species in flame gases. The mostprominent consequence of this is that the production of NH2, the key intermediatein selective noncatalytic reduction processes for NO suppression, can be regardedas reasonably well understood. This allows us to focus on the subsequentreactions of NH2 (and later chemistry) to develop improved understanding ofthese processes. We will address this topic several times in later sections.

2.6 QRRK TREATMENTS

2.6.1 H + NH2 −→ NH + H2

There is one new measurement (Davidson et al. 1990b) to supplement the sparsedata base on this reaction, which has also been reviewed by Baulch et al. (1992).These results are compared to the tentative Hanson-Salimian recommendation inFig. 9. For completeness, the earlier measurements of Yumura and Asaba (1980)are also included. The Davidson et al. expression is consistent with the recommen-dation of Hanson and Salimian. The experimental scatter, however, suggests thatadditional study of this reaction is needed. The mechanism of this radical-radicalreaction could be either direct hydrogen transfer or a complex recombination-elimination path. Spin conservation on the ground state NH3 surface wouldrequire that the combination-elimination pathway produce 1NH + H2, a channelappreciably higher in energy than the one producing ground state 3NH + H2. Apotential energy diagram for this pathway on the 1NH3 surface is shown in Fig. 10;the QRRK parameters are listed in Table 4.

TABLE 2.4. H + NH2 −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) H + NH2 −→NH3 1.2×1015 T−0.4 0(–1) NH3 −→H + NH2 2.6×1015 454(2) NH3 −→ 3NH + H2 6.0×1013 444(3) NH3 −→ 1NH + H2 1.8×1015 550

Molecular constants of NH3 adduct: ν1 = 1101 cm{1 (degen=2.10) ν2 = 2430 cm{1(2.33)ν3 = 3973 cm{1(1.57); σ = 3.49 A and ε/K = 310.

Rate coefficients: k1 based on H + CH3 rate coefficient expression of Tsang et al. (1986);k−1 from microscopic reversibility (MR); k2 A from TST estimate; Ea estimated to givea 42 kJ/mol barrier for the reverse reaction; k3 via k−3 and MR based on the 1CH2 + H2reaction.

Page 29: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 153

12

13

0.4 0.6 0.8 1.0 1000 K / T

log

( k

/ cm

3 mol

–1 s

–1)

Davidson et al. (1990)Hanson-Salimian (1984)Baulch et al. (1992)Yumura et al. (1981)QRRK, triplet barrier 444 kJQRRK, triplet barrier 423 kJabstraction estimateabstraction (reverse)Linder et al 1995Xu et al (1995)

H + NH2 → NH + H2

FIGURE 2.9. Arrhenius plot for H + NH2 −→ NH + H2. The QRRK predictions ofreaction via combination–elimination are apparently somewhat low even with the tripletchannel barrier reduced by 21 kJ/mol. Both the radical–radical hydrogen transfer estimateand the one based on the reverse reaction are close to the measurements of Davidson etal. Two sets of ab initio calculations (Linder et al. 1995; Xu et al. 1995) for the directabstraction reaction are also included.

We are not aware of any quantum-mechanical computations for eliminationvia the triplet channel. Comparison with other non-spin-conserving processes(Yarkony 1993) suggests that the barrier is about 40 kJ/mol above the enthalpychange for this channel. One would anticipate little or no barrier above theenergetic constraint for singlet NH elimination, because the reverse reaction of1NH insertion should be analogous to the isoelectronic 1CH2 insertions, wherebarriers are observed to be very low. Even assuming the barrier to 3NH + H2 tobe 10 kJ/mol lower than that for dissociation of the complex back to the reactantsH + NH2, the much higher A-factor for the simple fission pathway results in

Page 30: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

154 Anthony M. Dean, Joseph W. Bozzelli

Ent

halp

y (k

J/m

ol)

NH2 + H

NH3

1NH + H2

3NH + H2

50

450

FIGURE 2.10. Potential energy diagram for the recombination-elimination pathway forNH2 + H on the ground state surface of NH3. The triplet channel pathway is shown by thedashed line.

dissociation of the adduct back to reactants being the dominant channel. Thus thecalculated rate coefficient for the 3NH + H2 channel is lower than the expressionreported by Davidson et al. These results are shown in Fig. 9. Lowering thebarrier to the 3NH channel by 21 kJ/mol, so that the reverse barrier is now only21 kJ/mol, serves to increase the rate coefficient slightly, but, as seen in Fig. 9,this expression is still well below the Davidson et al. measurements. This findingleads us to conclude that the rate coefficient expression measured by Davidson etal. can not be rationalized in terms of a recombination-elimination mechanism.

Another possible mechanism for this reaction is direct hydrogen transfer withouttransient formation of an intermediate that can be stabilized. The estimationprocedures discussed in Section 4 were used for both directions of this reaction,with the k−9 estimate derived on the assumption that hydrogen abstractions by NHcan be expected to have A-factors similar to hydrogen abstractions by OH. Theactivation energy for k−9 was estimated from reactions with similar enthalpiesof reaction. Both approaches lead to expressions with similar values for k9,but somewhat greater temperature dependence, than were found by Davidson etal. (Figure 9) The agreement does support direct hydrogen abstraction, and nopressure dependence of the rate coefficient is to be expected.

Similar conclusions were drawn from theoretical investigations of the directabstraction mechanism by Linder et al. (1995) and Xu et al. (1995), both ofwhom found a temperature dependence similar to that found from our estimationprocedure. We suggest use of the expression we derived for the forward reaction

Page 31: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 155

rate coefficient

k9 = 4.8× 108T 1.5 exp(−3995/T ) cm3mol−1s−1,

which is in reasonable agreement with the measurements of Davidson et al.

2.6.2 HO2 + NO −→ NO2 + OH

A low-temperature measurement (Jemi-Alade and Thrush 1990) appeared after theHanson-Salimian review. It is consistent with the recommendation of Atkinsonet al. (1989), which bases its recommended temperature dependence on the mea-surements of Howard (1980) over the range 423–1271 K with a slight difference inthe pre-exponential term. The Hanson-Salimian recommendation was also basedon the measurements of Howard. These expressions are compared in Fig. 11.

We suggest using the expression recommended by Atkinson et al. with thetemperature range extended to 1250 K:

k10 = 2.2× 1012 exp(+240/T ) cm3mol−1s−1.

TABLE 2.5. HO2 + NO −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) HO2 + NO −→HOONO 5.6×1012 0(–1) HOONO −→HO2 + NO 8.5×1014 120(2) HOONO −→OH + NO2 2.0×1014 87(3) HOONO −→HONO2 7.1×1011 105

(–3) HONO2 −→HOONO 1.4×1013 218(4) HONO2 −→OH + NO2 1.3×1015 206

Molecular constants of HOONO adduct: ν1 = 338 cm{1(degen=3.56) ν2 = 1271 cm{1(4.30)ν3 = 3994 cm{1(0.14); molecular constants of HONO2 (nitric acid): ν1 = 622 cm{1(4.55)ν2 = 1422 cm{1(3.52) ν3 = 3868 cm{1(0.93); σ = 4.78 A and ε/K = 486 for both adducts.

Rate coefficients: k1 this evaluation—see references in text; k−1 from microscopicreversibility (MR); k2 via k−2 and MR, , k−2 set equal to the OH + NO rate coefficientof Atkinson et al. (1989); k3 TST with loss of 2 rotors and optical isomer, ∆S‡ = –35.6J/K; Ea is an estimate for the lower limit [Ea = 128 kJ/mol for NH2NO −→ HNNOH(Melius and Binkley 1984b)]; k−3 MR; k4 via k−4 and MR; k−4 same as k−2.

One might expect this reaction to have an addition-elimination mechanismvia the HOONO adduct, as shown in the potential energy diagram of Fig. 12.The products OH + NO2 are about 33 kJ/mol and the HOONO adduct about120 kJ/mol more stable than the reactants (Melius 1988). Treatment of thissystem as a chemically-activated reaction with the kinetic and thermodynamic

Page 32: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

156 Anthony M. Dean, Joseph W. Bozzelli

12.5

13.0

1 2 3 41000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Hanson-Salimian (1984)

Atkinson et al. (1989)

Jemi-Alade and Thrush (1990)

Howard (1980)

QRRK (0.01 atm)

QRRK (10 atm)

HO2 + NO → NO2 + OH

FIGURE 2.11. Arrhenius plot for HO2 + NO −→ NO2 + OH. The recommendedexpression is that of Atkinson et al. (1989) extended to 1250 K. Note that the QRRKcalculations indicate virtually no dependence on pressure.

assignments shown in Table 5 gives the rate coefficients shown in Fig. 11. Bothdissociation channels of the adduct are expected to have loose transition statesand correspondingly high Arrhenius A-factors for unimolecular dissociation. Thelower energy of the OH + NO2 channel results in rapid dissociation of the adductvia this channel rather than dissociation to the original reactants. Furthermore,dissociation is much more rapid than collisional stabilization. Thus one expectsto see little pressure dependence even though the reaction proceeds via a transientcomplex. The predicted rate coefficient for the NO2 + OH channel is essentiallyinvariant over a factor of 1000 in pressure; even at 10 atm the stabilization channelcontributes only a very small fraction to the total rate (Fig. 11).

Page 33: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 157

–50

0

Ent

halp

y (k

J/m

ol)

HO2 + NO

HOONO

OH + NO2

50

100

150

FIGURE 2.12. Potential energy diagram for the recombination-elimination pathway forHO2 + NO. The curve forming HONO2 (nitric acid, at−134 kJ/mol) is not shown.

Another possible product channel not shown on Fig. 11 would be HNO + O2,which can occur via isomerization of the HOONO adduct to HN(O)OO followedby dissociation. However, this isomerization would be expected to have a relativelylow A-factor, because of the loss of internal rotors in the transition state, and alsoa high barrier to form a diradical, which eliminates this route from consideration.A third possible channel is isomerization to nitric acid, HONO2, which would bestabilized at low temperatures and dissociate to OH + NO2 at high temperatures.This isomerization would also be expected to have a lower A-factor. To ourknowledge the barrier has not been calculated; by analogy to the NH2NO systemone can estimate that it would be greater than the barrier to OH + NO2. Resultsfrom the QRRK analysis with the barrier set at the reasonable lower limit of105 kJ/mol indicated that the HONO2 channels, both stabilization and subsequentreaction to OH + NO2, are about 1000 times less important than the OH + NO2channel via the HOONO adduct.1

The observed negative temperature dependence can be attributed to fall-offeffects, i.e., the growing importance of the dissociation of the addition complexback to reactants. At higher temperatures the larger A-factor for adduct dissoci-ation back to reactants leads to increasing contributions from this channel and apredicted decrease in the observed rate coefficient. The somewhat lower A-factorfor dissociation of the complex to OH + NO2 (cf. Table 5) is based upon themeasured expression for the reverse reaction rate coefficient.

1 The reverse reaction OH + NO2, however, would be expected to form nitricacid at a rate comparable to forming the HOONO adduct. It is not an importantreaction in combustion for the reasons discussed in Chapter 1.

Page 34: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

158 Anthony M. Dean, Joseph W. Bozzelli

Another possible explanation for the observed negative temperature dependenceof the rate coefficient would be that the high-pressure addition rate coefficient hasa slight negative temperature dependence. If this is assumed to be the case, thenthe computed QRRK rate coefficient expressions could be brought into betteragreement with experiment in the middle of the temperature range. We did notpursue this assumption because the purpose of our analysis was not to replicatethe experimental rate coefficient but only to show that the observed behavior isconsistent with an addition-elimination reaction. The analysis strongly suggeststhat one does not expect any pressure effects in this system even though it appearsnot to be a direct reaction. The rate coefficient expression suggested by Atkinsonshould therefore be applicable at all pressures of interest in combustion.

2.6.3 H + N2O −→ Products

A potential energy diagram, following Marshall et al. (1987), for this reaction isshown in Fig. 13. Aspects of it have been addressed in several investigations(Miller and Melius 1992a; Durant 1994; Melius 1993; Walch 1993a). Although adistinction was made in some of these studies between cis- and trans-HNNO, theoverall features are in each case similar to those shown in Fig. 13. We selecteda barrier height of 38 kJ/mol for addition to form HNNO, representative of thevalues reported by the cited authors, and a barrier of 77 kJ/mol for direct additionto form NNOH, again close to the reported values. Following Melius (1993) weuse a barrier of 123 kJ/mol for the isomerization reaction of HNNO to NNOH.

NNH + O

NH + NO

N2 + OH

NNOHHNNO

H + N2O

250

Ent

halp

y (k

J/m

ol)

50

150

350

450

FIGURE 2.13. Potential energy diagram for H + N2O. Note that both adducts can beformed directly from the reactants.

Page 35: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 159

There are four product channels:

H+N2O −→ HNNO (11a)

H+N2O −→ N2 +OH (11b)

H+N2O −→ NH+ NO (11c)

H+N2O −→ NNH+ O . (11d)

Fig. 13 suggests that the major channels, particularly at lower temperatures, wouldbe HNNO or N2 + OH. At higher temperatures the initially formed adduct couldhave sufficient energy to dissociate to one of the higher energy exit channels.

Reaction 11b has been reasonably well characterized. Hidaka et al. (1985)used a computer model in conjunction with their shock tube observations inN2O–Ar and N2O–H2–Ar mixtures, while Fontijn and coworkers (Marshall etal. 1987; Marshall et al. 1989) measured the reaction directly in a high-temperaturephotochemical experiment. There is a difference in the higher-temperature ratecoefficients in the two photochemical studies. The authors point out that their ear-lier work employed a capped inlet which might have led to premature dissociationof some of the N2O, thus introducing some error. As a result, we only considerthe later data. These rate coefficients, as well as an earlier shock- tube modelinganalysis by Dean et al. (1980), are compared to the recommended expressionsgiven by Hanson and Salimian (1984) and Tsang and Herron (1991) in Fig. 14.

Data for the NH + NO channel (Reaction 11c), which has usually been studiedin the reverse direction, as described below, are more scattered. The more recentinformation is shown in Fig. 15. Cattolica et al. (1982) used measured NH andOH flame profiles in N2O/H2 flames to obtain estimates for this rate coefficient.The two shock tube studies of the reverse reaction (Roose 1981; Mertens etal. 1991a) report markedly different expressions. Also included in this plot is theexpression obtained by using the recommended expression given by Miller andBowman (1989) for the reverse reaction; their assignment is based primarily uponcomparison of computed and experimental N2O profiles in ammonia flames.

A confusing aspect of the NH + NO reaction is that it can yield multipleproducts:

NH+NO −→ N2 +OH (11e)

NH+NO −→ H+N2O (11f)

NH+NO −→ NNH+O . (11g)

Mertens et al. (1991a) report a branching fraction for the N2 channel k11e/ktot =0.19 � 0.10 at 3000 K from measurements of NH and OH. Shock tube measure-ments of emissions from electronically excited OH and NH in HNCO–NO–Armixtures (Yokoyama et al. 1991a; Yokoyama et al. 1991b) were used with acomplex model to infer a branching fraction of 0.32 � 0.07 for the N2 channelat 3500 K. However, an analysis of ammonia flames (Vandooren et al. 1991)suggested that the N2 channel is 4–6 times faster than the N2O channel at 2000 K.The more direct nature of the Mertens et al. study suggests that their data are themost reliable.

Page 36: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

160 Anthony M. Dean, Joseph W. Bozzelli

7

9

11

13

0.5 1.0 1.5 2.0 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Hikada et al. (1985)

Hanson-Salimian (1984)

Dean et al. (1980)

Marshall et al. (1989)

Tsang-Herron (1991)

QRRK (N2 + OH)

H + N2O → N2 + OH

FIGURE 2.14. Arrhenius plot for H + N2O −→ N2 + OH. The recommended expressionis that of Marshall et al. (1989) for temperatures above 1000 K. The QRRK calculationsinclude contributions to this channel from both the HNNO and the NNOH adducts.

At low temperatures, the measurements of the overall rate coefficient exhibitquite good agreement, but there is controversy with respect to the products. Har-rison et al. (1986) suggest that their measurements are best interpreted assumingthat the OH + N2 channel is the dominant one. Yamasaki et al. (1991) reportdirect measurements of OH at 300 K and claim OH + N2 is the exclusive channel.They also reported failure to detect H, supporting this assignment. However, morerecent work from that laboratory (Okada et al. 1994) reverses this assignment,now reporting that the H + N2O channel contributes 65% and OH + N2 30%. Thesenew assignments are consistent with a study by Durant (1994), who reported thatH + N2O is the dominant channel at room temperature and the branching ratio is0.8� 0.4. Lillich et al. (1994) report that the OH channel contributes 15� 5% at300 K. Calculations by Harrison and Maclagan (1990) have been used to support

Page 37: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 161

9

11

13

0.4 0.6 0.8 1.01000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Cattolica et al. (1982)

Roose (1981, reverse)

Mertens et al. (1991, reverse)

Miller-Bowman (1989, reverse)

QRRK (NH + NO)

H + N2O → NO + NH

FIGURE 2.15. Arrhenius plot for H + N2O −→ NO + NH. The “reverse” notation signi-fies that the rate coefficient was obtained via the equilibrium constant from measurementsof the reverse reaction rate coefficient. The QRRK calculations are in agreement with theexpression recommended by Miller and Bowman (1989).

the claim of OH + N2 as the dominant channel. However, other calculations byMelius and Binkley (1984b) and by Miller and Melius (1992a) indicate that theN2O channel is the dominant one over the entire temperature range. Modelingefforts in low pressure, stoichiometric H2–N2O–Ar flames (Sausa et al. 1993) haveused the branching fraction reported by Miller and Melius. Given this ambiguityconcerning the low temperature branching fractions, we have plotted the variousreported expressions for the total rate coefficient for NH + NO in Fig. 16. Althoughthe reported rate coefficients span a relatively narrow range around 1×1013

cm3mol−1s−1, there is a large variation in reported temperature dependence. Thehigh temperature shock tube results exhibit a substantially different temperature

Page 38: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

162 Anthony M. Dean, Joseph W. Bozzelli

dependence than the lower temperature data and the calculations of Miller andMelius. R�omming and Wagner (1996) reported shock tube measurements that gavea rate coefficient of about 1.0×1013 cm3mol−1s−1over the temperature range 1200to 2200 K. This result together with those of Mertens et al. (1991) and Lillich etal. (1994) suggests a minimum rate coefficient near 1400 K.

As evident from the potential energy diagram, this reaction is not a simple one.We have attempted to gain understanding of it by means of a QRRK analysis. Theinput parameters are listed in Table 6.

TABLE 2.6. H + NNO −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1a) H + NNO −→ HNNO 8.5×1013 38(–1a) HNNO −→ H + NNO 2.6×1013 104(1b) H + NNO −→ NNOH 1.3×1014 77

(2) HNNO −→ NH + NO 4.0×1015 209(3) HNNO −→ O + NNH 4.9×1015 258(4) HNNO−→NNOH−→N2+OH 6.8×1012 123

Molecular constants of HNNO adduct: ν1 = 567 cm{1 (degen=1.99) ν2 = 1174 cm{1 (2.49)ν3 = 2517 cm{1 (1.53); σ = 3.77 A and ε/K = 280.

Rate coefficients: k1a this study with Ea being a consensus value, cf. text; k−1a

microscopic reversibility (MR); k1b this study, Ea consensus value, cf. text; A basedupon H + NO2 from Ko et al. (1991); this channel, effectively a direct route to N2 + OH,is added to the chemical activation component as described in the text; k2 via k−2 and MRwith A−2 = 3.46×1013 and Ea(-2) = 0 from Harrison et al. (1986); k3 via k−3 and MRwith k−3 = 7.5×1013 [based upon O + C2H5, Tsang and Hampson (1986)]; k4 TST,∆S‡

= –13.8 J/K, Ea from Melius (1993)—this is the rate controlling step for the N2 + OHreaction.

We also explored use of a more complex reaction system, accounting separatelyfor the cis- and trans-isomers of HNNO, as described by both Walch (1993a) andDurant (1994); the results were very similar to those presented below. We preferthe simpler analysis, since there remains considerable ambiguity as to the relativeenergies of the isomers, cf. Walch (1993a). This reaction system is unusual inthat the addition reaction to form HNNO has a higher A-factor than the reverseunimolecular dissociation. Over the temperature range 300 to 2500 K the averageratio of Aforward/Areverse for H + NNO −→ HNNO is 3.2. Usually this ratio ismuch less than unity, strongly favoring the reverse reaction. For example, the ratiofor NH + NO −→ HNNO is 0.0086. This unusual case of a higher A-factor foraddition than for the (reverse) dissociation is a result of the unusually low standardentropy of the reactants H + N2O combined with loss of a rotor in the dissociation.The HNNO adduct will have a significantly higher probability for stabilization or

Page 39: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 163

12.5

13.0

13.5

1 2 3 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Mertens et al. (1991)

Harrison et al. (1986)

Miller-Bowman (1989)

Baulch et al. (1992)

Yamasaki et al. (1991)

Vandooren et al. (1991)

Miller-Melius (1992)

QRRK (total)

Lillich et al. (1994)

Hack et al. (1994)

NH + NO → Products

FIGURE 2.16. Arrhenius plot for NH + NO −→ Products. The QRRK predictions areseen to be in good agreement with the measured overall rate coefficients. Note the changein the predicted temperature dependence above 2000 K.

reaction to other channels (as opposed to redissociating to H + N2O) relative tomany of the other systems discussed elsewhere in this chapter. In particular, thisincludes formation of the higher energy products NH + NO, since this channel hasa more typical A-factor.

2.6.4 H + N2O −→ N2 + OH and H + N2O −→ HNNO

The products N2 + OH can be formed from two adducts. Addition of H to formNNOH will lead immediately to N2 + OH, and the rate coefficient for productionof N2 + OH via this channel is simply the rate of formation of the complex;no chemical activation calculation is needed. This rate coefficient is added tothat obtained from the chemical activation analysis for addition of H to form

Page 40: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

164 Anthony M. Dean, Joseph W. Bozzelli

HNNO to obtain the total for this channel. The high temperature predictions,above 1000 K, are in good agreement with the data for the N2 + OH channel(cf. Fig. 14). This is particularly encouraging, since no parameters were adjustedin the calculation. However, this prediction does not account for the observedrate coefficient at lower temperatures. This strongly non-Arrhenius temperaturedependence has been attributed to tunneling by Marshall et al. (1987). They pointout that tunneling is consistent with the observed isotope effect. Although theseresearchers rule out stabilization as a major pathway on the basis of their observedlack of pressure dependence, our calculations suggest that stabilization may wellaccount for the lower temperature data.

In Fig. 17 we compare our predictions for all channels to the data of Marshallet al. The calculations were done for a pressure of 0.5 atm and Ar buffer gas,conditions typical of those used in the experiments. Below 900 K, stabilizationof the HNNO adduct is predicted to be important. (As discussed above, the lowA-factor for dissociation of this adduct back to reactants makes this channelparticularly susceptible to stabilization.) The temperature range over whichstabilization dominates coincides with the region of significant non-Arrheniusbehavior seen in the experiments of Marshall et al. and might account for theunusual behavior. The strong temperature dependence of the data, as well as thelimited number of experiments at constant temperature and various pressures byMarshall et al., might obscure the predicted dependence on pressure. Additionalexperiments are needed to identify the source of the higher observed rate at lowertemperatures. In the interim, we cautiously suggest use of a calculated stabilizationrate coefficient. At 1 atm N2

k11a = 1.3× 1025T−4.48 exp(−5420/T ) cm3mol−1s−1

for the temperature range from 300 to 1200 K. Over the 600 to 1200 K temperatureinterval the rate coefficient for this channel is close to the low pressure limit, sothat it can be scaled linearly with pressure.

Given the ambiguity concerning the lower temperature behavior, we are reluc-tant to conclude that all of the observed loss of reactant in this regime can beattributed to production of N2 + OH. Above 1000 K, however, it appears that themeasurements of Marshall et al. describe this reaction well. (Cf. Fig. 17.) Thuswe recommend use of the high-temperature component of their rate expression

k11b = 2.2× 1014 exp(−8430/T ) cm3mol−1s−1

for the temperature range from 1000 to 2000 K. (We have extended the hightemperature range since this expression is consistent with the shock tube data.)The QRRK analysis suggests there should be no effect of pressure on the ratecoefficient for this channel at higher temperatures.

Page 41: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 165

7

9

11

13

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

) Marshall et al. (1989)

HNNO

NH + NO

O + NNH

OH + N2 (total)

P = 0.5 atm Ar

H + N2O → Products

FIGURE 2.17. Arrhenius plot for the various channels of H + N2O −→ Productsas predicted by QRRK calculations for 0.5 atm Ar. Note the importance of the HNNOstabilization channel at lower temperatures. Near 2500 K the NH + NO channel has grownto become almost as important as the OH + N2 channel.

2.6.5 H + N2O −→ NH + NO

The predicted rate coefficient for the NH + NO channel is compared to experi-mental measurements in Fig. 15. We predict the rate coefficient for this channelto be independent of pressure over the range 10−3 to 100 atm. The calculationsindicate that at higher temperatures the NH + NO channel becomes increasinglyimportant and is comparable to the N2 + OH channel near 2500 K (cf. Fig. 17).Selection of a preferred rate coefficient is difficult. The data of Mertens et al. were

Page 42: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

166 Anthony M. Dean, Joseph W. Bozzelli

obtained from measurements of the reverse reaction rate; we suspect ambiguity inassigning product channels (see below), which would lead to a lower activationenergy for k11c and bring that expression closer to the one obtained from detailedmodeling (of the reverse reaction) by Miller and Bowman as well as that calculatedin this work. Our analysis leads to rate coefficients close to those given by theMiller-Bowman expression. We suggest using our calculated expression:

k11c = 8.5× 1020T−1.62 exp(−17800/T ) cm3mol−1s−1

for the temperature range from 300 to 4000 K.

2.6.6 H + N2O −→ NNH + O

Very little information is available on the NNH + O channel. The results inFig. 17 suggest it could become important at higher temperatures. Its smaller ratecoefficient is primarily due to a higher barrier. Our results can be described by

k11d = 2.4× 1019T−1.26 exp(−23700/T ) cm3mol−1s−1

for the temperature range from 300 to 4000 K. This expression is much lowerthan one obtains using the Miller-Bowman estimate for the reverse reactionrate coefficient. The calculations indicate that this channel has no pressuredependence.

2.6.7 NH + NO −→ Products

The potential energy surface shown in Fig. 13, with the same QRRK parametervalues used for the H + N2O calculations, was also used to study the reaction NH+ NO −→ Products. The results (Fig. 16) show that the features reported inthe experiments, including the increase in rate at high temperature, are accountedfor.

Predictions for the various channels are shown in Fig. 18. Stabilization channelsare predicted to be unimportant even at the lowest temperatures. The major productchannel, except at the highest temperatures, is predicted to be H + N2O, and ourpredicted rate coefficient expression agrees well with the calculations of Millerand Melius. Our analysis leads to a somewhat less important N2 + OH productchannel than predicted by Miller and Melius. At 3000 K, we predict that k11e/ktot= 9%, as compared to 19 � 10% measured by Mertens et al.; Miller and Meliuspredict 29%. At room temperature, we predict 9%, compared to the Lillich etal. measurement of 15 � 5% and the Miller-Melius prediction of 19%. We notethat the reports of N2 + OH as the dominant channel at lower temperatures wouldsuggest that the barrier to the HNNO⇀↽ NNOH isomerization used in both thiswork and by Miller and Melius is too high. While a detailed evaluation of theeffect of this barrier on the product distribution is beyond the scope of this chapter,we note that changes of the magnitude needed would require a decrease in the

Page 43: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 167

11

12

13

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

OH + N2

H + N2O

O + NNH

total

NH + NO → Products

FIGURE 2.18. Arrhenius plot for the various channels of NH + NO −→ Products aspredicted by QRRK calculations for 1 atm N2. The change in temperature dependencenear 2000 K is due to the onset of the O + NNH channel; at lower temperatures H + N2Ois predicted to be dominant. As discussed in the text, the isomerization barrier betweenHNNO and NNOH would need to be significantly lower to have N2 + OH dominate atlower temperature.

barrier height well beyond the expected uncertainty in the Melius calculation.Such a conclusion is consistent with the data supporting that the N2O channel isdominant at room temperature. It is also reinforced by the theoretical analysis ofBradley et al. (1995), who report a branching fraction of 13±3% for the N2 + OHchannel and show that decreasing the isomerization barrier by 40 kJ/mol increasesthe branching fraction by less than a factor of 2.

An additional feature of interest in Fig. 18 is the predicted importance of theNNH + O channel at the highest temperatures. The results suggest that NNH +

Page 44: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

168 Anthony M. Dean, Joseph W. Bozzelli

O is a major product in the Mertens et al. shock tube experiments. This channelis more important for the NH + NO reaction than the H + N2O reaction since theavailable energy is appreciably higher for the NH + NO reaction, thus allowingeasier access to the endothermic NNH + O channel. This seems to be the mostplausible explanation for the observed temperature dependence in the shock tubeexperiments. This reaction should be included in high temperature models; weshow in Section 7.3 that it can have significant implications for NO production.

None of these channels are predicted to depend on pressure. Given the amountof energy in the adducts, as well as the low-energy dissociation channels, thisbehavior is expected—the unimolecular dissociation events are much too rapid tobe intercepted by collisions. Our predicted rate coefficients are given by

k11e = 1.4× 1017T−1.49 exp(−660/T ) cm3mol−1s−1

k11f = 3.0× 1018T−1.65 exp(−720/T ) cm3mol−1s−1

k11g = 1.7× 1014T−0.20 exp(−6140/T ) cm3mol−1s−1

for the temperature range from 300 to 4000 K.1

2.6.8 NH + O2 −→ Products

Most of the available high temperature rate coefficients for this reaction have beeninferred from applications of kinetic models to flames (Miller et al. 1983; Deanet al. 1984; Bian et al. 1990; Tsang et al. 1991; Vandooren et al. 1991). Thetotal rate coefficient was measured in a direct experiment behind shock waves(Mertens et al. 1991a) and a direct measurement at lower temperatures was madeby Hack et al. (1985b), who concluded that OH + NO is the dominant pathwaybelow 573 K. Baulch et al. (1992) recommend the Hack expression for the lowertemperature regime. Later data (Hennig et al. (1993); Lillich et al. (1994)) in theintermediate temperature regime capture the marked curvature in the Arrheniusplot. Miller and Melius (1992a) calculated rate coefficients for two channels over awide temperature range, showing that HNO + O is the dominant high temperaturepathway while OH and NO are the major products at low temperatures. Theexperimental data are summarized in Fig. 19. The results inferred from themodeling studies are notably higher than the direct observations of Mertens et al.

A potential energy diagram for this system, based on the calculations of Millerand Melius (1992a), is shown in Fig. 20.2 Input parameters for the QRRK

1 The expression for k11f is included only for illustration; since reaction (11f) isthe reverse of reaction (11c), it should not be added to a reaction mechanism thatalready includes reaction (11c).2 For completeness, the additional pathway of HONO dissociation to H + NO2shown by the dashed line in Fig. 20 was considered; it is not expected to competewith direct dissociation from HNO2, because the A-factor for isomerization toHONO is appreciably smaller than the A-factor for direct dissociation.

Page 45: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 169

10

11

12

13

1 2 3 41000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Dean at al. (1984) HNO+O

Hack et al. (1985) NO+OH

Mertens et al. (1991) total

Bian et al. (1990) NO+OH

Vandooren et al. (1991) NO+OH

Hennig et al. (1993) total

Lillich et al. (1994) total

NH + O2 → Products

FIGURE 2.19. Arrhenius plot for NH + O2 −→ Products. The direct measurementsof Mertens et al. (1991) and Hennig et al. (1993) suggest that the higher rate coefficientvalues derived from more complex systems are in error. The NO + OH channel appears todominate at lower temperatures.

–100

3NH + O2

1HNOO

1HNO2 1HONO

2OH + 2NO

2H + 2NO2

3O + 1HNO3HNOO

Ent

halp

y (k

J/m

ol)

0

100

200

400

300

FIGURE 2.20. Potential energy diagram for NH + O2. The dashed line indicates a reactionpathway not considered by Miller and Melius (1992a).

Page 46: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

170 Anthony M. Dean, Joseph W. Bozzelli

calculations are listed in Table 7, and the results are compared to literatureexpressions in Fig. 21. The low A-factor for HONO formation suppresses OH+ NO production, since this is the only path to these products. Although onemight expect a route to OH + NO via an H-shift in HNOO (to form NOOH, whichrapidly dissociates), the estimated heat of formation of NOOH places it about 96kJ/mol higher than HNOO; thus the barrier for an H-shift to form NOOH is toohigh for this pathway to be important.

TABLE 2.7. NH + O2 −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) 3NH + O2 −→ 1HNOO 2.0×1012 0(–1) 1HNOO −→ 3NH + O2 2.0×1014 112

(2) 1HNOO −→HNO + O spin forbidden; cf. triplet input(3) 1HNOO −→ 1HNO2 4.9×1012 120

(–3) 1HNO2 −→HNOO 4.9×1013 418(4) 1HNO2 −→H + NO2 5.1×1014 319(5) 1HNO2 −→HNO + O spin forbidden; cf. triplet input(6) 1HNO2 −→ 1HONO 7.7×1013 205

(–6) 1HONO −→HNO2 4.8×1012 221(7) 1HONO −→H + NO2 3.0×1014 327(8) 1HONO −→OH + NO 8.5×1015 203(9) 3NH + O2 −→ 3HNOO 6.0×1012 21

(–9) 3HNOO −→NH + O2 2.2×1014 2(10) 3HNOO −→HNO + O 5.0×1014 28

Molecular constants of 1HNOO adduct: ν1 = 798 cm{1 (degen=1.83) ν2 = 801 cm{1

(1.85) ν3 = 1838 cm{1 (2.33); molecular constants of 1HNO2 adduct: ν1 = 1099 cm{1

(degen=1.97) ν2 = 1107 cm{1 (1.91) ν3 = 2553 cm{1 (2.12); molecular constants of1HONO: ν1 = 548 cm{1 (degen=1.99) ν2 = 1219 cm{1 (2.86) ν3 = 3450 cm{1 (1.15); σ =3.49 A and ε/K = 350 for all adducts.

Rate coefficients: k1 from 3CH2 + O2, Darwin et al. (1989) and Bohland et al. (1984); k−1by microscopic reversibility (MR); k3 TST, loss of 1 rotor, N–H stretch, gain of N–HO andO–OH bends, Ea(3) set to be 8 kJ/mol above entrance channel (from Miller and Melius1992), (variation of barrier height was checked, see text); k−3 MR; k4 via k−4 and MR withk−4 =1.3×1013 exp(–1100/T ) based on H + C2H4; k6 TST, Ea(6) from Miller and Melius(1992), degeneracy = 2; k−6 MR; k7 via k−7 and MR with k−7 = 1.3×1014exp(-180/T )(Ko et al. 1991)—included for completeness (not considered by Miller and Melius); k8via k−8 and MR with k−8 = 1.9×1013, Atkinson et al. (1992); k9 3×k1 because of tripletsurface, barrier estimated; k−9 MR; k10 A10 via MR using A−10 = 2.0×1013 for O +CH3O (Herron 1988) and barrier from the adjusted barrier of Miller and Melius (1992).

Page 47: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 171

10

11

12

1 2 3 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Hack et al. (1985) NO + OH Mertens et al. (1991) total Hennig et al. (1993) total Lillich et al. (1994) total Miller-Melius (1992) HNO + O Miller-Melius (1992) NO + OH QRRK NO+OH, 115 kJ QRRK NO+OH, 120 kJ QRRK HNO + O

NH + O2 → Products

FIGURE 2.21. Comparison of QRRK predictions to measurements and calculations forNH + O2. The QRRK calculations bracket the lower temperature data for the NO + OHchannel. Small changes in the assumed isomerization barrier lead to the large changes in theQRRK rate coefficients, as generally found when the isomerization barrier is comparableto the entrance channel energy.

Possible product channels include

3NH+ O2 −→ 1HNOO (12a)

−→ 1HNO2 (12b)

−→ 1HONO (12c)

−→ OH+ NO (12d)

−→ H+ NO2 (12e)

−→ 3HNOO (12f)

−→ HNO+O . (12g)

Page 48: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

172 Anthony M. Dean, Joseph W. Bozzelli

Fig. 21 shows that the QRRK prediction for production of O + HNO on thetriplet surface agrees with the calculations of Miller and Melius and the data ofMertens et al. The dominant singlet channel produces OH + NO, in accord withthe observations, but the predicted rate coefficient is below the data. Becausethe barrier height for isomerization of HNOO to HNO2 is comparable to the welldepth, however, the QRRK results are extremely sensitive to its value. A decreaseof only 5 kJ/mol is sufficient to bring calculations and data into agreement.

7

9

11

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

1HNOO

H + NO2 via HNO2

OH + NO

H + NO2 via HONO

HNO + O

QRRK total

P = 1 atm N2

NH + O2 → Products

FIGURE 2.22. QRRK predictions for NH + O2−→ Products at 1 atm N2. Stabilizationis predicted to be important at 1 atm and lower temperatures, while the HNO + O channeldominates at high temperatures.

The QRRK results can be used to explore the effects of temperature and pressureon the rate coefficient of this reaction. Fig. 22 summarizes the predicted effectof temperature at 1 atm N2. Near room temperature, substantial production of

Page 49: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 173

the 1HNOO adduct is predicted. This channel would be two to three orders ofmagnitude slower in the low-pressure experiments of Hack et al., making theOH + NO channel dominant, as observed. Production of 1NHOO falls off withincreasing temperature, as typically observed for radical addition reactions withoutlow energy exit channels. In addition to formation of OH + NO, one expects thatsmaller amounts of H + NO2 would be produced. As expected, the production ofH + NO2 via HONO is much less than via HNO2, for the reasons discussed above.At higher temperatures, production of HNO + O via the triplet surface dominates.Fig. 23 shows that most of the reaction channels are independent of pressure andthat at 1500 K stabilization is unimportant up to pressures of 100 atm.

6

8

10

12

–3 –2 –1 0 1 2 log (P/atm)

log

( k

/ cm

3 m

ol–1

s–1

)

1HNOO

H + NO2 via HNO2

OH + NO

H + NO2 via HONO

O + HNO

total

T = 1500 K

NH + O2 → Products

FIGURE 2.23. QRRK predictions of the effect of pressure at 1500 K for NH + O2−→Products. Rate coefficients for all major channels are independent of pressure, so stabiliza-tion should be unimportant for combustion applications.

Page 50: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

174 Anthony M. Dean, Joseph W. Bozzelli

The predicted absence of a pressure effect for the major channels at hightemperatures allows one to use the literature expressions for rate coefficients forvirtually all conditions of interest in combustion. We suggest use of the Hack etal. expression for the OH + NO channel

k12d = 7.6× 1010 exp(−770/T ) cm3mol−1s−1

and the Miller-Melius recommendation for the HNO + O channel

k12g = 4.6× 105T 2 exp(−3270/T ) cm3mol−1s−1 .

For completeness, we suggest use of the QRRK estimate for stabilization of the1HNOO adduct

k12a = 3.7× 1024T−5 exp(−1155/T ) cm3mol−1s−1 ,

which is valid for N2 at 1 atm. One can readily scale this expression to otherpressures—since the calculations show this channel is in the low pressure limit,its rate coefficient is directly proportional to pressure. The QRRK estimate for theH + NO2 channel is

k12e = 2.3× 1010 exp(−1250/T ) cm3mol−1s−1 .

The QRRK calculations show that the other channels are not important.

2.6.9 NH2 + O2 −→ Products

The NH2 + O2 reaction has received a considerable amount of attention and hasbeen included in several mechanisms of ammonia oxidation. It is known to berelatively slow in comparison to other reactions of NH2, with some authors (Meliusand Binkley 1984a) favoring an endothermic pathway to NH2O + O, while othersmodeling ammonia oxidation and Thermal DeNOx (Fujii et al. 1981; Branchet al. 1982; Dean et al. 1982; Kimball-Linne and Hanson 1986) suggest thatthe products are HNO + OH formed by isomerization and decomposition of theNH2O2 adduct. (Melius and Binkley concluded that the adduct did not have apotential well sufficiently deep to allow a rapid isomerization path.) Miller andBowman (1989) followed earlier modeling studies in assuming that the productsare HNO + OH. The literature was sufficiently ambiguous in 1984 for Hansonand Salimian to be unable to make a recommendation. Since then there have beenadditional experimental studies as well as a detailed QRRK analysis (Bozzelli andDean 1989).

The experimental studies included a discordant variety of low temperaturemeasurements. Lesclaux (1984), Patrick and Golden (1984), Michael et al.(1985), and Tyndall et al. (1991) all reported a slow rate, with only upper limitsmeasured, while Hack and Kurzke (1985a) reported rate coefficients some two

Page 51: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

FIGURE 2.24. Potential energy diagram for NH2 + O2. The high barrier for isomerizationof NH2OO to HNOOH, coupled with the relatively low A–factor, makes the HNO + OHpathway unlikely. The additional barrier for H2O elimination makes the NO + H2O channeleven less likely.

orders of magnitude higher. Baulch et al. (1992) recommend the lower valueat room temperature. Hennig et al. (1995) measured NH2 decay as well as OHand O production behind incident shock waves in N2H4–O2–Ar mixtures between1450 and 2300 K. They report that the radical formation channels contribute lessthan 15% of the overall reaction, and that the main reaction products are NO andH2O.

Fig. 24 shows a potential energy diagram for this reaction. There are fivepossible channels:

NH2 +O2 −→ NH2OO (13a)

−→ NH2O+O (13b)

−→ HNOOH (13c)

−→ HNO+OH (13d)

−→ NO+H2O (13e)

The combination of an extremely shallow well for the NH2OO adduct and thehigh barrier (171 kJ/mol) calculated by Melius and Binkley for the intramolecularhydrogen shift implies that the energy required for simple bond fission to formNH2O + O (151 kJ/mol) is actually lower than that required for isomerization.This isomerization barrier is consistent with what one would estimate assumingthat the barrier height is the sum of the energy for hydrogen abstraction and thering strain energy in the cyclic transition state. This assumption suggests a barrier

Page 52: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

176 Anthony M. Dean, Joseph W. Bozzelli

of about 54 + 109 = 163 kJ/mol. The lower barrier for bond fission, coupledwith the higher A-factor for this channel, would suggest that the fission channeldominates. Similarly, one expects formation of the molecular products NO + H2Oto be unlikely since any HNOOH formed would decompose easily to HNO + OH.The shallow wells for both adducts suggest that stabilization is unlikely, especiallyat high temperatures.

We repeated our earlier QRRK analysis (Bozzelli and Dean 1989) using updatedthermodynamic parameters and barrier estimates. The results were similar to ourearlier ones, suggesting that the reaction is in the low pressure limit for pressuresbelow 10 atm and that the major channel for all conditions below 1000 K isstabilization, a non-reaction because dissociation of the NH2OO adduct, withonly a 25 kJ/mol potential well, is rapid. These results suggest negligible netstabilization, in agreement with most of the reported experimental results. Wecould not reconcile our results with data reported by Hack and Kurzke. The QRRKcalculations suggest instead that the major high temperature channel is NH2O +O, with virtually no effect of pressure. Under typical conditions of temperatureand pressure, one can describe this channel with:

k13b = 2.5× 1011T 0.48 exp(−14900/T) cm3mol−1s−1.

The QRRK rate coefficient for the channel to HNO + OH is:

k13d = 6.2× 107T 1.23 exp(−17700/T) cm3mol−1s−1,

substantially smaller than k13b. The calculated rate coefficient for the NO + H2Ochannel is several orders of magnitude lower than that for HNO + OH. Neitherthe NH2O + O nor the HNO + OH channel is affected by change in bath gas overthe temperature or pressure ranges considered. Furthermore, the stabilization andthe NH2O + O channels are not affected by plausible changes in the barrier toHNOOH. The barrier for the intramolecular H-shift would need to lowered by over40 kJ/mol for the HNO + OH channel to compete with the NH2O + O channel.Such a drastic change seems unwarranted. Thus the QRRK analysis, confirmingthe earlier arguments of Melius and Binkley, seems to rule out dominance of theHNO + OH channel. This conclusion is also consistent with the observations ofHennig et al. that the NH2O + O channel is more important than HNO + OH,although their analysis suggests the difference between the two channels is lessthan we calculate. In any event, considering that HNO + OH was assumed tobe the only channel in earlier ammonia oxidation modeling, e.g., that of Millerand Bowman (1989), it is clear that these ammonia oxidation modeling needs tobe revisited. The above results, as well as those on NH2 + HO2 discussed in thefollowing section, indicate that NH2O should be included in ammonia oxidationmechanisms.

A major unresolved issue is the importance of the NO + H2O channel. TheQRRK analysis suggesting that this channel is unimportant is in conflict withthe interpretation of Hennig et al. that this is the dominant pathway. Their

Page 53: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 177

conclusion was derived from their observation that NH2 decay is fast relative toproduction of H and OH. We cannot offer a satisfactory alternate explanation oftheir observations. However, our analysis of that experimental system suggeststhat it is surprisingly complex. For example, a sensitivity analysis for NH2 decayin their experiments indicates that the most important reaction is NH2 + NH2−→NH3 + NH, not any reactions of NH2 + O2. Dominance of the NO + H2O channelwould have far-reaching consequences for ammonia oxidation, as discussed inSections 8.1 and 8.3. Inclusion of this channel in the mechanism causes substantialchain termination and forces the reaction to proceed more slowly than observed.

2.6.10 NH2 + HO2 −→ Products

The NH2 + HO2 reaction has received relatively little attention. The potentialenergy surface for recombination-elimination (Figure 25) shows three possiblechannels:

NH2 +HO2 −→ NH2OOH (14a)

−→ NH2O+OH (14b)

−→ HNO+H2O (14c)

Another possible pathway is the disproportionation reaction:

NH2 +HO2 −→ NH3 +O2 (14d)

There are no direct experimental studies. Two photochemical studies (Cheskisand Sarkisov 1979; Lesclaux 1984) on the NH3–O2 system led to estimates ofthe overall rate coefficient for this reaction of 3.0 and 1.5 ×1013 cm3mol−1s−1,respectively. One theoretical study (Pouchan et al. 1987) was reported whereinthe principal focus was the isomerization channel (14c); the results indicated thatit is not significant. Lesclaux, however, concluded the opposite, that the HNO +H2O product channel is probably the most important—he neglected the NH2O +OH channel, arguing that it is not energetically accessible. This latter conclusionis surprising, as our thermodynamics (cf. Fig. 25) suggest that these products areappreciably lower in energy than the reactants. Baulch et al. (1992) recommend1.6 ×1013 molcm−3s−1 over the range 300 to 400 K for loss of reactant. Usingthe DHT method of Section 4 to estimate the disproportionation rate coefficientleads to

k14d = 9.2× 105 T 1.94 exp(+580/T ) cm3mol−1s−1

At 300 K this expression gives only about 4 × 1011 cm3mol−1s−1, much slowerthan observed, suggesting that reaction 14d is not the pathway for this reaction atlow temperature.

This system was studied using the QRRK approach by Bozzelli and Dean(1989). The most important feature of the reaction is that the NH2O + OH productchannel is 92 kJ/mol below the lowest initial energy of the NH2OOH adduct.

Page 54: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

178 Anthony M. Dean, Joseph W. Bozzelli

250

HNO + H2O

NH2O + OHE

ntha

lpy

(kJ/

mol

)

NH2 + HO2

NH2OOH

150

50

–50

–150

FIGURE 2.25. Potential energy diagram for NH2 + HO2. The high barrier for isomer-ization of NH2OOH, coupled with the relatively low A-factor, makes the HNO + H2Opathway unlikely in spite of its greater exothermicity.

While the H2O + HNO products are lower in energy than NH2O + OH, the barrierto H2O formation is high, as discussed above, so that this reaction is unimportant.

The calculations indicate that production of NH2O + OH is the only importantreaction channel. The rate coefficient for forming these products is nearly constantover the temperature range from 200 to 1900 K and at pressures to above 10 atm.The rate coefficient for these conditions can be described by

k14b = 2.5× 1013 cm3mol−1s−1

The results are insensitive to the thermochemistry used; increasing the heat offormation of NH2O by 42 kJ/mol only increases the contribution of the stabilizationchannel from 0.1 to 1 percent of the collisions at 200 K and 1 atm. The barrier forthe H-atom shift, 65 kJ/mol above the initial reactant energies, combined with thepresence of lower energy dissociation channels (to products and back to reactants),essentially eliminates the HNO + H2O channel from consideration. Anotherreason for this is the lower A-factor expected for the tight transition state for theshift relative to the simple bond fission transition states. Thus the isomerizationchannel can only become favorable if the thermodynamics are changed to lowerthe shift barrier by about 210 kJ/mol, which is inconsistent with other 4 memberring transition states.

Our calculations are consistent with the limited experimental measurementsand provide an explanation for high rate coefficient, i.e., a large A-factor andzero barrier is expected for radical–radical recombination. However, we predict acompletely different product set on the basis of the low energy exit channel. The

Page 55: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 179

250

NH2 + O

NH2OHNOH

NO + H2

HNO + H

NH + OH

Ent

halp

y (k

J/m

ol)

50

150

350

450

FIGURE 2.26. Potential energy diagram for NH2 + O. The dashed lines illustrateadjustments to barrier heights that affect the branching ratio of products.

largest error in determination of this rate coefficient is estimating the high pressurelimit A-factor for the NH2 + HO2 channel. The expression we chose for it, 2.5×1013 cm3mol−1s−1, gives rate coefficients approximately halfway between theexperimental results of Lesclaux and of Cheskis and Sarkisov.

2.6.11 NH2 + O −→ Products

The NH2 + O reaction system has been analyzed theoretically by Melius andBinkley (1984a), who calculated energies for the various possible molecularadducts (NH2O and HNOH) and for the barriers leading to the various dissociationchannels. Their potential energy surface is shown in Fig. 26. The followingchannels are possible:

NH2 + O −→ NH2O (15a)

−→ HNOH (15b)

−→ HNO+H (15c)

−→ NH+OH (15d)

−→ NO+H2 (15e)

Sufficient energy is available in the initial NH2O adduct that it can readilydissociate to HNO + H or unimolecularly isomerize by an H-atom shift anddissociate to NH + OH or undergo concerted H2 elimination to form NO + H2.

Rate coefficients for channels (15c) and (15d) have been measured at roomtemperature by Dransfeld et al. (1984) using a flow reactor; NH and OH were

Page 56: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

180 Anthony M. Dean, Joseph W. Bozzelli

monitored by laser magnetic resonance and HNO by laser induced fluores-cence. The rate coefficient for the overall reaction was reported to be 5.3×1013

cm3mol−1s−1, with 87% of the reaction found to proceed via the HNO + Hchannel, in agreement with the predictions of Melius and Binkley. The productinternal energy distribution was probed in a molecular beam study by Patel-Misraand Dagdigian (1991). The observed low level of rotational excitation of HNOwas taken as evidence of a barrier to formation of HNO from the adduct. TheCohen-Westberg review (1991) recommends the Dransfeld et al. expression forthe total rate coefficient and the 87% branching fraction to HNO + H. They includeNO + H2 as a minor channel with a rate coefficient slightly smaller than for theNH + OH channel.

This reaction was also studied by Bozzelli and Dean (1989). The potential en-ergy surface suggests that stabilization should be unimportant at typical pressures,since low energy exit channels are available. The QRRK results are consistentwith this expectation; stabilization is calculated to contribute less than 1% atpressures below 10 atm. The QRRK analysis predicted that the HNO + H channeldominates, with the HNOH isomer contributing approximately twice as much asthe NH2O isomer at low temperatures. The HNOH contribution shifts slightly(about 20%) toward the NH + OH channel at higher temperatures because of thehigher A-factor for dissociation of HNOH to NH + OH and the higher barrier tothis channel. Using the barrier of Melius and Binkley for formation of HNO +H from HNOH dissociation (the dashed curve in Fig. 26), the NH + OH channelwas found to account for about 30% of the total reaction products, higher thanobserved by Dransfeld et al. The Melius and Binkely barrier for the HNO + Hchannel, however, implies a 42 kJ/mol barrier for the reverse reaction of H–atomaddition to HNO, which is higher than most H–atom additions to unsaturated bondsystems. In H + C2H4, for example, the observed barrier is in the 4 to 12 kJ/molrange (Kerr and Moss 1981) and reactions of H with NO and HCO have evenlower barriers. Upon lowering the barrier to the HNO + H channel so that thereverse reaction had a barrier of 15 kJ/mol (the solid curve in Fig. 26), the QRRKanalysis gave the branching ratio measured by Dransfeld et al.

Bozzelli and Dean (1989) also pointed out other barrier adjustments that wouldgive results consistent with the observations. In spite of these uncertaintiesconcerning the potential energy surface the qualitative picture is clear. The overallreaction is fast and the room temperature value of Dransfeld et al. is also applicableat higher temperatures. Thus we recommend

k15(total) = 5.3× 1013 cm3mol−1s−1.

The measurements at 295 K by Adamson et al. (1994), who found k15(total) tobe 3.9 × 1013 cm3mol−1s−1, with about 5 to 8% branching into the NH + OHchannel, are in reasonable agreement with the Dransfeld et al. measurements. Thebranching ratio would be expected to have only a modest temperature dependenceand (15c) to be the dominant channel at all temperatures. We recommend use of

Page 57: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 181

the Dransfeld et al. values, i.e.,

k15c = 4.6× 1013 cm3mol−1s−1

k15d = 7× 1012 cm3mol−1s−1.

In spite of the fact that the NO + H2 channel is very exothermic, the high barrierfor concerted elimination, coupled with the correspondingly low A-factor, makesthis channel unimportant.

The potential energy surface computed by Wolf et al. (1994), who distinguishedbetween cis and trans-HNOH, agrees in most of its features with the one computedby Melius and Binkley.

Direct abstraction from NH2 by O would also produce NH + OH. Using theDHT method described in Section 4 leads to

k15d2 = 3.3× 108 T 1.5 exp(−2555/T ) cm3mol−1s−1.

The contribution of this channel is negligible at low temperatures, but exceedsthat for the addition pathway at 2000 K. A similar conclusion was reached byDuan and Page (1995) on the basis of theoretical calculations for the abstractionchannel. Their expression

k = 0.87T 4.01 exp(842/T ) cm3mol−1s−1

is within a factor of 2 of our expression from 300 to 2500 K.

2.6.12 NH2 + OH −→ Products

Little is known experimentally about this reaction. The expressions of Hanson andSalimian (1984) and Miller and Bowman (1989) are compared to earlier work ofCheskis and Sarkisov (1979) and an upper limit measurement by Diau et al. (1990)in Fig. 27. Although all measurements cluster near 1×1013 cm3mol−1s−1, there isdisagreement about the temperature dependence,suggesting that multiple channelsmay be involved.

As with the H + NH2 reaction considered earlier, this reaction could be eitherdirect hydrogen transfer or an addition-elimination reaction. For the direct route,the expected lowest energy products would be 3NH + H2O, a channel that is spin-forbidden for the addition-elimination route via a singlet adduct. The potentialenergy diagram is shown in Fig. 28. Possible product channels for the addition-elimination route include

NH2 +OH −→ NH2OH (16a)

−→ 3NH+ H2O (16b)

−→ 1NH+ H2O (16c)

−→ NH2O+H (16d)

−→ HNOH+H (16e)

Page 58: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

182 Anthony M. Dean, Joseph W. Bozzelli

11

12

13

1 2 3 1000 K / T

log

( k

/ cm

3 mol

–1 s

–1)

Hanson-Salimian (1984) NH + H2O Miller-Bowman (1989) NH + H2O Diau et al. (1990) upper limit Cheskis-Sarkisov (1979) NH + H2O QRRK total at 1 atm N2 QRRK total at 300 Torr He abstraction estimate

NH2 + OH → Products

FIGURE 2.27. Comparison of QRRK predictions to rate coefficient measurements forNH2 + OH. Also included is our estimate for the rate coefficient of the abstraction pathwayto form NH + H2O. The results suggest a change in mechanism from adduct formation atlow temperature to hydrogen abstraction at high temperature.

The input parameters for the analysis are shown in Table 8. If one assumesthe barrier to the spin-forbidden lowest energy products (3NH + H2O) is about255 kJ/mol, one obtains the total rate coefficient predictions for the addition-elimination pathway shown in Fig. 27. (Use of a lower barrier results in evenhigher total rate coefficients at the lower temperatures, clearly higher than theupper limit reported by Diau et al. in 300 Torr helium.) The QRRK predictions athigher temperatures are considerably lower than the literature values.

Our calculations suggest that the dominant low-temperature pathway is forma-tion of the NH2OH adduct, while 3NH and H2O are the high-temperature products.

Page 59: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 183

NH2 + OH

NH2OH

3NH + H2O

H + NH2O

H + HNOH

Ent

halp

y (k

J/m

ol)

–50

50

150

250

350

1NH + H2O

FIGURE 2.28. Potential energy diagram for NH2 + OH. The barrier assignment for3NH + H2O was made to achieve consistency with the lower temperature measurements ofDiau et al. (1990).

Above 1000 K adduct formation is at the low pressure limit to at least 10 atm. TheQRRK expressions for adduct formation in N2 are

k16a = 1.8× 1032 T−6.91 exp(−2070/T ) cm3mol−1s−1 0.1 atm

= 3.9× 1033 T−7.00 exp(−2235/T ) cm3mol−1s−1 1.0 atm

= 5.6× 1034 T−7.02 exp(−2700/T ) cm3mol−1s−1 10 atm.

Our conclusion that adduct formation dominates at lower temperatures differsfrom that of Cheskis and Sarkisov, who report NH and H2O to be the products.

We estimated the rate coefficient for direct hydrogen transfer using the DHTprocedure described in Section 4 to be

k16b = 2.4× 106 T 2 exp(−25/T ) cm3mol−1s−1.

This result, which is close to the expression used by Miller and Bowman (1989),has been included in Fig. 27.

Our analysis suggests that the high-temperature reaction is direct H-atomtransfer; thus no pressure dependence of the rate coefficient is expected at hightemperatures, as concluded in Section 6.1 for the H + NH2 reaction.

Page 60: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

FIGURE 2.29. Potential energy diagram for the molecular and radical output channels ofNH2 + NH2.

Page 61: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 185

2.6.13 NH2 + NH2 −→ Products

There are limited high temperature measurements relevant to this reaction (David-son et al. 1990b; Dean et al. 1984). Low temperature measurements by Stothardet al. (1995) indicate that H2 production dominates. Because there must be a lowenergy barrier for this channel, formation of H2 and singlet H2NN are indicated;formation of H2 and N2H2 would be expected to have a much higher barrier, asshown Fig. 29.

Possible product channels include

NH2 +NH2 −→ N2H4 (17a)

−→ H2NN+ H2 (17b)

−→ N2H2 + H2 (17c)

−→ N2H3 + H (17d)

Davidson et al. derived a rate coefficient from a model used to describe theirshock tube measurements of ammonia pyrolysis and proposed

NH2 +NH2 −→ NH3 +NH. (17e)

Dean et al. proposed that the recombination reaction has two sets of products,N2H3 + H and N2H4, on the basis of radical profiles observed in rich ammoniaflames. Miller and Bowman (1989) suggested that the high temperature productsare N2H2 + H2. These results are summarized in Fig. 30. With the notableexception of Stothard et al. (1995) most of the lower-temperature studies reportthat the product is N2H4, but the reported rate coefficients vary widely (Mallardet al. 1993).

TABLE 2.9. NH2 + NH2 −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) NH2 + NH2 −→ NH2NH2 3.1×1013 0(–1) NH2NH2 −→ NH2 + NH2 1.5×1016 279(2) NH2NH2 −→ N2H3 + H 4.3×1014 321(3) NH2NH2 −→ N2H2 + H2 8.7×1012 414(4) NH2NH2 −→ H2NN + H2 1.5×1012 238

Molecular constants of NH2NH2 adduct: ν1 = 653 cm{1 (degen = 3.88) ν2 = 1315 cm{1

(4.52) ν3 = 3495 cm{1 (3.60) σ = 4.2 A and ε/K = 205.

Rate coefficients: k1 Back and Yokada (1973) and Sarkisov et al. (1984); k−1 microscopicreversibility (MR); k2 via k−2 and MR with k−2 = 1.0×1014 from H + C2H3 of Duran etal. (1988); k3 estimated using Ea = ∆r H◦ + 300 kJ/mol, based upon C2H4 −→ C2H2 +H2 1,2 elimination, A -factor from TST, loss of 1 rotor, degeneracy = 4; k4 A -factor fromLaufer et al. (1983) for H2C2 + H2, Ea adjusted to fit low pressure data of Stothard et al.(1995).

Page 62: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

186 Anthony M. Dean, Joseph W. Bozzelli

This reaction was analyzed using the QRRK method. The input parameters arelisted in Table 9 and the results are compared to the reported rate coefficients inFig. 30. The barrier for channel 17b was adjusted to 238 kJ/mol so as to agreewith the low-temperature, low-pressure data of Stothard et al. (1995), which isconsistent with the reverse reaction, a singlet radical inserting into H2, havinga low barrier. The calculations indicate that the dominant product channel forthe conditions of the Stothard et al. experiments is indeed 17b, for about 80%of the total rate, consistent with their finding that H2 is the major product. Thepressure of the experiments (about 1 mbar) was too low for N2H4 stabilization tobe important.

Several conclusions emerge from these comparisons. The channels that arisefrom adduct formation, i.e., 17a–d, are predicted to have rate coefficients appre-ciably lower than the observations of Davidson et al. However, the rate coefficientthey reported is consistent with a hydrogen atom abstraction reaction. An estimatefor this rate coefficient obtained with the DHT method of Section 4 is seen in Fig. 30to be similar in magnitude and temperature dependence to their measurements.It is also consistent with the rate coefficient for the reverse reaction reported byR�ohrig et al. (1994). We recommend that the Davidson et al. expression

k17e = 5.0× 1013 exp(−5000/T ) cm3mol−1s−1

be used for this rate coefficient.Our calculations also suggest that recombination to produce hydrazine (17a)

is important up to quite high temperatures and should be explicitly consideredin kinetic models. Figure 31 shows the predicted rate coefficients at 1500 K inN2 as functions of pressure. We suggest the following expressions be used fortemperatures from 600 to 2500 K at the specified pressures with N2 as the bathgas:

k17a = 2.0× 1046T−10.93 exp(−5030/T ) cm3mol−1s−1 0.1 atm

= 5.6× 1048T−11.30 exp(−5980/T ) cm3mol−1s−1 1.0 atm

= 3.2× 1049T−11.18 exp(−7040/T ) cm3mol−1s−1 10 atm.

These rate coefficient expressions cannot be extrapolated to room temperature.Our QRRK calculations at room temperature are in good agreement with theresults of Patrick and Golden (1984). The next most important channel at lowertemperatures is 17b, which implies that one has to account for the subsequentreactions of singlet H2NN. This channel is also significant at lower pressures andhigher temperatures. (Figure 31) QRRK calculations for N2 bath gas give

k17b = 2.4× 1020T−2.91 exp(−1075/T ) cm3mol−1s−1 0.1 atm

= 1.2× 1021T−3.08 exp(−1695/T ) cm3mol−1s−1 1.0 atm

= 2.3× 1019T−2.54 exp(−2105/T ) cm3mol−1s−1 10 atm.

At higher temperatures production of N2H3 + H also contributes. Because theN2H3 product will decompose rapidly to produce a second H atom, this channel

Page 63: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 187

8

10

12

1.0 2.0 3.01000 K / T

log

( k

/ cm

3 mol

–1 s

–1)

Davidson et al. (1990) NH + NH3 Dean et al. (1984) N2H3 + H Dean et al. (1984) N2H4 Miller-Bowman (1989) N2H2 + H2 QRRK N2H4QRRK N2H3 + H QRRK H2NN + H2abstraction estimate NH + NH3

P = 1 atm N2

NH2 + NH2 → Products

FIGURE 2.30. Comparison of QRRK predictions to high temperature rate coefficientmeasurements for NH2 + NH2. Also included is an estimate for the direct hydrogenabstraction pathway to form NH + NH3.

can become a doubly significant H-atom source. Its rate coefficient is nearlyindependent of pressure. At 1 atm in N2 the QRRK result is

k17d = 1.2× 1012T−0.03 exp(−5075/T ) cm3mol−1s−1.

Figure 31 shows that the molecular channel forming N2H2 and H2 is unimportantfor all conditions, as expected from the high barrier for concerted elimination.Because this channel would have a tight transition state and correspondingly lowA-factor, it is unlikely that the QRRK estimate could be sufficiently in error tomake this channel important.

Page 64: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

188 Anthony M. Dean, Joseph W. Bozzelli

7

9

11

13

–2 0 2log (P / atm

log

( k

/ cm

3 mol

–1 s

–1)

N2H4 N2H2 + H2

H2NN + H2 N2H3 + H

NH + NH3

T = 1500 K

NH2 + NH2 → Products

FIGURE 2.31. Predicted effect of pressure at 1500 K from the QRRK calculations forNH2 + NH2−→ Products. At pressures above 1 atm, stabilization is predicted to beimportant even at high temperatures.

2.6.14 NH2 + NO −→ Products

This pivotal SNCR reaction has been the subject of extensive analysis since theHanson and Salimian review. The product distribution is of critical importanceand has been intensively studied. Most studies have focused on the two channels

NH2 +NO −→ N2 +H2O (18a)

−→ NNH+OH (18b)

Modeling results are particularly sensitive to the branching ratio, since 18ais a chain termination reaction while reaction 18b, a source of OH as well

Page 65: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 189

as H, from NNH dissociation, is a chain branching reaction. There is nowreasonable agreement on the overall rate coefficient (Atakan et al. 1989; Atkinsonet al. 1989; Bulatov et al. 1989; Pagsberg et al. 1991; Baulch et al. 1992; Diauet al. 1994; Wolf et al. 1994), and there is growing evidence that the branchingratio (β = k18b/(k18a + k18b) is smaller than previously believed, particularly atlower temperatures. For example, six recent investigations (Dolson 1986; Hall etal. 1986; Silver and Kolb 1987; Atakan et al. 1989; Bulatov et al. 1989; Pagsberget al. 1991) all report β values in the range from 0.1 to 0.15 at room temperature;Baulch et al. (1992) recommend β = 0.12 at 298 K. The temperature dependenceof β is controversial. Atakan et al. and Bulatov et al. report that it increasesslightly with temperature, while Park and Lin (1996) and Halbgewachs et al.(1996) report stronger dependences. This information is summarized in Fig. 32,where for completeness the recommendation of Atkinson et al. (1989) has alsobeen included. The rate coefficients for the channel producing OH were obtainedby multiplying the total rate coefficient and branching ratio from the individualstudies cited.

In addition to the uncertainty regarding the temperature dependence of β, asecond concern about this reaction is that H atom production has not been observed.Because of the expected short lifetime of NNH produced in reaction 18b, H atomproduction coincident with OH should be observable. No H has been seen,however, and experiments by Unfried et al. (1990) show that this cannot beattributed to reactions of NNH with NO, at least at room temperature. It is possiblethat some of the observed OH might come from sources other than reaction 18b.(Cf. Unfried et al., loc. cit.) Stephens et al. (1993) found the value of β to increasewith temperature but also reported that the total contribution of the two channelsappeared to decrease at higher temperatures, possibly indicating the onset of anadditional reaction channel. However, the modeling study by Diau et al. (1994)suggests that this phenomenon may result from secondary reactions.

A potential energy diagram for this system, based primarily on the calculationsof Melius and Binkley (1984b), is shown in Fig. 33. The low barrier for H–atomshift, through which the initial NH2NO adduct can form trans-HNNOH, suggeststhat stabilization is relatively unimportant in this system. Trans-HNNOH canisomerize to cis-HNNOH or dissociate directly to NNH + OH. A significantfraction of the NNH + OH yield comes from the trans form, with the cis formdecomposing primarily to N2 and H2O. The branching ratio is controlled by thedifferences in both A-factors and barriers for these channels. Walch (1993), usinga more extensive basis set, obtained results similar to those of Melius and Binkley.Diau and Smith (1996) used Walch’s calculation as the basis for a microcanonicalvariational TST analysis of the reaction.

The input parameters for the QRRK calculation are listed in Table 10. Theresults at 0.01 atm, similar to the pressure in most of the experiments, arecompared to experimental observations in Fig. 34. Also shown in Fig. 34 arethe calculated values of Diau and Smith (1996). Much of the difference betweenthe two predictions for the OH channel can be traced to differences in the total

Page 66: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

190 Anthony M. Dean, Joseph W. Bozzelli

11

12

13

1 2 3 41000 K / T

log

( k

/ cm

3 mol

–1 s

–1)

Atakan et al. (1989) totalBulatov et al. (1989) total Atkinson (1989) totalBaulch et al. (1992) totalPagsberg et al. (1991) totalDiau et al. (1994) totalSilver-Kolb (1987) OHHall et al. 1986 OHBulatov et al. (1989) OH Atakan et al. (1989) OHPark-Lin (1996) OH

NH2 + NO → Products

FIGURE 2.32. Arrhenius plot of experimental measurements of the total and OHproduction rate coefficients for NH2 + NO −→ Products.

rate coefficient, as the values computed for the branching ratio as a function oftemperature agree with one another. The agreement with respect to the overallrate coefficient is reasonable, while the calculations overpredict the amount ofOH produced at higher temperatures. A small increase in the standard enthalpyof formation of NNH would improve the high temperature fit but lead to a lowerbranching ratio than observed at room temperature. Nonetheless, this comparisonsuggests that a straightforward QRRK analysis does capture the intrinsic natureof this system. In particular, it supports that the QRRK predictions as functionsof pressure should serve to indicate how the rate coefficients for the variouschannels are affected by pressure. Fig. 35 illustrates the predicted effect ofpressure at 1200 K, a typical temperature for Thermal DeNOx applications.The total rate coefficient for NH2 + NO disappearance is insensitive to pressure

Page 67: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 191

H2O + N2

N2O + H2

NNH + OH

c-HNNOHt-HNNOHNH2NO

NH2 + NO

–250

Ent

halp

y (k

J/m

ol)

–50

150

350

FIGURE 2.33. Potential energy diagram for NH2 + NO.

TABLE 2.10. NH2 + NO −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) NH2 + NO −→ NH2NO 1.0×1013 0(–1) NH2NO −→ NH2 + NO 1.9×1015 194(2) NH2NO −→ H2 + N2O 6.3×1013 315(3) NH2NO −→ HNNOH (trans) 1.5×1013 128

(–3) HNNOH(trans) −→ NH2NO 6.0×1012 122(4) HNNOH(trans) −→ NNH + OH 5.0×1015 193(5) HNNOH(trans) −→ HNNOH(cis) 1.3×1014 151

(–5) HNNOH(cis) −→ HNNOH(trans) 1.4×1014 144(6) HNNOH(cis) −→ H2O + N2 5.5×1013 91(7) HNNOH(cis) −→ NNH + OH 5.6×1015 185

Molecular constants of NH2NO adduct: ν1 = 506 cm{1 (degen=2.96) ν2 = 1435 cm{1

(1.43) ν3 = 1457 cm{1 (4.61); molecular constants of HNNOH(cis) adduct: ν1 = 634cm{1 (degen=3.29) ν2 = 802 cm{1 (0.41) ν3 = 1506 cm{1 (5.30); molecular constants ofHNNOH(trans) adduct: ν1 = 789 cm{1 (degen=4.42) ν2 = 1213 cm{1 (3.96) ν3 = 2914cm{1 (0.62); σ = 3.97 A and ε/K = 436 for all adducts.

Rate coefficients: k1 review of literature, this study; k−1 microscopic reversibility (MR);k2 nonsymmetric TST, Ea from C.F. Melius—average of 2 TST structures calculated; k3TST, Radicalc estimate (Bozzelli et al. 1993), Ea from Melius et al. (1984b); k−3 MR;k4 via k−4 and MR with k−4 as for OH + CH3, A(–4) = 4.5×1013, Ea(–4) = 0, Washida(1980); k5 TST, Radicalc estimate (Bozzelli et al. 1993) with Ea from C.F. Melius; k−5MR; k6 TST, Radicalc estimate (Bozzelli et al. 1993) with Ea from C.F. Melius; k7 fromk−7 by MR—see k4.

Page 68: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

192 Anthony M. Dean, Joseph W. Bozzelli

12

13

1 2 31000 K / T

log

( k

/ cm

3 mol

–1 s

–1)

Atakan et al. (1989) total Park-Lin (1996) total QRRK total Diau-Smith (1996) total Atakan et al. (1989) OH Park-Lin (1996) OH QRRK OH Diau-Smith (1996) OH

P = 0.01 atm N2

NH2 + NO → Products

FIGURE 2.34. Comparison of QRRK predictions to the measurements of Atakan et al.(1989) for NH2 + NO.

over a wide range. Only at pressures approaching 10 atm is the effect of thestabilization channel noticeable. At room temperature the same transition appearsat pressures closer to 1 atm. Thus one should be able to use the measured lowpressure expressions for the rate coefficients for most applications of interest. Forcompleteness, rate coefficients for formation of NH2NO are included in Table 19.

Our recommendation for the rate coefficient of the OH channel

k18b = 3.5× 1010T 0.335 exp(+385/T ) cm3mol−1s−1

is based on a fit to the average of the measurements of Atakan et al. and Park andLin. The rate coefficient for the N2 channel

k18a = 4.7× 1012T−0.247 exp(+605/T ) cm3mol−1s−1

is a fit to the difference between k18b and the expression for the total ratecoefficient recommended by Baulch et al. (1992) for the temperature interval

Page 69: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 193

7

9

11

13

–2 0 2log (P/atm)

log

( k

/ cm

3 m

ol–1

s–1

)NH2NON2O + H2t-HNNOH NNH + OH via t-HNNOHc-HNNOH NNH + OH via c-HNNOHN2 + H2Ototal

T = 1200 K, M = N2

NH2 + NO → Products

FIGURE 2.35. Predicted effect of pressure at 1200 K from QRRK calculations for NH2+ NO −→ Products. The dominant pathways are virtually independent of pressure untilabove 10 atm.

from 300 to 1000 K. Because of the uncertainty in the temperature dependence ofβ, extrapolation to higher temperatures should be done with caution. At 1250 K,these expressions give β = 0.28, significantly lower than the 1273 K value (0.58)reported by Halbgewachs et al. and the temperature-independent 0.51 value usedby Miller and Bowman (1989) for modeling Thermal DeNOx. As discussed inSection 8.2, such high β values are actually not needed to account for ThermalDeNOx kinetics; a β value near 0.3 at 1250 K is sufficient.

2.6.15 CH3 + NO −→ Products

As CH3 is usually present in much larger concentrations than other hydrocarbonradicals in flames, this reaction may provide an important route for the removal of

Page 70: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

194 Anthony M. Dean, Joseph W. Bozzelli

11

12

13

–2 –1 0 1log (P/atm)

log ( k / cm

3 mol

–1 s

–1)

Davies et al. (1991) 296 K, Ar Wallington et al. (1992) 298 K, SF 6 Vakhtin et al. (1990) 298 K, He Jodkowski et al. (1993) 298 K, acetone Kaiser (1993) 297 K, Ar Davies et al. (1991) 407 K, Ar Davies et al. (1991) 509 K, Ar

CH 3 + NO → Products

FIGURE 2.36. Low temperature measurements of CH3 + NO rate coefficients.

NO (Haynes 1978). There is quite good agreement among low temperaturemeasurements of the recombination channel rate coefficient (Washida 1980;Vakhtin and Petrov 1990; Davies et al. 1991; Wallington et al. 1992; Jodkowski etal. 1993; Kaiser 1993). (Figure 36) The temperature and pressure dependence andthe effect of buffer gases is what one expects for a simple recombination. At highertemperatures, there have been several recent measurements (Wolff and Wagner1988; Hoffmann et al. 1990; Yang et al. 1993; Lifshitz et al. 1993; Hennig and

Page 71: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 195

Wagner 1994). These data and earlier measurements are summarized in Fig. 37.Hoffmann et al. measured production of OH and, using earlier measurementsfor the total rate coefficient, reported that OH production accounts for about 40%of the total methyl radical loss. They also presented evidence that the observedmethyl loss at high temperatures cannot be attributed to formation of the adduct.Lifshitz et al. interpreted their single-pulse shock tube experiments in terms ofHCN + H2O as the major products. Yang et al. measured loss of CH3 andinferred a rate coefficient expression for the total reaction rate. Their expressionextrapolates to a value close to that preported by Lifshitz et al., and they suggestthat HCN and H2O may be the major products in both studies. Hennig et al.measured H and HCN profiles and report that the contribution of the OH channelis 20–25% of that of the main channel. Overall, the data manifest appreciablescatter in both total rate and product distribution.

A potential energy level diagram for CH3NO based on the calculations of Melius(1993) is shown in Fig. 38. Possible reaction channels include

CH3 +NO −→ CH3NO (19a)

−→ CH2NHO (19b)

−→ CH2NOH (19c)

−→ H2CN+OH (19d)

−→ HCN+H2O (19e)

This system is similar in some respects to NH2 + NO, with which it is isoelectronic.In both cases, isomerization via H–atom shifts is followed by a two-channeldissociation, one into radical fragments and one into stable species. However, thehigher barriers for isomerization of CH3NO make stabilization and dissociationof the adduct back to reactants much more likely for this case. Isomerizationto CH2NOH can occur either directly or stepwise via CH2NHO. The inputparameters for the QRRK calculations are listed in Table 11. Transition stategeometries and entropies were calculated with semi-empirical methods,1 and therate coefficients, including temperature-dependent pre-exponential factors, werederived from transition state theory. The similarity with the NH2 + NO systemcontinues in that the rate coefficient for production of the initial adduct was takento be the same.

1 The calculations were done with the MOPAC 6.0 program and PM3 parame-terization. Transition states were located with the TS keyword; RHF was usedfor singlet species and UHF for doublet species. Transition states were confirmedby having one imaginary vibration frequency, by the atom positions, and byanimation of the structure as it changed along the reaction coordinate. Entropyand heat capacity values were computed in MOPAC with the THERMO optionfor vibrations and overall rotation; contributions from frequencies correspondingto internal rotations were replaced by Pitzer-Gwinn estimates of the rotationalentropy and heat capacity.

Page 72: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

196 Anthony M. Dean, Joseph W. Bozzelli

8

9

10

11

0.5 0.7 0.91000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Wolff-Wagner (1988) total Yang et al. (1993) total Haynes (1978) total Hennig-Wagner (1994) total Lifshitz et al. (1993) HCN + H2O Hennig-Wagner (1994) HCN + H2O Hoffmann et al. (1990) H2CN + OH Hennig-Wagner (1994) H2CN + OH

CH3 + NO → Products

FIGURE 2.37. High temperature rate coefficient measurements for CH3 + NO.

A comparison of these calculations to the room temperature data for Ar bath gasis shown in Fig. 39. The QRRK rate coefficient predictions obtained using a normalvalues for collisional energy transfer (1E = 2.6 kJ/mol) are significantly lowerthan observed. Davies et al. (1991) encountered a similar difficulty in analyzingthe fall-off regime and had to invoke unusually efficient energy transfer to accountfor their data. They suggested that the falloff data could be accounted for byrecombination to form a triplet, followed by rapid intersystem crossing to formthe stablized singlet adduct. Figure 39 also illustrates that even the strong collisioncase (β = 1) does not fit the observed falloff. Somewhat better agreement can beforced by assuming greater stability of CH3NO. Figure 39 includes a calculation in

Page 73: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 197

–150

–50

50

150

250

Ent

halp

y (k

J/m

ol)

350

CH3 + NO

CH3NOCH2NHO

CH2NOH

H2CN + OH

HCN + H2O

FIGURE 2.38. Potential energy diagram for CH3 + NO −→ Products.

TABLE 2.11. CH3 + NO −→ Products

Reaction A/s−1 or cm3mol−1s−1 m Ea /kJ mol−1

(1) CH3 + NO −→ CH3NO 1.0×1013

(–1) CH3NO −→ CH3+ NO 1.8×1015 154(2) CH3NO −→ CH2NHO 6.8×109 1.25 198

(–2) CH2NHO −→ CH3NO 2.2×1011 0.98 238(3) CH2NHO −→ CH2NOH 3.1×1010 1.22 184

(–3) CH2NOH −→ CH2NHO 9.7×1010 0.96 222(4) CH3NO −→ CH2NOH 2.9×1010 1.00 243

(–4) CH2NOH −→ CH3NO 2.9×1012 0.47 321(5) CH2NOH −→ HCN + H2O 2.3×1011 0.94 238(6) CH2NOH −→ CH2N + OH 4.3×1016 259

Molecular constants of CH3NO adduct: ν1 = 658 cm{1 (degen=4.62) ν2 = 1867 cm{1

(5.84) ν3 = 3999 cm{1 (1.54); molecular constants of CH2NHO adduct: ν1 = 250 cm{1

(degen=2.22) ν2 = 1119 cm{1 (6.29) ν3 = 3278 cm{1 (3.49); molecular constants ofCH2NOH adduct: ν1 = 384 cm{1 (degen=3.37) ν2 = 1542 cm{1 (6.26) ν3 = 3999 cm{1

(2.37); σ = 4.25 A and ε/K = 410 for all adducts, an average of values for C2H5OH andCH3CN.

Rate coefficients: k1 Kaiser (1993); k−1 microscopic reversibility (MR); k2 TST; k−2 MR;k3 TST; k−3 MR; k4 TST; k−4 MR; k5 TST; k6 via k−6 and MR with k−6 based on OH +CH3 with A(–6) = 4.5×1013 and Ea(–6) = 0, Washida (1980).

Page 74: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

198 Anthony M. Dean, Joseph W. Bozzelli

11

12

13

–2 –1 0 1log (P/atm)

log

( k

/ cm

3 m

ol–1

s–1

)

Davies et al. (1991) 296 K, Ar Kaiser (1993) 297 K, Ar QRRK Ar QRRK β = 1 QRRK Ar, –21

CH3 + NO → Products

FIGURE 2.39. Comparison of QRRK predictions for CH3 + NO to experimental falloffmeasurements in Ar buffer gas near room temperature. The line marked QRRK [Ar] usedthe normal energy transfer parameter for Ar,1E = 2.6 kJ/mol. The line marked QRRK [β= 1] was calculated using a strong collision assumption. Increasing the CH3NO well depthby 21 kJ/mol, QRRK [Ar, –21], while using the usual Ar parameters, results in stabilizationcomparable to invoking the strong collision assumption.

which the well depth for CH3NO was increased by 21 kJ/mol while using the usualcollisional stabilization parameters; the QRRK results are similar to those for thestrong collision case. Use of an even deeper well results in an approximatelyhorizontal displacement of the calculated curves toward lower pressure; the sloperemains steeper than observed. Similar difficulties arise in comparisons to thedata at 407 and 509 K. Tunneling through the isomerization barrier would alsoincrease the low-temperature stabilization rate coefficient computed using theMelius (1993) potential energy surface parameters. While ambiguities remain inunderstanding the low-temperature rate coefficient, the data do show that adductformation is important under the conditions studied.

Page 75: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 199

7

9

11

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

) Lifshitz et al. (1993) HCN + H2O

Hennig-Wagner (1994) HCN + H2O

Hennig-Wagner (1994) H2CN + OH

Yang et al. (1993) total

Miller et al. (1998) total

QRRK CH3NO

QRRK H2CN + OH

QRRK HCN + H2O

P = 1 atm N2

CH3 + NO → Products

FIGURE 2.40. Arrhenius plot for the various channels of CH3 + NO −→ Products aspredicted by QRRK calculations for 1 atm N2. Formation of the stabilized adduct CH3NOis predicted to dominate at lower temperatures, CH2N + OH formation to dominate athigher temperatures, and the HCN + H2O channel to be unimportant at all temperatures.The Miller et al. (1998) calculation considers only the HCN + H2O and H2CN + OHchannels.

Figure 40 illustrates the QRRK predictions for the various reaction channels asa function of temperature at 1 atm N2. Stabilization of the adduct dominates atlower temperatures as expected, while H2CN + OH is predicted to be the majorhigh temperature channel. The transition between CH3NO formation and H2CN +OH formation suggested by Fig. 40, however, is misleading. The computed bonddissociation energy of CH3NO is only 160 kJ/mol, which implies that dissociationof CH3NO back to CH3 and NO is rapid. For this reason, the CH3NO formationchannel can be neglected at the high-temperature conditions of the studies shownin Fig. 37. Even at 1000 K and 10 atm, where Fig. 40 would suggest that CH3NO

Page 76: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

200 Anthony M. Dean, Joseph W. Bozzelli

formation would be the dominant channel, modeling studies for a mixture initiallycontaining 20 ppm CH3 and 2% NO confirmed that the predicted profiles for CH3loss and H2CN formation are but slightly affected if the CH3NO formation channelis omitted from the model.

The calculations predict in particular that the dominant high temperature prod-ucts are H2CN and OH rather than HCN and H2O. This prediction was supportedby additional calculations in which the barrier to formation of HCN and H2Owas reduced by 42 kJ/mol. Even then the H2CN + OH channel was found tobe the most important one above 1000 K. The main reason is the much higherA-factor for dissociation of CH2NOH to H2CN and OH than for isomerization ofCH2NOH and dissociation to HCN and H2O. Using the somewhat higher A-factorfor isomerization proposed by Lifshitz et al. would not change this result.

Some of the high-temperature results are included in Fig. 40 for comparison.Several difficulties are apparent, especially with respect to the importance of theHCN + H2O channel. While the QRRK calculations suggest that it is unimportant,it is reported by Lifshitz et al. to be the only channel and by Hennig and Wagner tobe the dominant one. One way to reconcile this situation would be to assume thatsecondary reactions lead to the observed HCN. In the longer time scale Lifshitzet al. experiments, subsequent reactions of H2CN (cf. Section 7.20) wouldreadily account for the observed HCN production. The faster Hennig and Wagnerexperiments entailed observations of H-atom profiles and were interpreted asmeasures of the rate of OH + H2CN production on the assumption of rapid H2CNdecomposition. There may be complications with this assumption. For example,using the H2CN rate coefficient expressions discussed in Section 7.20 leads to anH2CN dissociation lifetime of about 1 µs for the Hennig and Wagner conditions,sufficient for loss through secondary reactions such as recombination with NO,which is present in high concentrations. Any such reactions would reduce theappearance rate of H atoms and lead to an underestimation of the rate of H2CNand OH appearance. The rate measurements based on HCN formation may becomplicated by presence of vibrationally excited HCN, as the authors point out.Additional experiments and more detailed modeling are needed to resolve thesecomplications. In any event the QRRK results are clear in their support of OH andH2CN being the main products of the CH3 + NO reaction at high temperatures,and the predicted rate coefficient is in reasonable agreement with the observedreaction rate.

The conclusions of Miller et al. (1998) are in agreement with these predictions,as seen in Fig. 40. Their calculated total rate coefficient for sum of the twobimolecular channels is close to the combined QRRK rate coefficients, and theyconclude tentatively that the branching ratio should favor the H2CN + OH channel.

(Nguyen et al. 1996 report a QRRK analysis based upon ab initio molecularelectronic structure calculations that predicts HCN + H2O to be the major productchannel; however, these workers did not investigate the H2CN + OH channel andreport a total rate constant an order of magnitude smaller than the experimentalobservations at high temperatures.)

Page 77: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 201

QRRK calculations over an extended pressure range show that both the HCN+ H2O and the H2CN + OH channel are insensitive to pressure at the highertemperatures while the stabilization channels scale linearly with pressure. (Atpressures above 10 atm,there is some falloff in all channels except CH3NO.) Underall conditions the major channels are either formation of CH3NO, which dominatescompletely at lower temperatures, or H2CN + OH. For low temperatures we adoptthe limiting rate coefficients derived by Kaiser for the stabilization channel in Arbath gas

k0,19a = 1.3× 1018cm6mol−2s−1

k∞,19a = 1.0× 1013 cm3mol−1s−1 .

Using these values we tested the Gilbert et al. (1983) and the Stewart et al.(1989) fitting formulas against the data shown in Fig. 36. The Stewart et al.formula (Section 3.1) with a = 0.03, b = −790 and c = 1.0 was found to givesignificantly better fits for the temperature range of the data, 300 to 500 K. Athigher temperatures, where the other channels begin to have an influence, wesuggest using the QRRK estimates, in cm3mol−1s−1 units, for the stabilizationchannel in N2:

k19a = 3.6× 1035T−8.25 exp(−2420/T ) 0.1 atm

= 1.0× 1037T−8.38 exp(−2630/T ) 1.0 atm

= 4.6× 1041T−9.39 exp(−4160/T ) 10 atm.

While these expressions underpredict the observed low temperature stabilizationrate coefficient, they should account properly for the temperature and pressuredependence over the whole range of these variables. As discussed earlier, rapiddissociation of CH3NO makes modeling results insensitive to the stabilization ratecoefficient at temperatures over 1000 K.

The QRRK estimate for the primary high-temperature reaction channel is

k19d = 2.2× 109T 0.75 exp(−5900/T ) cm3mol−1s−1.

The QRRK estimate for the HCN + H2O channel is

k19e = 4.9× 108T 0.46 exp(−6240/T ) cm3mol−1s−1.

2.6.16 CH3 + N −→ Products

This reaction provides a pathway for production of HCN and ultimately NO. Ithas been extensively studied at low temperatures (Stief et al. 1988; Marston etal. 1989a; Marston et al. 1989b), including measurement of the branching ratiobetween the major product channels

CH3 + N −→ H2CN+H (20a)

−→ HCN+H2 . (20b)

Page 78: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

202 Anthony M. Dean, Joseph W. Bozzelli

The reaction is observed to be very fast, with about 90% of the reaction producingH2CN + H and the remaining 10% producing HCN + H2 via a spin-forbiddenchannel. The total rate has also been measured in shock tube experiments between1600 and 2000 K (Davidson and Hanson 1990b). These data are summarized inFig. 41. The high temperature rate coefficients are significantly lower than onewould expect from a linear extrapolation of the lower temperature results.

Figure 42 shows the relevant potential energy level diagram. Other channelsbesides those discussed in the experiments include

CH3 +N −→ 3CH3N (20c)

−→ 3CH2NH (20d)

−→ 1CH2NH (20e)

−→ HCNH+H . (20f)

It can be seen that the possible reaction pathways are rather complex. Toimprove clarity, solid lines were used to indicate paths to allowed reaction productchannels and isomerizations; dashed lines show the spin-forbidden reactions. The3CH3N formed can dissociate to H2CN + H, isomerize to 3CH2NH, undergo aspin forbidden molecular elimination of H2 or dissociate back to reactants. The3CH2NH can dissociate to H2CN + H or to HC=NH + H or undergo intersystemcrossing to 1CH2NH. The adduct well is sufficiently deep, about 290 kJ/mol, toput the barriers for isomerization and dissociation below the entrance channel.Thus one expects rapid reactions. Input parameters for the QRRK calculationsare listed in Table 12.

These calculations indicate that H2CN + H is the dominant channel at alltemperatures and pressures with very slight rate decreases as the temperatureis increased due to an increased reverse reaction rate, i.e., dissociation back toreactants. The QRRK predictions are compared to the observations in Fig. 41.Although the temperature dependence of the low temperature data is not accountedfor, the overall agreement is acceptable. One possible explanation of the lowtemperature results, suggested by Steif and co-workers, is the participation of aseparate channel, formation of 1CH3N, via an energy barrier, followed by rapidisomerization to 1CH2NH, yielding an adduct with about 530 kJ/mol excess energythat could readily dissociate via a loose transition state to H2CN + H or HC=NH+ H. The isomerization would have to have a higher Arrhenius A-factor thanthe reverse dissociation, and adduct formation would still require an unusuallyhigh rate coefficient to explain the rapid increase in rate with temperature. Thispathway was not included in the QRRK calculations.

Page 79: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 203

13.75

14

1 2 3 4 51000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Stief et al. (1988)

Marston et al. (1989)

Davidson-Hanson (1990)

QRRK (total)

CH3 + N → Products

FIGURE 2.41. QRRK predictions and experimental results for CH3 + N.

50

250

450

650

Ent

halp

y (k

J/m

ol)

CH3 + N

3CH3N

3CH2NH

H2C=NHHCN + H2

H2CN + H

HCNH + H

FIGURE 2.42. Potential energy diagram for CH3 + N −→ Products. Spin-forbiddenpathways are indicated by dashed lines.

Page 80: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

204 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.12. CH3 + N −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) CH3 + N −→ 3CH3N 8.0×1013 0(–1) 3CH3N −→ CH3 + N 1.9×1015 293(2) 3CH3N −→ CH2=N + H 9.3×1013 148(3) 3CH3N −→ HCN + H2 7.0×1013 188(4) 3CH3N −→ 3CH2=NH 1.4×1014 205

(–4) 3CH2=NH −→ 3CH3N 7.0×1013 163(5) 3CH2=NH −→ CH2=N + H 2.9×1013 89(6) 3CH2=NH −→ HC=NH+H 8.7×1013 118(7) 3CH2=NH −→ 1CH2=NH 1.0×1012 51

(–7) 1CH2=NH −→ 3CH2=NH 2.0×1012 348(8) 1CH2=NH −→ HCN + H2 8.0×1013 389

Molecular constants of 3CH3N adduct: ν1 = 799 cm{1 (degen=2.61) ν2 = 1426 cm{1

(3.39) ν3 = 2772 cm{1 (3.00); molecular constants of 3CH2=NH adduct: ν1 = 779 cm{1

(degen=3.27) ν2 = 1487 cm{1 (3.03) ν3 = 3419 cm{1 (2.70); molecular constants of1CH2=NH adduct: ν1 = 884 cm{1 (degen=2.81) ν2 = 2046 cm{1 (5.59) ν3 = 3986 cm{1

(0.60); σ = 3.49 A and ε/K = 350 for all adducts.

Rate coefficients: k1 was estimated by assuming that the dynamics of CH3N adductformation resemble those of the reaction O + CH3, for which Baulch et al. (1992) adoptedthe rate coefficient 8.4×1013 of Slagle et al. (1987); k−1 microscopic reversibility (MR);k2 from k−2 and MR with A(–2) = 9.35×1012, (0.5×A-factor for H + C2H4) and Ea(–2)= 8 kJ/mol from Mallard et al. (1993); k3 TST, Ea from Gordon et. al (1986); k4 TST,loss of symmetry, gain of rotor, Ea from Gordon et al. (1986)—cf. Walia and Kakar(1990,1991); Melius calculates ∆r H◦= 59 kJ/mol; k−4 MR; k5 via k−5 and MR with k−5taken to be the same as k−2; k6 via k−6 and MR with k−6 taken to be the same as k−2; k7triplet–singlet conversion, Ea from Gordon et al. (1986) and our estimate for A; k−7 MR;k8 TST, Ea based on 1,2 H2 elimination from C2H4 (Yoshimine 1989).

Fig. 43 illustrates the effect of pressure on the various pathways at 1500 K.The rate coefficients for the major channels are virtually independent of pressure,indicating that there is no need to assign pressure-dependent rate coefficients forthis system. The most important channel is H2CN + H, giving about 80% ofthe total reaction. Our results indicate that HCN + H2 is the next importantchannel with about 13% of the total rate, followed by HC=NH + H at about 5%of the total rate. These results are consistent with the experimental observationsthat about 90% of the reaction produces H–atoms. As expected, most of theH2CN + H is produced from the initially-formed 3CH3N, since isomerization to3CH2NH has both a lower A-factor and a higher barrier than dissociation. Theslower isomerization also accounts for the much slower rate of production of

Page 81: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 205

4

6

8

10

12

14

–2 0 2log (P/atm)

log

( k

/ cm

3 m

ol–1

s–1

)

3CH3N H2CN + H via CH3N HCN + H2 via CH3N 3H2CNH H2CN + H via 3H2CNH HCNH + H H2C=NH HCN + H2 via H2C=NH total

T = 1500 K

CH3 + N → Products

FIGURE 2.43. Effect of pressure at 1500 K for N + CH3−→ Products. The faster ratesare virtually independent of pressure until above 100 atm.

HCNH + H. In much the same way, most of the HCN + H2 is produced directlyfrom the 3CH3N via the spin-forbidden dissociation. The HCN and the HC=NHchannels increase slightly with increasing temperature, as shown in Fig. 44. Giventhe extremely rapid rate, we suggest the QRRK results be used to estimate thebranching ratios at the high temperatures of interest in combustion. These resultsare, in cm3mol−1s−1 units,

k20a = 6.1× 1014T−0.31 exp(−145/T )

k20b = 3.7× 1012T 0.15 exp(+45/T )

k20f = 1.2× 1011T 0.52 exp(+185/T ) .

Page 82: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

206 Anthony M. Dean, Joseph W. Bozzelli

8

10

12

14

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

3CH3N H2CN + H via CH3N HCN+H2 via CH3N 3H2CNH H2CN + H via 3H2CNH HCNH + H H2C=NH HCN + H2 via H2C=NH total

P = 1 atm N2

CH3 + N → Products

FIGURE 2.44. Arrhenius plot for CH3 + N −→ Products predicted by QRRK calculationsfor 1 atm N2 buffer gas. The rate coefficients for the dominant channels are all predictedto be essentially independent of temperature.

2.6.17 CH3 + NH2 −→ Products

This reaction can lead to recombination and other products

CH3 +NH2 −→ CH3NH2 (21a)

−→ CH2NH2 +H (21b)

−→ CH3NH+H (21c)

−→ H2C=NH+H2 . (21d)

Rate coefficients were estimated using the QRRK method and the parametersgiven in Table 13.

Page 83: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 207

The results at 1 atm N2 are shown in Fig. 45. The dominant channel is formationof the stabilized adduct up to very high temperatures. Of the various eliminationchannels, the formation of CH2NH2 + H is favored. As seen in Fig. 46, even at1500 K stabilization is predicted to dominate at pressures above 0.1 atm.

TABLE 2.13. CH3 + NH2 −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) CH3 + NH2 −→ CH3NH2 2.3×1013 0(–1) CH3NH2 −→ CH3 + NH2 1.2×1016 354(2) CH3NH2 −→ C

.H2NH2 + H 2.0×1015 395

(3) CH3NH2 −→ CH3N.

H + H 1.8×1015 420(4) CH3NH2 −→ H2C=NH + H2 4.9×1013 433

Molecular constants of CH3NH2 adduct: ν1 = 576 cm{1(degen=3.12) ν2 = 1416 cm{1(6.46)ν3 = 3121 cm{1(4.92); σ = 3.77 A and ε/K = 364.

Rate coefficients: k1 geometric mean of CH3 + CH3 and NH2 + NH2 from Mallard et al.(1993); k−1 microscopic reversibility (MR); k2 via k−2 and MR with k−2 = 3.6×1013

based on H + C2H5 from Baulch et al. (1992); k3 same as k2; k4 TST with loss of 1 rotor,reaction degeneracy = 6, barrier based on C2H6 −→ C2H4 + H2.

The calculations suggest a slight pressure dependence for the eliminationchannels, but at the temperatures where these are important this effect is sufficientlysmall that it can be neglected. Because of the importance of the stabilizationchannel one must consider subsequent reactions of the monomethylamine adduct.The QRRK analysis suggests the rate coefficient expressions in cm3mol−1s−1

units:

k21a = 1.3× 1054T−12.72 exp(−7855/T ), 0.1 atm N2, 600–2500 K

k21a = 5.1× 1052T−11.99 exp(−8450/T ), 1.0 atm N2, 600–2500 K

k21a = 1.6× 1047T−10.15 exp(−7895/T ), 10 atm N2, 600–2500 K

k21b = 1.4× 1014T−0.43 exp(−5590/T )

k21c = 4.4× 1013T−0.31 exp(−8375/T )

k21d = 4.8× 1011T−0.20 exp(−9765/T )

The only experimental work on this system we are aware of (Higashiharaet al. 1987; Hwang et al. 1990; Lifshitz et al. 1991) used extensive com-puter simulations to derive elementary reaction rate coefficients applicable tomonomethylamine pyrolysis and oxidation. For typical experimental conditionsused by these authors to derive the CH3NH2 thermal dissociation rate coefficient(1% CH3NH2 in argon at 1 atm and 1400 K) the QRRK analysis gives a fallofffrom the high pressure limit of about a factor of 10, somewhat less than half of theamount they derived from their experiments. Because the thermochemistry used

Page 84: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

208 Anthony M. Dean, Joseph W. Bozzelli

8

10

12

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

CH3NH2

CH2NH2 + H

CH3NH + H

H2C=NH + H2

CH4 + NH

CH2 + NH3

P = 1 atm N2

CH3 + NH2 → Products

FIGURE 2.45. Arrhenius plot for the various channels of CH3 + NH2−→ Productspredicted by QRRK (filled symbols) and DHT (open symbols) calculations for 1 atmN2 buffer gas. Stabilization is predicted to be the dominant channel up to quite hightemperatures; the rate of the CH2NH2 + H channel does not equal the stabilization rateuntil 1800 K.

by these authors is different from ours and because of the complicated chemistrythat they had to describe, extensive reanalysis of their data would be required todetermine how well our QRRK rate coefficients for this and the other unimolecularreactions involved account for their data.

In addition to adduct formation followed by stabilization or dissociation inreactions 21b–21d, H-atom transfer can be significant:

CH3 +NH2 −→ CH4 +NH (21e)

−→ CH2 +NH3 (21f)

Page 85: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 209

8

10

12

–2 0 2log (P/atm)

log

( k

/ cm

3 m

ol–1

s–1

)

CH3NH2 CH2NH2 + H CH3NH + H H2C=NH + H2 CH4 + NH CH2 + NH3

T = 1500 K

CH3 + NH2 → Products

FIGURE 2.46. Effect of pressure at 1500 K predicted by QRRK calculations for CH3 +NH2−→ Products. Even at this relatively high temperature stabilization dominates at N2buffer gas pressures above 0.1 atm.

The DHT method described in Section 4 leads to the rate coefficient expressions,in cm3mol−1s−1 units,

k21e = 2.8× 106 T 1.94 exp(−4635/T )

k21f = 1.6× 106 T 1.87 exp(−3810/T ) .

These estimates are included in Figs. 45 and 46. It can be seen that these channelstogether with the chemical activation channel producing CH2NH2 and H dominateat high temperatures.

Page 86: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

210 Anthony M. Dean, Joseph W. Bozzelli

2.6.18 CH2 + N2 −→ Products

The methylene radical reacts differently in its singlet 1CH2 and triplet 3CH2 forms.The more stable, by 38 kJ/mol, triplet form is commonly designated just as CH2.

2

4

6

8

10

0.5 0.7 0.91000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Blauwens et al. (1978) NH + HCN Miller-Bowman (1989) NH + HCN QRRK CH2NN QRRK NH + HCN

3CH2 + N2 → Products

FIGURE 2.47. Comparison of QRRK predictions for reaction in 1 atm N2 to earlier ratecoefficient estimates for 3CH2 + N2.

Attack of triplet methylene on molecular nitrogen was postulated as one of thesources of Prompt NO in rich hydrocarbon flames by Blauwens et al. (1978) andHayhurst and Vince (1980). However, the kinetics and energetics of such flamesmake it difficult to distinguish experimentally between this reaction and CH + N2.Miller and Bowman (1989) use a much smaller rate coefficient than Blauwenset al., as shown in Fig. 47. Fig. 48 illustrates the potential energy level diagramfor the reaction based upon the parameters listed in Table 14. Possible product

Page 87: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 211

250

350

450

550

650

750

3CH2 + N2

3CH2NN

CH2N + N

3HC=NNH

HCN + 3NH

Ent

halp

y (k

J/m

ol)

FIGURE 2.48. Potential energy diagram for 3CH2 + N2. All products are separated fromthe reactants by large energy barriers.

channels include:

CH2 +N2 −→ CH2N2 (22a)

−→ H2CN+ N (22b)

−→ HC=NNH (22c)

−→ HCN+NH (22d)

There is an initial barrier of 66 kJ/mol for addition and a relatively shallowwell, about 160 kJ/mol, below the top of the barrier to CH2N2. Subsequentisomerization of CH2N2 requires about 220 kJ/mol relative to the reactants.Although isomerization would rapidly lead to HCN + 3NH, the high barrier wouldbe expected to block this channel. Thus one expects the most likely fates of theinitially-formed adduct to be stabilization and/or dissociation back to reactants.

Page 88: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

212 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.14. 3CH2 + N2 −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) 3CH2 + N2 −→ 3CH2NN 2.0×1012 66(–1) 3CH2NN −→ 3CH2 + N2 4.6×1014 159(2) 3CH2NN −→ 3CH2N + N 7.0×1015 429(3) 3CH2NN −→ 3HC=N–NH 1.1×1013 314

(–3) 3HC=N–NH −→ 3CH2NN 4.3×1011 25(4) 3HC=N–NH −→ HCN + 3NH 1.8×1013 7

Molecular constants of 3CH2NN adduct: ν1 = 616 cm{1 (degen=2.92) ν2 = 1163 cm{1

(3.64) ν3 = 2706 cm{1 (2.45); molecular constants of 3HC=N–NH adduct: ν1 = 342 cm{1

(degen=2.45) ν2 = 401 cm{1 (0.59) ν3 = 1300 cm{1 (5.46); σ = 3.49 A and ε/K = 298,from CH2CO, for both adducts.

Rate coefficients: k1 from 3CH2 + O2, Bohland et al. (1984) and Darwin et al. (1989), Ea

from Melius (1988); k−1 from microscopic reversibility (MR); k2 via k−2 and MR withk−2 = 7.1×1013, based on N + NH2 of Whyte and Phillips (1983); k3 TST, with loss ofhindered rotor, degeneracy 2, and Ea =∆r H◦ + 25 kJ/mol; k−3 MR; k4 from k−4 and MRwith k−4 = k1.

The QRRK parameters listed in Table 14 lead to the results compared to theBlauens et al. and Miller and Bowman expressions in Fig. 47. As expected,formation of the stabilized adduct CH2N2 is predicted to be the dominant channel,but even that channel is well below the expression obtained by Blauwens etal. Moreover, the stabilization rate is deceiving, because this adduct dissociatesprimarily by reverse reaction back to CH2 + N2, which must be considered in anymechanism where the forward reaction is included. The equilibrium constant shiftsthis system toward the dissociation channel at higher temperature, suppressingthe contribution of the stabilization channel. The NH + HCN channel is muchslower, close to the expression proposed by Miller and Bowman. The QRRKanalysis indicates that the results at other pressures (other than for the stabilizationchannels) are similar to those shown in Fig. 47 for 1 atm N2. Thus the QRRKcalculation is in agreement with the proposal of Miller and Bowman that this ratecoefficient is much smaller than would be needed to explain Prompt NO formationas a result of triplet methylene reacting with nitrogen. As a consequence, the dataof Blauwens et al. are probably best interpreted in terms of Prompt NO beingformed in the reaction CH + N2. We recommend use of the QRRK results forthe stabilization channel and the Miller-Bowman expression for reaction (22d),giving in cm3mol−1s−1 units for N2 bath gas

k22a = 9.3× 1030T−7.01 exp(−9935/T ) 0.1 atm

= 1.6× 1032T−7.07 exp(−10050/T ) 1.0 atm

= 4.3× 1033T−7.18 exp(−10500/T ) 10 atm

k22d = 1.00× 1013 exp(−37240/T ) .

Page 89: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 213

HC

N N

CH

N N

H

Ent

halp

y (

kJ/m

ol)

HCN + 1NH

1CH2 + N2

1CH2NN

+ H

250

350

450

550

650

750

1CH2 + N2

1HCN=NH

FIGURE 2.49. Potential energy diagram for 1CH2 + N2. Formation of all distinguishableproducts is very endothermic.

An energy diagram for the reaction of 1CH2 with N2 is shown in Fig. 49. Itforms a cyclic intermediate complex having a shallow well, about 100 kJ/mol,and dissociation back to reactants is its dominant unimolecular reaction. There issome stabilized adduct production at higher pressures, but the shallow well andthe high A-factor for dissociation, about 1016 s−1, imply a half life of less than10 ps at 1000 K. Considering the probable concentrations of radical species suchas OH, CH3, O or H which could attack this adduct, further important bimolecularreactions of this intermediate are unlikely.

2.6.19 3CH2 + NO −→ Products

Until recently, rate coefficient measurements for this reaction were confined tolow temperatures (Laufer and Bass 1974; Vinckier and DeBruyn 1979; Darwinet al. 1989; Seidler et al. 1989). In these experiments the product distributionswere not determined, although Seidler et al. concluded that OH and H werenot the main products, i.e., less than 10% of the total. In later work Atakan etal. (1992) extended the temperature range up to 1025 K and also measured the rateof production of some of the reaction products. In particular, they reported thatthe HCN + OH channel contributes virtually nothing at 300 K but increases to 63

Page 90: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

214 Anthony M. Dean, Joseph W. Bozzelli

11.5

12

12.5

13

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Darwin et al. (1989)Laufer-Bass (1974)Vinckier-Debruyn (1979)Seidler et al. (1989)Atakan et al. (1992)Atakan et al. (1992) OHMiller-Bowman (1989) HCNO + HBauerle et al. (1995) HCNO + HBauerle et al. (1995) HCN + OHQRRK total

CH2 + NO → Products

FIGURE 2.50. Comparison of QRRK predictions for the total rate coefficient to measure-ments for 3CH2 + NO −→ Products.

±25% of the total products at 1025 K. In contrast to the conclusions of Seidler etal., Grussdorf et al. (1997) report that the dominant channel produces HCNO + H(about 84%), with HCN + OH accounting for about 15%. Bauerle et al. (1995)measured H, O and OH appearance rates in shock waves over the temperaturerange from 1100 to 2600 K and concluded that HCNO + H and HCN + OH aremajor product channels, with both rate coefficients increasing with temperature.These results, along with the expression used by Miller and Bowman in their flamemodeling, are shown in Fig. 50. A potential energy diagram for the reaction isshown in Fig. 51. Possible channels include:

3CH2 + NO −→ CH2NO (23a)

−→ H2CyCON (23b)

−→ HNCHO (23c)

−→ NCHOH (23d)

Page 91: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 215

CN

O

H

H

CH2NO

Ent

halp

y (k

J/m

ol)

3CH2 + NO

0

HCNOH

HNCHO

H2NCO

NCHOHNH2 + CO

HNCO + H

HCN + OH

HOCN + H

HCNO + H

H2CN + O

100

200

300

400

FIGURE 2.51. Potential energy diagram for 3CH2 + NO.

−→ H2NCO (23e)

−→ HCNOH (23f)

−→ HCNO+ H (23g)

−→ HOCN+ H (23h)

−→ HCN+ OH (23i)

−→ HNCO+ H (23j)

−→ NH2 + CO (23k)

−→ H2CN+O (23l)

The adduct CH2NO well is quite deep, about 300 kJ/mol, and all reaction barriersexcept the one to H2CN + O are lower in energy than the entrance channel. Thus,one expects that the unimolecular reactions of the initially-formed adduct willbe rapid and stabilization will be minimal. The CH2NO adduct can dissociate toHCNO + H, dissociate to H2CN + O, or isomerize. There is a high energy pathwayto HCN + OH via HCNOH. Alternatively, there is a lower energy pathway viathe cyclic intermediate H2CyCON to HNCHO. This in turn can isomerize furtherto NCHOH or H2NCO or dissociate to HNCO + H. Finally these isomers candissociate to the indicated products. The QRRK parameters used are listed inTable 15.

Page 92: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

216 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.15. 3CH2 + NO −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) CH2 + NO −→ CH2NO 2.0×1013 0(-1) CH2NO −→ CH2 + NO 7.7×1015 297(2) CH2NO −→ H + HCNO 9.3×1013 229(3) CH2NO −→ CyH2CON 1.5×1013 79

(-3) CyH2CON −→ CH2NO 1.9×1013 25(4) CyH2CON −→ HNCHO 9.0×1013 121

(-4) HNCHO −→ CyH2CON 5.2×1013 279(5) HNCHO −→ HNCO + H 7.2×1013 84(6) HNCHO −→ NCHOH 1.7×1013 130

(-6) NCHOH −→ HNCHO 7.3×1012 141(7) HNCHO −→ H2NCO 1.5×1013 88

(-7) H2NCO −→ HNCHO 2.6×1012 182(8) NCHOH −→ HCN + OH 1.8×1014 120(9) NCHOH −→ HOCN + H 1.8×1013 157

(10) H2NCO −→ NH2 + CO 3.5×1012 123(11) H2NCO −→ HNCO + H 1.2×1013 149(12) CH2NO −→ HCNOH 4.4×1013 237

(-12) HCNOH −→ CH2NO 8.4×1012 151(13) HCNOH −→ HCN + OH 1.1×1014 15(14) CH2NO −→ H2CN + O 3.8×1015 318

The QRRK frequencies in cm{1 units (ν1,ν2,ν3) and degeneracies for the adducts areCH2NO: (616,2.92; 1163,3.64; 2706, 2.45); CyH2CON: (819,3.76; 823,2.45; 2868,2.79);HNCHO: (251,0.92; 910,5.67;3450,2.40); NCHOH: (250,0.88; 903,5.70; 3448,2.41);H2NCO: (251,1.68; 1016,4.36; 2744,2.46); HCNOH: (350,3.32; 1473, 4.43; 3996,1.24);σ = 3.97 A and ε/K = 436.

Rate coefficients: k1 average of Darwin et al. (1989) and Siedler et al. (1989); k−1 frommicroscopic reversibility (MR); k2 MR with A−2 = 1.5×1013 from H + C3H7, Allara etal. (1980), Ea(–2) = 9 kJ/mol from H + C2H4; k3 TST with loss of hindered rotor, ∆S‡ =–9.3 J/K, Ea =∆r H◦ + 25 kJ/mol; k−3 MR; k4 TST with 1,2 H shift,∆S‡ = 0, degeneracyof 2, Ea = ring strain of 113 kJ/mol + Ea(abstraction) of 8 kJ/mol; k−4 MR; k5 MR withA−5 = 1.5×1013 from 0.5(H + C2H4) and barrier from Melius (1988)—cf. text; k6 TST,∆S‡ = –16.7 J/K and Ea from Melius (1988)—cf. text; k−6 MR; k7 TST,∆S‡ = –9.2 J/K,Ea from Melius (1988)—cf. text; k−7 MR; k8 MR with k−8 = 1.2×1013 exp(–1060/T )from Miller et al. 1986; k9 MR with k−9 = 1.5×1013 exp(–1060/T ), from 0.5(H + C2H4);k10 MR with k−10 = 5.0×1011 exp(–3300/T ), from CH3 + CO (Anastasi et al. 1982);k11 MR with k−11 = 1.5×1013 exp(–1060/T ), from 0.5(H + C2H4); k12 TST with ∆S‡

= –6.2 J/K, degeneracy of 2, Ea from Melius (1988)—cf. text; k−12 MR; k13 MR withA−13 = 1.5×1013 (Miller et al. 1986) and barrier from Melius (1988); k14 MR with k−14= 8.0×1013 (Tsang et al. 1986).

Page 93: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 217

The predicted rate coefficient for loss of reactant is compared to the observationsin Fig. 50. The predictions show less of a negative temperature dependence thanobserved at low temperatures, but a sharper falloff near combustion temperatures.The magnitude of the rate coefficient is fixed by the rate of adduct formation, andwe have chosen this to match the measurements of Seidler et al. and Darwin et al.

Fig. 52 shows QRRK predictions for the various channels at 1 atm pressure.The major one is predicted to be HNCO + H, with most coming directly fromHNCHO rather than through isomerization to H2NCO followed by dissociation.The next most important channel is HCNO + H, followed by OH + HCN.

Grussdorf et al., however, report that HCNO is the dominant channel. Thepotential energy curves shown in Fig. 51 suggest that decreased production ofHNCO relative to HCNO would be favored by an increased isomerization barrierfor the conversion of the cyclic H2CON to HNCHO, i.e., decreasing the rate ofreaction (4) in Table 15. For HCNO production to dominate, however, this barrierwould have to be increased by more than 40 kJ/mol.

It is even more difficult to reconcile a non-H or non-OH channel dominating atlow temperature, as reported by Seidler et al.; the only likely such channel at lowtemperature would be formation of NH2 and CO. For this pathway to dominate,the barrier between HNCHO and HNCO + H must be substantially higher, or theone between HNCHO and H2NCO much lower, than computed. The A-factor fordissociation is expected to be substantially larger than for isomerization,so the onlyplausible way to increase the selectivity to favor NH2 and CO production wouldbe by adjusting these barriers. Adjustments of other barriers were also tested.For example, increasing the barrier for isomerization of HNCHO to H2NCO by50 kJ/mol decreases the production rate of NH2 and CO by only about a factorof 2, with little effect on the other channels. Similarly, increasing the barrier fordissociation of HNCHO to HNCO and H by 20 kJ/mol has virtually no effectat all. Given the structure of the Fig. 51 potential energy diagram, such smallchanges are to be expected. In all cases, the energy of the initially formed CH2NOadduct is so much higher than the barriers that only small changes can be effectedby adjusting them.

Near 1000 K, we find that the branching ratio to OH + HCN is only about14%, consistent with the report of Grussdorf et al., and well below that reportedby Atakan et al. One would need very large changes in the barriers to allow OH+ HCN to be the major channel. For example, lowering the barriers for bothCH2NO −→ HCNOH and HNCHO −→ NCHOH by 21 kJ/mol (these are therate-limiting steps for both the OH + HCN channels) increases this channel byabout a factor of two, but HNCO + H is still dominant. At higher temperatures, O+ H2CN becomes important. It is thus difficult to reconcile the QRRK calculationswith the high temperature observations of Bauerle et al.; the only channel witha positive temperature dependence was found in the computations to be the oneleading to H2CN and O. The channels producing H and OH all become slowerwith increasing temperature.

Page 94: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

218 Anthony M. Dean, Joseph W. Bozzelli

9

11

13

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

CH2NO

HCNO + H

H2CN + O

HOCN + H

NH2 + CO

HCN + OH

HNCO + H

P = 1 atm N2

3CH2 + NO → Products

FIGURE 2.52. Arrhenius plot for the rate coefficients of various channels of 3CH2 + NO−→ Products as predicted by QRRK calculations for 1 atm N2 buffer gas. Stabilizationis relatively unimportant and many different products are expected.

Reaction exothermicity is not the controlling factor for any of the channels.For example, one sees a substantial contribution from HCNO + H, a relativelyhigh energy channel. This can be traced to the fact that this channel is directlyformed from the initially-formed adduct, and the resulting large A-factor (relativeto isomerization) offsets the larger energy barrier. Similar considerations applyto O + H2CN. Conversely, the lowest energy channel, NH2 + CO, is onlyaccessible via a series of isomerizations, each with a relatively tight transitionstate, making it relatively unimportant. The most important stabilization channelis insignificant compared to the dissociation channels, even at 1 atm pressure andlow temperature.

Page 95: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 219

Because stabilization is found to be relatively unimportant, the effect of pressureis small. The calculations indicate that pressures well above 10 atm are required forstabilization to be significant even at low temperatures. The negative temperaturedependence is also problematic. Our analysis (Fig. 50) suggests a more gradualdecline at lower temperatures—followed by a steeper drop as one gets into thefalloff region—than was observed by Vinckier et al. The work of Atakan et al.,however, shows a temperature dependence closer to our calculations. We arereluctant to use QRRK analysis to revise a measured temperature dependence.It would be helpful if this rate coefficient could be remeasured over a widertemperature range, particularly at temperatures above 1000 K. In the interim, wesuggest use of the measured temperature dependence of Atakan et al., but with alarger A-factor to reflect the faster room temperature measurements of Siedler etal. and Darwin et al.; i.e.,

k23(total) = 1.0× 1013 exp(+190/T ) cm3mol−1s−1

for moderately high temperatures and pressures of less than 10 atm, wherestabilization is unimportant. In view of the differences between the calculationsand the available data we reluctantly recommend, in cm3mol−1s−1 units,

k23g = 3.8× 1013T−.36 exp(−290/T )

k23i = 2.9× 1014T−.69 exp(−380/T )

k23j = 3.1× 1017T−1.38 exp(−640/T )

k23k = 2.3× 1016T−1.43 exp(−670/T )

k23l = 8.1× 107T 1.42 exp(−2070/T ) .

There is a clear need for further experimental and theoretical work on the 3CH2 +NO reactions.

2.6.20 CH + N2 −→ Products

The reaction of CH with N2, in flames and in other high temperature oxidationsystems, is important in coupling hydrocarbon oxidation chemistry with nitrogenchemistry. By forming HCN + N it generates two species that can both reactfurther to form NO. As discussed in Section 6.18, it is unlikely that the CH2 + N2reaction plays a significant role in NO formation, leaving CH + N2 as the mainroute to Prompt NO. Because CH tends to react via insertion, forming two newbonds, one has to account for chemical activation effects in these very exothermicreactions.

The experimental rate coefficient measurements are shown in Fig. 53. Bermanand Lin (1983) studied the pressure and temperature dependence and interpretedtheir results in terms of adduct formation. An RRKM analysis was used to

Page 96: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

220 Anthony M. Dean, Joseph W. Bozzelli

8

9

10

11

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Berman-Lin (1983) 100 TorrDean et al. (1990)Lindackers et al. (1990)Medhurst et al. (1993) 100 TorrMatsui et al. (1985) Becker et al. (1996) 20 TorrQRRK 20 TorrQRRK 100 TorrQRRK 20 Torr, –42 kJQRRK 100 Torr, –42 kJ

CH + N2 → Products

FIGURE 2.53. Comparison of QRRK predictions to rate coefficient measurements for CH+ N2−→ Products in Ar buffer gas. Two sets of QRRK calculations, differing in the welldepths assigned to the adducts, are shown. Assigning the adduct well depths calculatedby Manaa and Yarkony (1991, 1992) substantially underpredicts the low temperature data.Assigning wells 42 kJ/mol deeper results in rate coefficient values only slightly lower thanobserved at low temperature. Both sets correctly predict the increasing rate coefficient athigher temperatures.

describe the observed falloff in rate with increasing temperature. Becker etal. (1989) measured the CH decay rate at room temperature and 2 Torr. Themeasurements were later extended to higher temperatures (Becker et al. 1992a;Medhurst et al. (1993); Becker et al. (1996). The data are consistent, showing boththe temperature and the pressure dependences expected for an adduct formationmechanism. Although the first report of Becker et al. (1992a) showed a definiteminimum near 700 K, their 1996 data (included in Fig. 53) show no such minimum;the authors consider the earlier data at temperatures over 560 K to be in error.

The higher temperature measurements (Matsui and Yuuki 1985;Dean et al. 1990;Lindackers et al. 1990) all show increases in rate coefficient with increasingtemperature. As pointed out by Berman and Lin, this behavior is consistent

Page 97: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 221

N

C

N

H

CH + N2

Ent

halp

y (k

J/m

ol)

HCNN

HCN + N

400

500

600

700

(HCN2)

FIGURE 2.54. Potential energy diagram for the CH + N2. The solid lines are based onManaa and Yarkony, (1991, 1992) and a barrier between adducts from Martin and Taylor(1993). The dashed lines show the result of making both adducts more stable by 42 kJ/mol.For QRRK calculations it was assumed that both adducts can be reached directly from thereactant channels.

with formation of an energized adduct HCNN that can be collisionally stabilized,dissociate back to reactants, causing the falloff in rate at intermediate temperatures,or at higher temperatures dissociate via an endothermic channel.

The initially formed adduct has a doublet electronic state, and so there mustbe curve crossing to a quartet surface to form HCN + N. This crossing wasstudied theoretically by Manaa and Yarkony (1991; 1992), who report that thecurve-crossing barrier is about 40 kJ/mol above the entrance channel. They alsoconclude that there are two adduct isomers, an acyclic HCNN and a cyclic insertionspecies HCN2, which have well depths of about 92 and 100 kJ/mol respectively.(Figure 54) Calculations by Martin and Taylor (1993) show a substantial (about235 kJ/mol) barrier between the isomers. Other calculations (Walch 1993c;Seidemann and Walch 1994; Seidemann 1994) indicate a significant entrancechannel barrier for HCN2 formation. We did not attempt to account for thisfeature. An additional complication is that Manaa and Yarkony report that theprobability for crossing from the doublet to the quartet surface is much less thanunity. Rodgers and Smith (1996) analyzed the system in terms of two separatereactions, a low temperature one forming the HCNN adduct with no entrancechannel barrier and a high temperature one forming HCN and N through the cyclicHCN2 structure. They assumed that the barrier to form HCN2 is rate-limiting.

Page 98: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

222 Anthony M. Dean, Joseph W. Bozzelli

Earlier attempts to characterize the well depth gave widely different results;Miller and Bowman (1989) use 75 kJ/mol while Berman and Lin use 222 kJ/mol.Medhurst et al. found that a 120 kJ/mol well was consistent with their data.Rodgers and Smith found that a well depth in the range from 105 to 145 kJ/mol,depending on the choice of collision efficiency, gave the best fits to experiment.The dashed lines in Fig. 54 show adjustments of the well depths made in attemptsto fit the falloff data. (V. infra.)

TABLE 2.16. CH + N2 −→ Products

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1a) CH + N2 −→ HCNN 1.3×1013 0(-1a) HCNN −→ CH + N2 1.6×1014 82, 124(1b) CH + N2 −→ HCN2 1.3×1013 0

(-1b) HCN2 −→ CH + N2 1.4×1015 92, 134(2) HCNN −→ HCN2 1.1×1013 209

(-2) HCN2 −→ HCNN 9.7×1013 220(3) HCN2 −→ HCN + N 8.0×1013 140, 182

Molecular constants of HCNN adduct: ν1 = 388 cm-1 (degen=3.02) ν2 = 900 cm-1 (0.19)ν3 = 1895 cm-1 (2.79); molecular constants of HCN2 adduct: ν1 = 414 cm-1 (degen=2.02)ν2 = 1612 cm-1 (3.33) ν3 = 3998 cm-1 (0.65); σ = 3.49 A and ε/K = 300 for both adducts.

Rate coefficients: k1a Berman and Lin (1983); k−1a from microscopic reversibility (MR),based on Manaa et al. (1992), varied to fit falloff data; k1b same as k1a ; k−1b MR based onManaa et al. (1992), varied to fit falloff data; k2 TST, ∆S‡ = –11.6 J/K, Ea from Martinand Taylor (1993); k−2 MR; k3 MR with A−3 for O + HCN from Perry et al. (1984) andbarrier from Manaa and Yarkony (1991), varied to fit falloff data.

The input parameters for the chemical activation calculation are listed inTable 16. Possible product channels include

CH+N2 −→ HCNN (24a)

−→ HCN2 (24b)

−→ HCN+N (24c)

Although the exit channel is uphill, it is energetically much more accessible than inthe analogous CH2 + N2 case. Both adducts can form directly from the reactants.The isomerization barrier is sufficiently high to inhibit production of HCN +N unless the insertion complex can be formed directly. However, since CH isnormally considered to insert, direct reaction is reasonable. We adjusted the welldepth to fit the falloff data of Berman and Lin and Becker et al. The results ofthe QRRK calculations for total loss of reactants are compared to the availabledata in Fig. 53. The QRRK predictions reflect the experimental observations inshowing both pressure and temperature dependence at lower temperature. Thedominant product channel is formation of stabilized adduct(s). The QRRK resultsunderpredict the extent of stabilization when the well depths of Manaa and Yarkony

Page 99: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 223

are used. Setting the wells 42 kJ/mol deeper substantially increased the amountof stabilization; the results are then only slightly lower than the data of Bermanand Lin at 100 Torr and Becker et al. at 20 Torr. This well depth is consistent withthat assigned by Medhurst et al. and by Rodgers and Smith.

9

11

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

HCN2

HCN + N

HCNN

total

P = 1 atm N2

CH + N2 → Products

FIGURE 2.55. Arrhenius plot for the various channels of CH + N2−→ Products in1 atm N2 as predicted by QRRK using the 42 kJ/mol deeper wells.

At higher temperatures, the calculations converge as the endothermic channelto HCN + N becomes accessible. (Figure 55) This becomes the dominant channeland the rate coefficient no longer exhibits a negative temperature dependence.The calculations predict this dominance to set in at a somewhat lower temperaturethan observed in the experiments. This could be accomodated by adjusting thebarrier for surface crossing. In view of the uncertainties in the surfaces and theprobability for surface crossing, however, such adjustments are unwarranted. Theobserved kinetic behavior is in any event consistent with a general picture ofadduct formation and stabilization at low temperature, an initial rate decrease withtemperature as adduct dissociation becomes faster, and then a rate increase with

Page 100: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

224 Anthony M. Dean, Joseph W. Bozzelli

7

9

11

–2 0 2log (P/atm)

log

( k

/ cm

3 m

ol–1

s–1

)

HCN2

HCN + N

HCNN

total

T = 1500 K

CH + N2 → Products

FIGURE 2.56. Predicted effect of pressure at 1500 K from QRRK calculations (usingthe 42 kJ/mol deeper wells) for CH + N2−→ Products. Even at 1500 K, stabilizationbecomes more important than HCN + N above 3 atm.

temperature when the adduct has enough energy to make surface crossing andHCN + N formation significant. At very high pressures the stabilization channelis computed to continue to be important until significantly higher temperaturesthan studied by Berman and Lin.

These effects are illustrated in Figs. 55 and 56 for N2 as bath gas. (The adjustedwell depth was used for these calculations, as this gave much better agreementwith the stabilization data.) Pressure effects could be important when includingthis reaction in kinetic models of high pressure combustors; as shown in Fig. 56,stabilization is important at pressures above 1 atm even up to 1500 K.

The disagreement between the two most direct high temperature measurementsis disconcerting. Dean et al. measure a higher Ea as well as a faster rate, a factorof 2 at 3150 K, than Lindackers et al. As seen in Fig. 56, one expects no pressuredependence for the HCN + N channel. We suggest using the expression of Deanet al.

k24c = 4.4× 1012 exp(−11060/T) cm3mol−1s−1

and the QRRK estimates for the stabilization channel in cm3mol−1s−1 units, for

Page 101: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 225

300 to 2500 K in N2 bath gas

k24a = 2.3× 1027T−5.78 exp(−1230/T ) 0.1 atm

= 3.6× 1028T−5.84 exp(−1320/T ) 1 atm

= 1.8× 1030T−6.02 exp(−1735/T ) 10 atm.

At temperatures above 1000 K, reaction 24a is close to the low pressure limit forpressures of 10 atm or less (cf. Fig. 56 for 1500 K). The still lower stabilizationrate for formation of the cyclic complex can be ignored.

2.6.21 CH + NO −→ Products

The reaction of CH with NO is important in flame chemistry because CH, alongwith other CHx radicals, provides a path for NO destruction. Its contribution toNO destruction compared to other CHx radicals is an important question that stillneeds to be resolved. Experimental data on this reaction are shown in Fig. 57.Several groups have measured the rate at room temperature (Butler et al. 1981;Wagal et al. 1982; Lichtin et al. 1984; Okada et al. 1993), while two groups havemeasured the rate over extended ranges of low temperatures (Berman et al. 1982;Becker et al. 1993) . There is only one direct high-temperature measurement(Dean et al. 1991). All of these data present a consistent picture of a very rapidreaction with essentially no temperature dependence. Both Dean et al. and Okadaet al. attempted to observe products of this reaction. In both cases, NH wasconfirmed to be a minor product (less than 10% and less than 15%, respectively);OH was observed to contribute less than 30% and less than 20%, respectively.These observations led Dean et al. to suggest that the major product is probablyHCN + O, while Okada et al. were unable to rule out that NCO could be a majorproduct. Miller and Bowman (1989) propose HCN + O as the sole products, butpoint out that yet another possible channel is CN + OH. Later work by Lambrechtand Herschberger (1994),using CD + NO, indicated that the major product channelis DCN + O (47.5±12.2%), with NCO + D at 18.8±5.5%, CN + OD less than7.5%, and the sum of CO + DN and DCO + N at 33.7±13.8%.

Figure 58 shows the potential energy surface for reaction on the triplet sur-face. Addition of CH to NO forms 3HCNO, which can form HCN + O viaa very low exit barrier or isomerize to form 3NCHO. NCHO has three dis-sociation routes forming either N, O, or H and the corresponding triatomicpartner. One expects these routes, with large A-factors and low energy exitchannels, to compete favorably with isomerization to HNCO or HOCN. If iso-merization occurs, dissociation to H + NCO, NH + CO or OH + CN canfollow. This plethora of products are all lower in energy than the reactants.The HCN + O channel is expected to dominate, because the direct route tothese products from the first adduct has a low barrier and a high A-factor. Thepossible product channels encompass a rich variety of bond rearrangements:

Page 102: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

226 Anthony M. Dean, Joseph W. Bozzelli

13.8

14.0

14.2

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

Wagal et al. (1982) Butler et al. (1981) Lichtin et al. (1984) Berman et al. (1982) Dean et al. (1991) Okada et al. (1993) Becker et al. (1993) QRRK (total)

CH + NO → Products

FIGURE 2.57. Comparison of QRRK predictions to rate coefficient measurements for CH+ NO −→ Products.

CH+ NO −→ 3HCNO (25a)

−→ 3NCHO (25b)

−→ 3HNCO (25c)

−→ 3HOCN (25d)

−→ O+HCN (25e)

−→ H+NCO (25f)

−→ N+HCO (25g)

−→ NH+ CO (25h)

−→ OH+ CN (25i)

Page 103: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 227

Ent

halp

y (k

J/m

ol)

CH + NO

3NCHO3HNCO

3HOCN

HCO + N

OH + CN

HCN + O

NCO + H

NH + CO

3HCNO

300

500

700

FIGURE 2.58. Potential energy diagram for CH + NO.

Table 17 describes the input parameters for this reaction on the triplet surface.In spite of the complexity of this system, several qualitative features are evident.First, the very low exit barriers suggest very little stabilization, implying littleexpected pressure dependence. Second, for the same reason, one expects therate to be very fast. Third, one expects the distribution of products to be drivenprimarily by the A-factors for the unimolecular processes. Finally, one suspectsthe product distribution might be more complex than that used by Miller andBowman.

These results are compared to the experimental measurements in Fig. 57.Reasonable agreement is observed, although the calculations suggest more falloffthan one would infer from the data of Dean et al. Figure 59 illustrates the predictedrate coefficients for the different pathways as functions of temperature.

For clarity only the sums of the three pathways leading to H + NCO andthe two pathways to HCN + O are shown in Fig. 59. (The major pathway forHCN + O is from HCNO; the major pathway for H + NCO is dissociation fromNCHO, followed by HNCO and a small amount from HOCN.) In agreementwith the observations of Lambrecht and Herschberger, the most important channelis predicted to be HCN + O, followed by HCO + N and NCO + H, with thediatomic product channels predicted to be least important. Overall, the agreementof these predictions with the detailed product measurements of Lambrecht andHerschberger is good. The only differences are that they observe slightly less O +HCN, 48% versus the QRRK prediction of 60%, and more H + NCO, 18% versusthe QRRK prediction of 8%. Our predictions for the sum of the NH + CO and theHCO + N channels is 26% compared to the observed 34%, and our prediction ofabout 6% for the OH + CN channel is consistent with the upper limit measurement

Page 104: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

228 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.17. CH + NO −→ Products (Triplet)

Reaction A/s−1 or cm3mol−1s−1 Ea/kJ mol−1

(1) CH + NO −→ 3HCNO 1.1×1014 0(–1) 3HCNO −→ CH + NO 1.2×1016 226(2) 3HCNO −→ O + HCN 2.4×1013 10(3) 3HCNO −→ 3NCHO 3.7×1013 42

(–3) 3NCHO −→ 3HCNO 5.0×1013 272(4) 3NCHO −→ O + HCN 4.7×1013 192(5) 3NCHO −→ H + NCO 3.6×1013 170(6) 3NCHO −→ 4N + HCO 1.4×1015 290(7) 3NCHO −→ 3HNCO 8.1×1013 91

(–7) 3HNCO −→ 3NCHO 7.8×1013 67(8) 3HNCO −→ NH + CO 1.4×1013 21(9) 3HNCO −→ H + NCO 3.4×1013 143

(10) 3NCHO −→ 3HOCN 8.1×1013 205(–10) 3HOCN −→ 3NCHO 7.2×1013 56

(11) 3HOCN −→ OH + CN 2.5×1015 97(12) 3HOCN −→ H + NCO 7.7×1013 5

Molecular constants of 3HCNO adduct: ν1 = 417 cm{1 (degen=2.12) ν2 = 1197 cm{1

(2.54) ν3 = 2924 cm{1 (1.33); molecular constants of 3NCHO adduct: ν1 = 722 cm{1

(degen=2.75) ν2 = 1724 cm{1 (2.86) ν3 = 3990 cm{1 (0.39); molecular constants of3HNCO adduct: ν1 = 396 cm{1 (degen=2.14) ν2 = 1228 cm{1 (2.39) ν3 = 3218 cm{1

(1.47); molecular constants of 3HOCN adduct: ν1 = 367 cm{1 (degen=1.78) ν2 = 924cm{1 (1.80) ν3 = 2776 cm{1 (2.42); σ = 3.8 A and ε/K = 232 for all adducts.

Rate coefficients: k1 — 0.75 of the value measured by Berman and Lin (1982) and Beckeret al. (1992b) to account for formation of the triplet product; k−1 from microscopicreversibilty (MR); k2 from A−2 = 6.0×1012, Mallard et al. (1993), Ea(2) assigned to beconsistent with the reaction exothermicity; k3 TST, ∆S‡ = –4.1 J/K, Ea based on three-member ring closure; k−3 MR; k4 from k−4 and MR with k−4 = 6.0×1012 exp(–3000/T )from Perry and Melius (1984); k5 from k−5 and MR with k−5 based on H + C2H5 fromTsang and Hampson (1986); k6 from k−6 and MR with k−6 taken as 4.0×1013, half therate coefficient used for N + CH3 adduct formation (cf. Table 16); k7 TST,∆S‡ = 4.1 J/K,partial gain of rotor, Ea for carbene H-shift; k−7 MR; k8 from k−8 and MR with k−8 =5.19×1011exp(–3250/T), based on 3CH2 + CO and CH3 + CO, cf. Anastasi et al. (1982);Ea agrees with a BAC/MP4 calculation of C.F. Melius (1993); k9 via k−9 and MR—cf.k5; k10 TST,∆S‡ = 4.1 J/K, partial gain of rotor, Ea based on carbene H–shift; k−10 MR;k11 via k−11 and MR with k−11= 4.0×1013 (Average of data reported by Mallard et al.1993); k12 via k−12 and MR—cf. k5.

Page 105: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 229

13

14

1 2 31000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

N + HCO

NH + CO

OH + CN

H + NCO (total)

O + HCN (total)

total

CH + NO → Products

FIGURE 2.59. Arrhenius plot for the various channels of CH + NO −→ Products aspredicted by QRRK calculations for 1 atm N2 buffer gas. The low energy exit channelsmake stabilization unimportant for this reaction; similarly, the effect of temperature is quitesmall.

of less than 7.5%. As discussed earlier, these results confirm that the energy ofthe exit channel is not necessarily the main factor influencing the rate. Here thehighest energy exit channel, HCO + N, is predicted to be much more importantthan the lowest, NH + CO. In this case, the higher barrier for HCO + N (272kJ/mol higher than for NH + CO) is compensated by a much higher A-factor. Theadditional isomerization required to produce NH + CO lowers the rate along thispath. The diminished role of energetic considerations in this reaction is made morepronounced by the low energies of all channels relative to the entrance channel.

Calculations over an extended pressure range confirm that pressure effectsshould be minimal for this system. The dissociation channels are found todominate up to pressures over 100 atm even at 300 K. Thus one does not expect

Page 106: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

230 Anthony M. Dean, Joseph W. Bozzelli

pressure to influence the branching ratios for this reaction.Based on these considerations we suggest for the total rate coefficient

k25(total) = 1.1× 1014 cm3mol−1s−1, from 300 to 3800 K,

in agreement with all the measurements. Following the observations of Lambrechtand Herschberger we suggest that the following branching ratios be used:

(O+HCN) k25e/k25(total) = 48%

(H+ NCO) k25f/k25(total) = 18%

(N+HCO) k25g/k25(total) = 26%

(NH+ CO) k25h/k25(total) = 5%

(OH+ CN) k25i/k25(total) = 3%.

The branching ratios vary only slightly with temperature, as can be seen fromFig. 59, and so the ones listed can be applied at all temperatures.

We also examined reactions on the singlet surface. The slower overall rateof adduct formation leads to lower rate coefficients for all channels. Two newproduct channels become possible, namely 1NH + CO and OH + CN; these areexpected to contribute less than 10% to the overall reaction.

2.7 OTHER REACTIONS OF INTEREST

In this section we consider other reactions that are important in high temperaturenitrogen chemistry. While most of them have been included in earlier nitro-gen chemistry mechanisms, we also discuss reactions of several species, e.g.,NH2O, H2NN and H2CN, that have not been included before. The discussionsare generally brief; appropriate references to the literature are given for thoseseeking further information. Our intent is to identify as many of the remainingpotentially important reactions as possible and to collect their rate coefficients—for those reactions for which recommendations can be made—in one place for theconvenience of the combustion modeling community.

2.7.1 Reactions of N-atoms

One expects N-atoms to participate in adduct formation reactions, as opposed tohydrogen abstractions, because the N–H bond dissociation energy is only about340 kJ/mol, meaning that hydrogen abstractions by N-atoms from molecules suchas hydrocarbons would be endothermic.

(a) N + O2 −→ NO + O.

This reaction was reviewed by Cohen (1992), who recommended

k26a = 9.0× 109T exp(−3270/T ) cm3mol−1s−1,

Page 107: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 231

close to the expression obtained using the Hanson-Salimian expression for thereverse reaction rate coefficient.

(b) N + OH −→ NH + O.

Cohen and Westberg (1991) propose

k26b = 6.4× 1012T 0.1 exp(−10700/T ) cm3mol−1s−1

based upon an estimated expression for the reverse (abstraction) reaction 27e2.The high barrier assigned to this reaction is consistent with its 92 kJ/molendothermicity. We suggest that this rate coefficient be calculated from thereverse reaction rate coefficient.

(c) N + OH −→ NO + H.

Natarajan et al. (1994) measured the reverse reaction rate behind shock wavesover the range 2560 to 4040 K. A forward rate coefficient expression wasobtained by fitting their data to an Arrhenius expression, invoking microscopicreversibility, and combining the result with the low-temperature recommenda-tion of Atkinson et al. (1989) to obtain

k26c = 1.1× 1014 exp(−565/T ) cm3mol−1s−1.

Essentially the same result was obtained by Qin et al. (1997) in a similarinvestigation. This channel is much faster than the NH + O channel, which inview of its 205 kJ/mol exothermicity is not surprising.

(d) N + CH −→ 3HCN −→ H + CN.

This channel is exothermic by 415 kJ/mol, sufficient to convert the initially-formed 3HCN to the ground state singlet, which subsequently dissociates toH + CN. The rate coefficient at room temperature was reported by Messing etal. (1981) to be 1.3× 1013 cm3mol−1s−1. Brownsword et al. (1996) measuredthe rate from 216 to 584 K and reported

k26d = 1.7× 1014 T−0.09 cm3mol−1s−1.

(e) N + CH2 −→ H2CN −→ HCN + H.

This is another exothermic reaction having an exit channel with energy lowerthan that of the initially formed adduct. We are not aware of any measurements.Miller and Bowman use

k26e = 5.0× 1013 cm3mol−1s−1,

which is close to the average of the rate coefficients for N + CH (reaction 26d)and N + CH3 (cf. Section 6.16).

Page 108: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

232 Anthony M. Dean, Joseph W. Bozzelli

(f) N + NH −→ N2 + H.

This reaction may be expected to proceed as formation of the NNH adductfollowed by rapid dissociation to N2 + H. Hack et al. (1994) studied it at 298 Kand reported

k26f = 1.5× 1013 cm3mol−1s−1.

(g) N + NH2 −→ N2 + H + H.

Another exothermic adduct dissociation, perhaps with NNH as a second inter-mediate dissociating to N2 + H, for which the rate coefficient is expected to belarge. This expectation is confirmed by two studies (Whyte and Phillips 1983a;Dransfeld and Wagner 1987) giving at 298 K

k26g = 7.1× 1013 cm3mol−1s−1.

(h) N + CN −→ C + N2.

Although Baulch et al. (1992) suggest 1.8 × 1014 cm3mol−1s−1 for this re-action, the room temperature measurement was only 6 × 1013 cm3mol−1s−1

(Whyte and Phillips 1983b). Dean et al. (1990) measured the reverse reactionrate in shock tube experiments and inferred the forward rate constant to be about2.6× 1013 cm3mol−1s−1 with little temperature dependence. Combining thetwo values gives

k26h = 2.4× 1013 exp(+280/T ) cm3mol−1s−1 .

2.7.2 Reactions of NH

Hydrogen abstraction from hydrocarbons by NH is expected to be slow, becausethe strength of the H–NH bond formed is only about 380 kJ/mol, meaning thatmost abstractions would be endothermic. As with N, adduct formation pathwaysare expected to be more important.

(a) NH + NH −→ Products.

This reaction was studied behind shock waves by Mertens et al. (1989), whoreported

k27a = 5.1× 1013 cm3mol−1s−1,

consistent with an exothermic combination to form N2H2 followed by rapiddissociation to NNH + H, with likely final products N2 + H + H.

Page 109: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 233

(b) NH + NH2 −→ N2H2 + H.

The most likely channel is formation of the N2H3 adduct followed by dissocia-tion to form N2H2 + H. The overall process is exothermic by 113 kJ/mol. Theshock tube experiments of Davidson et al. (1990b) were interpreted to yield

k27b1 = 1.5× 1015T−0.5 cm3mol−1s−1,

consistent with the room temperature measurement of Dransfeld et al. (1984)and with the expected rate for a process with an adduct formation mechanism.Adduct dissociation to form H2NN + H is less likely, since this channel is only33 kJ/mol exothermic. Another possibility would be the H abstraction to formN + NH3, exothermic by 113 kJ/mol. The DHT estimation method of Section4 gives

k27b2 = 9.2× 105 T 1.94 exp(−1230/T ) cm3mol−1s−1 .

(c) NH + OH −→ Products.

Cohen and Westberg (1991) reviewed this reaction. No data are available; theyestimate rate coefficients by analogy to OH + OH and suggest

NH+ OH −→ H+HNO, k27c1 = 2.0× 1013 cm3mol−1s−1,

consistent with the 71 kJ/mol exothermicity. Another pathway is direct hydro-gen transfer

NH+ OH −→ H2O+N ,

for which Cohen and Westberg suggest 2.0× 109T 1.2 cm3mol−1s−1. This issomewhat larger than the DHT estimate

k27c2 = 1.2× 106 T 2 exp(+245/T ) cm3mol−1s−1 .

(d) NH + H −→ H2 + N.

Baulch et al. (1992) suggest the value 1× 1013 cm3mol−1s−1, consistent withan exothermic direct abstraction reaction but not with an adduct formationpathway; adduct formation and dissociation would involve 2NH2, from whichdissociation to 4N + 1H2 is spin-forbidden. Davidson and Hanson (1990a)measured the reverse reaction rate from 1950 to 2850 K. Combining their resultwith the equilibrium constant gives

k27d = 3.5× 1013 exp(−870/T ) cm3mol−1s−1 ,

which is higher than the Baulch et al. expression at combustion temperatures.

Page 110: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

234 Anthony M. Dean, Joseph W. Bozzelli

(e) NH + O −→ Products.

Cohen and Westberg (1991) reviewed this reaction and suggested on the basisof collision theory arguments that k = 3.0× 1013 cm3mol−1s−1 for the H +NO channel, which is consistent with a very exothermic (–289 kJ/mol) adductformation and dissociation process. For the hydrogen transfer channel to formOH + N they suggest k = 3.0 × 1012 cm3mol−1s−1. Hack et al. (1994),however, report that the rate coefficient for this channel is less than 1.0× 1011

at 298 K. The DHT estimate is

k27e2 = 1.7× 108 T 1.5 exp(−1695/T ) cm3mol−1s−1.

This expression gives 3 × 109 cm3mol−1s−1 at 298 K, consistent with theHack et al. measurement, and 6 × 1012 cm3mol−1s−1 at 2000 K, close tothe Cohen and Westberg estimate. Sengupta and Chandra (1994) concludedon the basis of electronic structure and QRRK calculations that reaction 27e1proceeds by intermediate adduct formation, while both direct abstraction andadduct formation are significant for 27e2.

There have been several measurements of the total rate coefficient. Dransfeldet al. (1984) reported (5±2)×1013 cm3mol−1s−1 and Adamson et al. (1994)reported 4.2× 1013 cm3mol−1s−1 at room temperature. Mertens et al. (1991)reported 9.2×1013 cm3mol−1s−1 for temperatures between 2700 and 3400 K.Because the H-transfer channel is expected to be appreciably slower than thetotal for these measurements, we suggest that they should be assigned to the H+ NO channel and recommend

k27e1 = 6× 1013 cm3mol−1s−1.

(f) NH + CH3 −→ HNCH3 −→ H2CNH + H.

This reaction is overall exothermic by 188 kJ/mol, suggesting that its ratecoefficient should be similar to that for adduct formation. Using as an analogthe Baulch et al. (1992) expression for OH + CH3 we estimate

k27f = 4.0× 1013 cm3mol−1s−1.

Direct H-atom transfer is estimated to be at least two orders of magnitudeslower.

2.7.3 Reactions of NNH

We discussed NNH in Section 6.14 as a dissociation product of the cis- or trans-HNNOH isomers of the initially-formed adduct of the NH2 + NO reaction.

(a) NNH −→ N2 + H.

NNH is thought to be weakly bound and thus to dissociate rapidly; calculationsby Koizumi et al. (1991) indicate a tunneling lifetime of only about 3 ns

Page 111: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 235

for the ground vibrational state. The pressure dependence of the dissociationrate can be estimated by means of a simplified QRRK analysis. The highpressure rate coefficient for NNH dissociation was estimated from the reversereaction rate coefficient using an A-factor for H addition to N2 comparable toH addition to C2H2 and a barrier that gives a room temperature value for thehigh pressure rate coefficient consistent with the calculated tunneling lifetimefor the ground vibrational state. This barrier proved to be 20 kJ/mol. (It hadto be lower than the 31 kJ/mol calculated by Koizumi, as no attempt had beenmade to correct the QRRK rate coefficient for tunneling.) The subsequentQRRK analysis suggested that this reaction is at the low pressure limit forconditions of interest in combustion, with a rate coefficient given by k =1.3× 1014T−0.11 exp(−2500/T ) cm3mol−1s−1for N2 bath gas. At 1 atm theeffective first-order rate coefficient is computed to be 6×107 s−1 at 1000 K and1× 108 s−1 at 2000 K, the small temperature effect being due to the unusuallylow barrier. These values are much lower than the estimated high-pressure limitof over 1011 s−1 at these temperatures and indicate that this reaction is deep inthe falloff region. However, they are still much higher than the 104 s−1 valueused by Miller and Bowman (1989). We are not able to reconcile the Millerand Bowman value theoretically.

The QRRK treatment of NNH dissociation assumed that dissociation occursby thermal activation, the only consideration of tunneling being adoption ofa lowered dissociation barrier. An alternative approach is to use the energy-dependent tunneling rate coefficients computed by Koizumi et al. to estimate thedissociation rate directly. Two contributions have to be included: (1) A thermalactivation one with a 31 kJ/mol barrier height, corresponding to the calculationof Koizumi et al., and (2) a tunneling one at the rate calculated by Koizumi etal. The high pressure rate coefficient that results can be represented as k∞ =4.1×109T 1.13 exp(−2610/T ) s−1, with the tunneling contribution dominatingat all temperatures. At lower pressures, the analysis is complicated, becausethe pressure-independent term due to tunneling from the lowest vibrationalstate should be the lower limit to the rate coefficient. An approximate analysissuggests that the dissociation rate coefficients from the higher levels are fasterthan collisional stabilization rates; in other words, the reaction is near its low-pressure limit. Assuming that the thermal activation rate is approximatelykcoll [M] leads to

k28a ≈ 3× 108 + 1× 1013T 1/2 exp(−1540/T )[M] s−1,

where the collisional rate coefficient term represents activation to the first excitedvibrational state, which gives a dissociation rate coefficient of about 1.2 ×109 s−1 at 1 atm and 1000 K and 1.6 × 109 s−1 at 2000 K, substantiallylarger than the QRRK estimate given above. We recommend this expressionfor combustion modeling. This assignment is consistent with the study ofThermal DeNOx kinetics using isotopically labelled nitrogen by Duffy andNelson (1996), who report the lifetime of NNH to be less than 10 ns.

Page 112: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

236 Anthony M. Dean, Joseph W. Bozzelli

The ambiguity in rate coefficient assignment for this reaction caused by the un-certain tunneling contribution is mitigated by the consideration that if tunnelingis important it must be equally important for forward and reverse directions.Thus the net effect of tunneling is that this reaction is likely to be equilibratedunder most conditions of interest in combustion, so that the concentration ofNNH is essentially independent of the rate coefficient for dissociation. This isan important result—despite its large dissociation rate, NNH concentrationsin flames prove not to be negligibly small, particularly in air, where thelarge N2 concentration enhances the reverse reaction rate and can drive NNHconcentrations to near-equilibrium levels. (For flames in air at 1800 K and 1atm the equilibrium [NNH]/[H] ratio is about 2× 10−6; at 1200 K and 1 atmit is about 1 × 10−6.) There are therefore conditions where the equilibriumNNH concentration is high enough for it to participate in bimolecular reactions.In general, one expects that NNH is more likely to be equilibrated in hightemperature systems in air, where the fast rate of NNH formation quickly leadsto a near-equilibrium concentration.

One can estimate the importance of various reaction partners by estimatingvalues of kx [X], where kx is the rate coefficient for species X reacting withNNH. We consider such reactions with O2 and the dominant flame radicalsH, OH and O, which can attain mole fractions up to several percent in flamefronts, and with NH2, HO2 and NO, which are important under Thermal DeNOxconditions.

(b) NNH + O2 −→ Products.

The potential reactant at highest concentration is usually O2. A QRRK approachcan be used to estimate a rate coefficient for it. The HNNOO adduct formed candissociate back to reactants or undergo intramolecular hydrogen transfer to formNNOOH, which will dissociate to either HO2 + N2 or N2O + OH. The uncertainHNNOO well depth is probably somewhere in the range 70–90 kJ/mol. TheA-factor for hydrogen transfer, where the transition state involves loss of the O2rotation degree of freedom, is estimated to be about 4×1012 cm3mol−1s−1. Wechose an isomerization barrier of 88 kJ/mol, as used by Lin et al. (1992), and a92 kJ/mol well depth. (Assuming a shallower well significantly reduces the rateof product formation.) Partition of NNOOH to the two product channels wasestimated at 4:1, favoring HO2 + N2, from estimated transition state entropies.These calculations give in cm3mol−1s−1 units

NNH+O2 −→ HO2+N2, k28b1 = 1.2× 1012T−0.34 exp(−75/T )

NNH+O2 −→ N2O+ OH, k28b2 = 2.9× 1011T−0.34 exp(−75/T ) .

The value for the faster HO2 + N2 channel is only about 9×1010 cm3mol−1s−1

at 2000 K, suggesting a pseudo-first-order rate coefficient of about 1× 105 s−1

in air at 1 atm for the reaction of NNH with O2.

This analysis indicates that NNH + O2−→N2 + HO2 is generally too slow to beof interest. At first glance this is surprising, given that the reaction appears to

Page 113: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 237

be a simple exothermic hydrogen abstraction. Extensive analysis of analogousreactions shows, however, that it is not direct hydrogen transfer but rather theresult of an addition–isomerization–elimination sequence. Reactions of theform R + O2 −→ olefin + HO2 have received considerable attention, and thereis a growing consensus that they are not direct reactions. Kaiser (1995), forexample, showed that the ethylene yield from C2H5 + O2 decreases dramaticallywith increasing pressure. He interpreted this in terms of adduct formation, withthe increasing pressure stabilizing the energized adduct prior to isomerization.In theoretical treatments of such reactions (Bozzelli and Dean 1990; Wagner etal. 1990; Bozzelli and Dean 1993) the available data were reconciled in termsof adduct formation. Thus for the NNH + O2 reaction, even though the ratecoefficient for formation of the HNNOO adduct is large and the pathway to HO2+ N2 is exothermic, the rate limiting step is isomerization; the relatively shallowwell for adduct formation puts the barrier for isomerization only 4 kJ/mol lowerthan the entrance channel. Since the A-factor for redissociation to reactants ishigher than that for isomerization, dissociation to NNH + O2 dominates andleads to a low rate of N2 + HO2 formation.

(c) NNH + H −→ Products.

Recombination of H with NNH would form HNNH, while addition could formH2NN. NH production by breaking the N–N bond in HNNH is too endothermicto be important. Formation of the HNNH adduct liberates about 250 kJ/mol,which is probably less than the barrier for H2 elimination to form N2 andH2; in any event, the A-factor for this process should be much lower than forredissociation of the adduct to H + NNH. Stabilization of HNNH or H2NN wasalso found to be relatively unimportant.Another possibility is the direct reaction

NNH+H −→ H2 +N2. (28c).

for which the DHT estimate is

k28c = 2.4× 108 T 1.5 exp(+450/T ) cm3mol−1s−1 .

For an H atom mole fraction of 1% the pseudo-first-order rate coefficient of thisreaction at 2000 K and 1 atm would be about 2×106 s−1, substantially fasterthan reaction with 20% O2.

(d) NNH + OH −→ Products.

NNH reaction with OH to form HONNH is likely to form N2 + H2O viaconcerted elimination of H2O on the same potential energy surface consideredearlier (Section 6.14) for NH2 + NO. We analyzed this with the QRRK methodusing the parameters in Table 10. In the present case the surface is entered viaOH + NNH to form trans HNNOH. The analysis leads, in cm3mol−1s−1 units,to

NNH+OH −→ N2 + H2O, k28d1 = 2.4× 1022T−2.88 exp(−1230/T ).

Page 114: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

238 Anthony M. Dean, Joseph W. Bozzelli

At 2000 K the value of this expression is 4×1012 cm3mol−1s−1, significantlylower than the 5 × 1013 cm3mol−1s−1 value used by Miller and Bowman(1989). At lower temperatures, its value is higher, the decline with temperaturebeing due to normal fall-off effects, i.e., the complex dissociates faster. For anOH mole fraction of 1% the pseudo-first-order rate coefficient at 2000 K forthis reaction is about 2× 105 s−1.Direct hydrogen transfer can also lead to these products. The DHT estimate forits rate coefficient is

k28d2 = 1.2× 106 T 2 exp(+600/T ) cm3mol−1s−1 .

At 2000 K this expression is 6.5×1012 cm3mol−1s−1. At 300 K it is muchsmaller than k28d1. We suggest using both expressions together to reflect thechange in mechanism with temperature. The sum of both rate coefficients isstill about 5 times smaller than the Miller-Bowman value.

(e) NNH + O −→ Products.

The NNH + O reaction has three product channels, two of which have alreadybeen discussed as reverse reactions. The ONNH complex can dissociate toform H + N2O (the reverse of reaction 11d) or to NO + NH (the reverse ofreaction 11g). On the basis of the discussion in Section 6.7 we expect bothchannels to be fast, because both are lower in energy than NNH + O; we expecta low A-factor for the H + N2O channel to inhibit this pathway relative toNO + NH even though it is more exothermic. This is seen in the QRRK ratecoefficients, k−11g being about an order of magnitude larger than k−11d. At2000 K, k−11g is calculated to be about 7 × 1013 cm3mol−1s−1, leading toa pseudo-first-order rate coefficient of 4 × 106 s−1 for 1% O–atoms in air.The QRRK estimate in cm3mol−1s−1 units for the rate coefficient of the third(minor) channel producing OH + N2

NNH+O −→ N2+OH, k28e1 = 1.7× 1016T−1.23 exp(−250/T )

is about 1× 1012 cm3mol−1s−1 at 2000 K, significantly less than that for theNH + NO channel.Direct hydrogen transfer would give the same products. The DHT estimate forits rate coefficient is

k28e2 = 1.7× 108 T 1.5 exp(+450/T ) cm3mol−1s−1,

which is appreciably faster than the adduct formation pathway at high temper-atures. We accordingly suggest again that both rate coefficient expressions beused to reflect the change in mechanism with temperature.

In Section 6.7 we showed that NNH + O is a plausible high temperature channelof the NH + NO reaction and that the calculated temperature dependence for thischannel is consistent with the high temperature measurements of Mertens et al.

Page 115: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 239

(1991a). Specifically, Mertens et al. measured an activation energy of 54 kJ/molfor the total rate coefficient for the NH + NO reaction over the temperature range2220 to 3350 K. The NH + NO reaction has three likely product channels: H +N2O, N2 + OH and NNH + O. Only the latter is endothermic (1r H = 52 kJ/mol),and we expect the other two channels to have rate coefficients that decrease withtemperature (cf. Fig. 16). Thus the measured temperature dependence suggeststhat NNH + O is the dominant channel in the Mertens experiments. Acceptingthis assumption, their data imply, through microscopic reversibility, that k−11g isabout 1.3× 1014 cm3mol−1s−1, a factor of 2 above the QRRK estimate. Millerand Melius (1992) pointed out that this channel might be responsible for theobserved temperature dependence. The Mertens et al. experiments offer indirectbut convincing evidence that the reaction between NNH + O can be very rapid.The QRRK rate coefficient estimate being in reasonable agreement with their datasupports this conclusion.

(f) NNH + NH2 −→ Products.

The DHT estimate is

k28f = 9.2× 105T 1.94 exp(+580/T ) cm3mol−1s−1 ,

which is comparable to the H atom transfer reaction OH + NNH but slower thanthose of H and O.

(g) NNH + HO2 −→ Products.

The DHT estimate for H2O2 production is, in cm3mol−1s−1 units

NNH+ HO2 −→ N2 +H2O2, k28g1 = 1.4× 104T 2.69 exp(+805/T ) .

Adduct formation followed by dissociation to HNNO + OH is endothermic by8 kJ/mol. Assuming that the A-factor for this channel is comparable to that forredissociation to reactants gives, in cm3mol−1s−1 units,

NNH+HO2 −→ HNNO+OH, k28g2 = 2.4× 1013 exp(−855/T ) .

Page 116: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

240 Anthony M. Dean, Joseph W. Bozzelli

(h) NNH + NO −→ N2 + HNO.

We estimate

k28h = 1.2× 106T 2 exp(+600/T ) cm3mol−1s−1

by assuming the reaction with NO to be analogous to that with OH. Eventhough the the H–NO bond is about 290 kJ/mol weaker than the H–OH bond,abstraction by NO is still exothermic by about 230 kJ/mol.

The O + NNH −→ NH + NO reaction thus appears to be the most importantof the bimolecular NNH reactions for combustion, with a pseudo-first-order ratecoefficient at least an order of magnitude larger than the other bimolecular reactionsand the only one that has a rate of product formation comparable to that for adductformation. Roles of this reaction in NOx production in methane–air and hydrogen–air systems are discussed in Section 8 and by Bozzelli and Dean (1995).

No pressure dependence is expected for any of the radical reactions with NNH.Our conclusions that NNH can be partially equilibrated during combustion

processes and that its bimolecular reactions are important suggests that additionalattention should be given to them. Especially important are the reactions withOH, H and O2. Our conclusion that the rate coefficients of these reactions aresubstantially smaller than those of the reaction with O is based on analogies toother reactions. These analogies should be tested by theoretical and experimentalstudies.

2.7.4 Reactions of N2H2 (HNNH)

N2H2 is unusual in that its H–N bond dissociation energy is only 258 kJ/mol,quite small for a stable molecule. One therefore expects it to undergo hydrogentransfer reactions, and the NNH abstraction product to dissociate quickly to N2 +H. In some respects N2H2 might be expected to react like HNO, where the H–Nbond dissociation energy is also low, 210 kJ/mol. Addition reactions to N2H2are probably not important, as the adducts appear to have shallow potential wells.Consequently, the isomerization barriers are above the entrance channel, and theoverall rates to non-adduct products are slow.

(a) N2H2 + M −→ NNH + H + M.

To our knowledge there are no measurements of the dissociation rate of N2H2.The same products are accessible both by direct dissociation and via isomeriza-tion to H2NN followed by its dissociation. Because the isomerization barrieris not known, we estimated it by the G2 method (Foresman and Frisch 1995),which gave barriers in the 190 to 230 kJ/mol range. Based on this resultit appears that both dissociation pathways contribute. The following QRRKestimates, in s−1 units, apply to dissociation in N2 bath gas in the temperaturerange from 600 to 2500 K.

Page 117: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 241

k29a1 = 5.6× 1036 T−7.75 exp(−35400/T ) 0.1 atm

= 1.8× 1040 T−8.41 exp(−36900/T ) 1.0 atm

= 3.1× 1041 T−8.42 exp(−38300/T ) 10 atm

k29a2 = 1.6× 1037 T−7.94 exp(−35600/T ) 0.1 atm

= 2.6× 1040 T−8.53 exp(−36700/T ) 1.0 atm

= 1.3× 1044 T−9.22 exp(−38800/T ) 10 atm.

(b) N2H2 −→ H2NN.

The isomerization rate coefficients estimated by QRRK, again for nitrogen bathgas and over the temperature range from 600 to 2500 K, are, in s−1 units,

k29b = 9.2× 1038 T−9.01 exp(−34100/T ) 0.1 atm

= 2.0× 1041 T−9.38 exp(−34400/T ) 1.0 atm

= 1.3× 1045 T−10.13 exp(−35600/T ) 10 atm.

These results suggest that direct dissociation is faster than isomerization aboveabout 700 K at 1 atm.

(c) N2H2 + X −→ NNH + HX, with X = H, OH, O, NH2 or CH3.

We are not aware of any experimental measurements for these reactions. DHTestimates are included in Table 19. Linder et al. (1996) reported variationaltransition state calculations of the rate coefficients for the reactions with H,OH and NH2. Their expression for abstraction by H agrees quite well with theDHT estimate for the entire 300–2000 K range, while their OH and NH2 ratecoefficients are generally lower.

(d) N2H2 + NH −→ NNH + NH2.

Although appreciably less exothermic than the reaction with OH, the reactionwith NH is still exothermic by about 160 kJ/mol. The same rate coefficient aswe assigned to abstraction by OH is included in Table 19.

(e) N2H2 + NO −→ N2O + NH2.

This addition reaction can be considered to be similar to NO addition to HNO(Section 7.11). The A-factor is expected to be twice that for HNO addition.The adduct would be expected to undergo an H-shift with a lower barrier thanthat for HNO. This reasoning leads to

k29e = 4.0× 1012 exp(−6000/T ) cm3mol−1s−1.

Page 118: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

242 Anthony M. Dean, Joseph W. Bozzelli

2.7.5 Reactions of H2NN

This species is expected to be a major product of the NH2 + NH2 reaction (Section6.13). Like other singlet biradicals, e.g. 1CH2, H2NN is expected to undergoinsertion reactions. Its thermochemistry suggests that substituted hydrazines soformed will not have sufficient energy to dissociate into fragments.

1H2NN should not be considered as reactive as 1CH2, because 1H2NN isstabilized by the overlap of the nitrogen lone pair orbitals. This is reflectedin the shortening of the N–N bond length in the series N2H4 (1.43 A), NH2–NH(1.31 A), 1H2NN (1.18 A) and HN=NH (1.13 A). This stabilization suggests that1H2NN should be less reactive than 1CH2, i.e., more like pi-bonded species suchas C2H4 or carbonyls. Molecular electronic structure calculations were carriedout to characterize the transition states for reactions between radicals and 1H2NNand from them the corresponding rate parameters. Semi-empirical calculationswith the PM3 method indicate transition states and A-factors similar to radicaladdition reactions.

Reaction of 1H2NN with radicals can occur at two sites, addition betweenthe two N atoms or at the terminal N atom. The adducts formed often have anexothermic pathway in which beta-scission leads to formation of a double bond.Unimolecular reaction via this pathway is generally faster than stabilization undercombustion conditions.

(a) H2NN −→ Products.

QRRK calculations indicate that isomerization to the lower energy isomerHNNH (N2H2) dominates below 1500 K. This reaction has already beendiscussed in the reverse direction. As with the HNNH isomer, isomerizationmay or may not precede dissociation. The QRRK results for the two pathways,valid for dissociation in N2 bath gas from 600 to 2500 K, are in s−1 units,

k30a1 = 5.9× 1032 T−6.99 exp(−26100/T) 0.1 atm

= 9.5× 1035 T−7.57 exp(−27600/T) 1.0 atm

= 5.0× 1036 T−7.43 exp(−28800/T) 10 atm

k30a2 = 7.2× 1028 T−5.77 exp(−25500/T) 0.1 atm

= 3.2× 1031 T−6.22 exp(−26300/T) 1.0 atm

= 5.1× 1033 T−6.52 exp(−27300/T) 10 atm.

(b) H2NN + O2 −→ Products.

In this case adduct formation does provide sufficient energy to facilitate disso-ciation. Direct dissociation would produce NH2 + NO2, while isomerizationwould lead to HNN(O)OH, which could dissociate to H2O and N2O or to HNNOand OH. QRRK calculations suggest that the primary dissociation dominates

H2NN+ O2 −→ NH2 +NO2 (30b)

Page 119: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 243

with a rate coefficient given by the insertion

k30b = 1.5× 1012 exp(−3000/T ) cm3mol−1s−1

The A-factor is that measured by Laufer et al. (1983) for the reaction ofvinylidene with hydrogen. The dissociation barrier was estimated to be slightlylower than used for the H2NN + H2 channel of the NH2 + NH2 reaction.Although this barrier seems reasonable, additional experiments and theoreticalcalculations are needed. The rate of reaction 30b is greater than that of 30a inair at 1 atm for temperatures below 1900 K.

(c) H2NN + H −→ Products.

Both adduct formation and direct H-atom transfer have to be considered. Adductformation followed by exothermic dissociation to HNNH + H has the effectof isomerization. The A-factor, in cm3mol−1s−1 units, was estimated fromcalculated transition state geometries as described above.

H2NN+ H −→ N2H2 +H k30c1 = 1.8× 1010 T 0.97 exp(−2250/T ) .

The DHT estimate for the direct reaction, again in cm3mol−1s−1 units, is

H2NN+H −→ H2 +NNH k30c2 = 4.8× 108 T 1.5 exp(+450/T ) .

The direct reaction dominates.

(d) H2NN + O −→ Products.

The NH2NO adduct formed in this reaction is highly energized. QRRK analysisbased on the potential energy surface used for the NH2 + NO reaction (Figure33) shows that the H2NN + O reaction leads almost exclusively to NH2 andNO. The higher A-factor for this channel, relative to that for isomerization,governs the branching at such a high level of excess energy. The rate coefficientis therefore that of adduct formation, which we estimated from the calculatedtransition state geometries to be, in cm3mol−1s−1 units,

H2NN+ O −→ NH2 +NO k30d1 = 3.2× 109 T 1.03 exp(−1360/T ) .

The DHT estimate, likewise in cm3mol−1s−1 units,

H2NN+O −→ OH+ NNH k30d2 = 3.3× 108 T 1.5 exp(+450/T ) ,

is faster.

(e) H2NN + OH −→ Products.

Adduct formation followed by breaking of the O–H bond leads to NH2NO +H with an exothermicity of 54 kJ/mol. We estimated its rate coefficient on thebasis of a transition state calculation done with the PM3 semi-empirical method.

H2NN+ OH −→ NH2NO+H k30e1 = 2× 1012 cm3mol−1s−1.

Page 120: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

244 Anthony M. Dean, Joseph W. Bozzelli

The DHT estimate is, in cm3mol−1s−1 units,

H2NN+OH −→ H2O+ NNH k30e2 = 2.4× 106 T 2 exp(+600/T ) .

(f) H2NN + CH3 −→ Products.

Adduct formation can be followed by dissociation into two exothermic channels,one breaking a C–H bond and one breaking an N–H bond. The transition statetheory calculation leads to the rate coefficient expressions, in cm3mol−1s−1

units,

H2NN+ CH3 −→ H2NNCH2+H k30f1 = 8.3× 105 T 1.93 exp(−3270/T )

−→ HNNCH3+H k30f2 = 8.3× 105 T 1.93 exp(−3270/T ) .

The DHT rate coefficient expression, also in cm3mol−1s−1 units, is

H2NN+ CH3 −→ CH4 + NNH k30f3 = 1.6× 106 T 1.87 exp(−65/T ) ,

which is substantially faster than adduct formation.

(g) H2NN + NH2 −→ Products.

Adduct formation followed by breaking an N–H bond is the only exothermicchannel. The transition state rate coefficient expression, in cm3mol−1s−1 units,is

H2NN+ NH2 −→ HNNNH2+ H k30g1 = 7.9× 106 T 1.9 exp(+670/T ) ,

significantly larger than the DHT rate coefficient expression

H2NN+NH2 −→ NH3+NNH k30g2 = 1.8× 106 T 1.94 exp(+580/T ) .

(h) H2NN + HO2 −→ Products.

Adduct formation followed by breaking an O–O bond is 192 kJ/mol exothermic.The TST result is, in cm3mol−1s−1 units,

H2NN+HO2 −→ NH2NO+ OH k30h1 = 6.6×105 T 1.94 exp(−3550/T ) ,

while the DHT method gives

H2NN+ HO2 −→ NNH+H2O2 k30h2 = 2.9× 104 T 2.69 exp(+805/T ) .

This channel is 88 kJ/mol more exothermic than formation of N2H3 + O2.

Page 121: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 245

2.7.6 Reactions of N2H3

Dissociation of N2H3 to N2H2 + H is endothermic by about 226 kJ/mol, signifi-cantly more than similar beta scission reactions, and so N2H3 is an unusually long-lived radical. We expect that its most important reactions are adduct formationsand disproportionations; abstractions from hydrocarbons by N2H3 should be slow,as the strength of the H–N bond formed is only 322 kJ/mol.

(a) N2H3 −→ N2H2 + H.

QRRK analysis suggests that this reaction is in the falloff regime for typicalcombustion conditions. For N2 bath gas and temperatures in the range 600 to2500 K the QRRK analysis gives in s−1 units

k31a = 2.3× 1043T−9.55 exp(−32400/T ) 0.1 atm

= 3.6× 1047T−10.38 exp(−34700/T) 1.0 atm

= 1.8× 1045T−9.39 exp(−35300/T ) 10 atm.

These expressions are shown in Fig. 60.

–4

–1

2

5

0.5 1.0 1.5 1000 K / T

log

( k

/ cm

3 m

ol–1

s–1

)

0.1 atm N2

1.0 atm N2

10 atm N2

N2H3 → N2H2 + H

FIGURE 2.60. QRRK results for dissociation of N2H3.

Page 122: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

246 Anthony M. Dean, Joseph W. Bozzelli

(b) N2H3 + H −→ Products.

The excited hydrazine formed by this association reaction can dissociate bythe pathways shown in Fig. 29. QRRK analysis indicates that the dominantchannel is NH2 + NH2, the reverse of reaction 17d. Formation of H2NN + H2 isdisfavored by its lower A-factor, which outweighs the somewhat lower barrierheight. The DHT estimate for its rate coefficient is

N2H3 +H −→ N2H2 +H2 k31b = 2.4× 108 T 1.5 cm3mol−1s−1.

(c) N2H3 + O −→ Products.

There are two exothermic channels

N2H3 +O −→ NH2 + HNO, 1r H = −154 kJ (31c1)

NH2NO+ H, 1r H = −152 kJ. (31c2)

suggesting rapid reaction. This pair of reactions is analogous to the even moreexothermic pair of O + C2H5 reactions, where the adduct can form either H +CH3CHO or CH3 + CH2O. We adopt Baulch et al.’s recommendation for thetotal rate coefficient of O + C2H5 and assume that the two channels have equalrates.

k31c1 = 3× 1013 cm3mol−1s−1 k31c2 = 3× 1013 cm3mol−1s−1.

For the hydrogen atom transfer reaction producing N2H2 + OH the DHT estimateis

k31c3 = 1.7× 108 T 1.50 exp(+325/T ) cm3mol−1s−1.

(d) N2H3 + OH −→ Products.

The DHT estimate for the abstraction reaction producing H2O + N2H2 is

k31d1 = 1.2× 106 T 2 exp(+600/T ) cm3mol−1s−1.

The N2H3OH adduct has an exothermic product channel forming H2NN +H2O. The barrier for this channel is low, because H2NN is a singlet biradical;it is therefore analogous to H2NN + H2 formation from NH2 + NH2 (Section6.13). QRRK analysis indicates that this pathway dominates and that the ratecoefficient is that for adduct formation

k31d2 = 3× 1013 cm3mol−1s−1.

(e) N2H3 + CH3 −→ Products.

For the hydrogen transfer reaction producing N2H2 + CH4 the DHT estimate is

k31e1 = 8.2× 105 T 1.87 exp(−915/T ) cm3mol−1s−1.

Page 123: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 247

The adduct has a low energy exit channel to form H2NN + CH4. Despite thelower A-factor of this pathway it dominates the H-atom transfer pathway, therate coefficient being approximately that of adduct formation

k31e2 = 3× 1013 cm3mol−1s−1.

(f) N2H3 + NH2 −→ Products.

The DHT estimate for the reaction to form N2H2 + NH3 is

k31 f 1 = 9.2× 105 T 1.94 exp(+580/T ) cm3mol−1s−1.

The energetics for recombination channels are similar to those of the CH3 +CH3 reactions. QRRK analysis shows some falloff at high temperatures dueto dissociation of the adduct back to reactants, but formation of H2NN + NH3dominates with a rate coefficient approximately that of adduct formation:

k31 f 2 = 3× 1013 cm3mol−1s−1.

(g) N2H3 + HO2 −→ Products.

Adduct formation followed by rupture of the O–O bond to yield NH2NHO +OH is 105 kJ/mol exothermic. The rate coefficient is expected to be that ofadduct formation

k31g1 = 3× 1013 cm3mol−1s−1.

The DHT estimate for the H-atom transfer reaction forming N2H2 + H2O2 is

k31g2 = 2.9× 104 T 2.69 exp(+805/T ) cm3mol−1s−1

and that for the one forming N2H4 + O2 is

k31g3 = 9.2× 105 T 1.94 exp(−1070/T ) cm3mol−1s−1.

2.7.7 Reactions of N2H4

Hydrazine can be formed in ammonia flames by NH2 + NH2 association. Itsdissociation to N2H3 + H can be neglected relative to dissociation to NH2 + NH2,the reverse of reaction 17a. Dissociation to form H2NN + H2 can be important atlower temperatures. The QRRK estimates for this reaction in N2 bath gas in thetemperature range from 600 to 2500 K are in s−1 units

k32 = 4.0× 1044T−9.85 exp(−35900/T ) 0.1 atm

k32 = 5.3× 1039T−8.35 exp(−34900/T ) 1.0 atm

k32 = 2.5× 1039T−8.19 exp(−35100/T ) 10 atm.

Page 124: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

248 Anthony M. Dean, Joseph W. Bozzelli

The N–H bond being relatively weak, about 322 kJ/mol, suggests that thehydrogen can be easily abstracted, but few data are available. Using the DHTmethod we estimated the following expressions for abstractions from hydrazine:(in cm3mol−1s−1 units)

N2H4+H −→ N2H3 +H2 k32a = 9.6× 108T 1.5 exp(−2435/T )

+O −→ N2H3 +OH k32b = 6.7× 108T 1.5 exp(−1435/T )

+OH −→ N2H3 +H2O k32c = 4.8× 106T 2.0 exp(+325/T )

+CH3 −→ N2H3 + CH4 k32d = 3.3× 106T 1.87 exp(−2680/T )

+NH2 −→ N2H3 +NH3 k32e = 3.7× 106T 1.94 exp(−820/T ) .

These estimates can be compared to the limited experimental data at low tem-peratures. The estimate for abstraction by hydrogen atoms gives k = 1.5× 109

cm3mol−1s−1 at 300 K. This compares to experimental values ranging from1 to 2 ×1011 cm3mol−1s−1 at 300 K (Mallard et al. 1993). For abstractionby OH at room temperature, the above estimate gives 1.3× 1012 cm3mol−1s−1,whereas measurements of the total rate report values ranging from 1.3 to 3.9×1013

cm3mol−1s−1 (Mallard et al. 1993). An experimental study of the abstractionby NH2 (as a secondary reaction in the H + N2H4 system) by Gehring etal. (1971) gave k = 3 × 1011 cm3mol−1s−1 at 300 K, compared to our estimate1.5× 1010 cm3mol−1s−1. For CH3 the DHT estimate is 2.9×108 cm3mol−1s−1

at 400 K compared to the reported value of 1.8×108 cm3mol−1s−1(Mallard etal. 1993). These comparisons are disappointing, especially with regard tothe abstraction by H-atoms. The DHT estimates should be better at highertemperatures where the activation energy influence is diminished. There is someevidence that radical reactions with hydrazine may not be direct ones. For example,Vaghijiani (1996) reported that the OH yield in the reaction of O-atoms withhydrazine accounts for only 15% of the total reaction at 298 K. If all of theobserved OH were due to the direct abstraction reaction it would nonetheless beabout 20 times the DHT estimate.

2.7.8 Reactions of NO

Many reactions of NO have been considered earlier. For completeness we addseveral more here.

(a) NO + C −→ Products.

There are two likely product channels:

NO+ C −→ CO+N 1r H = −444 kJ (33a1)

NO+ C −→ CN+O 1r H = −121 kJ. (33a2)

There have been two studies of these reactions behind shock waves, Deanet al. (1991) reporting a total rate coefficient of 4.8 × 1013 cm3mol−1s−1

Page 125: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 249

and Lindackers et al. (1990) reporting 2.0 × 1013 cm3mol−1s−1. Dorthe etal. (1991) measured 1.6 × 1013 cm3mol−1s−1 at room temperature. Deanet al. reported that CN + O contributes 40% of the total rate and CO + Nthe remaining 60%. We suggest use of a simple average of the overall ratemeasurements together with the reported branching ratio, that is,

k33a1 = 1.7× 1013 cm3mol−1s−1

k33a2 = 1.1× 1013 cm3mol−1s−1 .

(b) NO + OH + M −→ HONO + M.

This reaction can serve as a sink for OH in the pre-ignition chemistry ofinternal combustion engines with exhaust gas recirculation, which could leadto effectively higher octane numbers. It is discussed as a reverse reaction inSection 7.15.

(c) NO + HCCO −→ Products.

Modeling studies, e.g., by Lindstedt et al. (1995), identify this reaction as amajor route for NO destruction in hydrocarbon flames. Boullart et al. (1994)reported a study of the rate over the temperature range from 290 to 670 K leadingto the rate coefficient expression k33c = 6×1013 exp(−350/T ) cm3mol−1s−1.

The major channel at 700 K was found to be HCNO + CO, with HCN + CO2making up the balance. Assuming that the Boullart et al. branching ratio istemperature independent leads to the two expressions, in cm3mol−1s−1 units,

NO+HCCO −→ HCNO+ CO k33c1 = 4.6× 1013 exp(−350/T )

−→ HCN+ CO2 k33c2 = 1.4× 1013 exp(−350/T ) .

A theoretical study of the reaction was reported by Nguyen et al. (1994).

(d) NO + C2H −→ Products.

This reaction was proposed to influence NOx chemistry in flames by Peeterset al. (1996), who reported an overall rate coefficient expression for thetemperature range from 295 to 440 K

k33d = 6.0× 1013 exp(−287/T ) cm3mol−1s−1 ,

which is similar to the expression for the analogous reaction of NO with HCCO,33(c). No information is available about the product distribution. Substantialrearrangement of the initially formed adduct would be required to reach theexothermic product channels forming HCN and CO (−630 kJ/mol) or CN andHCO (−170 kJ/mol). For this reason we hesitate to recommend likely productsand rate coefficients. Additional experiments to measure branching ratios areneeded for this reaction as well as for the reactions of NO with C2H3 and C2H5.

Page 126: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

250 Anthony M. Dean, Joseph W. Bozzelli

2.7.9 Reactions of NO2

(a) NO2 + H −→ NO + OH.

As shown in Fig. 20, H + NO2 can form HONO and HNO2 adducts, withthe likely final products from both being OH + NO. With this exothermic exitchannel one expects fast reaction and little effect of pressure. The reaction wasstudied by Ko et al. (1991) over the temperature range from 296 to 760 K. Wesuggest use of their rate coefficient expression

k34a = 1.3× 1014 exp(−180/T ) cm3mol−1s−1.

An interesting feature of this reaction is that its reverse would be expected toproduce primarily HONO, with some H + NO2 at high temperatures, as thechannel to H + NO2 is 121 kJ/mol endothermic. Consequently, there couldbe high pressure, lower-temperature combustion conditions where one needs toaccount for reactions of HONO, as discussed in Section 7.15.

(b) NO2 + O −→ NO + O2.

This reaction is similar to reaction 34a in that two different adducts can beformed without barriers and there is a low energy exit channel, here to NO + O2,so that one could form either ONOO or the symmetrical structure NO3. ONOOshould rapidly dissociate to form NO + O2, while there could be an appreciablebarrier for isomerization-dissociation or concerted elimination from NO3. Thelow temperature data show a small decrease in rate coefficient with increasingtemperature, suggesting adduct formation. This reaction was reviewed byBaulch et al. (1992), who recommend

k34b = 3.9× 1012 exp(+120/T ) cm3mol−1s−1

from 230 to 350 K. This expression gives k = 4.2 × 1012 cm3mol−1s−1 at2000 K, which can be compared to the recent report by Zuev et al. (1991c) thatk = 6.0× 1012 cm3mol−1s−1 over the range 1780 to 2300 K.

(c) NO2 + M −→ NO + O + M.

One expects thermal dissociation of this small molecule to be in the low pressurelimit over the pressure range of interest in combustion. R�ohrig et al. (1997)studied the reaction in shock waves from 1350 to 2100 K and derived the lowpressure limit expression

k34c = 4.0× 1015 exp(−30200/T ) cm3mol−1s−1

for M = Ar. This expression is in reasonable agreement with the one derivedby Tsang and Herron (1991) from the reverse reaction rate. Tsang and Herronpresent falloff parameters and third-body efficiencies as well.

Page 127: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 251

(d) NO2 + NH2 −→ Products.

Glarborg et al. (1995) reported as results of a modeling study of the NH3–NO2system from 850 to 1350 K the expressions in cm3mol−1s−1 units

NO2 +NH2 −→ N2O+H2O k = 3.2× 1018T−2.2

−→ NH2O+NO k = 3.5× 1012 .

Park and Lin (1997) studied the reaction over the temperature range from 300to 900 K and reported

ktotal = 8.1× 1016 T−1.44 exp(−135/T ) cm3mol−1s−1 ,

with the N2O channel being 19% of the total, independent of temperature.This branching ratio is consistent with the results of Lindholm and Hershberger(1997), who reported a branching ratio of 24% to N2O and H2O, 76% to NH2Oand NO. In contrast, extrapolation of the Glarborg et al. result to 300 K predicts77% for the N2O channel. On the other hand, the branching ratio reported byMeunier et al. (1996), 59% to N2O and H2O, 40% to NH2O and NO, are closerto the extrapolated Glarborg et al. expression. There remains ambiguity in thebranching ratio of this reaction. For combustion modeling we suggest usingthe expression of Park and Lin together with their temperature-independentbranching ratio, in cm3mol−1s−1 units

NO2 +NH2 −→ N2O+H2O k34d1 = 1.5× 1016 T−1.44 exp(−135/T )

−→ NH2O+NO k34d2 = 6.6× 1016 T−1.44 exp(−135/T ) .

(e) NO2 + CH3 −→ CH3O + NO.

Biggs et al. (1993) report for this reaction

k34e = 1.4× 1013 cm3mol−1s−1

at room temperature. The reaction is 70 kJ exothermic and little temperaturedependence is to be expected. At high pressure and low temperature, formationof the stabilized adducts CH3NO2 (nitromethane) and CH3ONO is possible.

2.7.10 Reactions of N2O

(a) N2O + OH −→ N2 + HO2.

Little information is available on the rate of this potentially important reaction.It was reviewed by Tsang and Herron (1991). The available data are allfrom experiments at low temperatures, and most report only upper-limit ratecoefficients. Tsang and Herron assign an activation energy of 42 kJ/mol to beconsistent with the upper-limit measurements. Miller and Bowman (1989) use

Page 128: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

252 Anthony M. Dean, Joseph W. Bozzelli

this activation energy together with an A-factor of 2×1012 cm3mol−1s−1, fourtimes lower than the Tsang-Herron upper-limit recommendation. The likelyproducts, particularly at lower temperatures, are HO2 and N2. Use of eitherof these expressions predicts formation of N2O at intermediate combustiontemperatures, where HO2 concentrations are large, and the reaction goes inreverse.Sausa et al. (1993) reported that their flame data are best interpreted with anexpression for this rate coefficient no larger than the one suggested by Millerand Bowman and that an even lower expression would also fit their data. Astudy of N2O decomposition by Allen et al. (1994) suggests that use of theMiller-Bowman expression predicts too much NO2 and too little NO. Theiranalysis showed that a better fit to the data could be obtained by lowering k35by a factor of 160; this reduced the production of HO2, which is responsible forconverting NO to NO2 under their conditions.Mebel et al. (1996) studied this reaction with molecular electronic structuremethods. An abstraction pathway to form N2 + HO2 as well as an additionpathway to form either a stabilized adduct HONNO or the dissociation productsHNO + NO were included. Their transition state theory abstraction ratecoefficient expression for the temperature range from 1000 to 5000 K,

N2O+OH→ N2+HO2 k35 = 1.3× 10−2 T 4.72 exp(−18400/T ) ,

in cm3mol−1s−1 units, is consistent with the upper limit of Allen et al. (1994,1995).Mebel et al. conclude that above 1400 K and 1 atm pressure the HONNOadduct rapidly dissociates to HNO + NO, while stabilization dominates below1000 K. The dissociation channel forming HNO + NO is accounted for as itsreverse reaction in the following section.The abstraction channel appears to dominate under combustion conditions.Even at 2000 K, however, k35 is only about 5× 109 cm3mol−1s−1, so it is notlikely that this reaction will be important under any combustion conditions.

2.7.11 Reactions of HNO

The H–NO bond strength in HNO is only 200 kJ/mol, intermediate between the lowbond dissociation energies associated with unstable radicals and normal covalentbond strengths. Because it has only 3 atoms, its dissociation rate is close to thelow pressure limit throughout the combustion temperature range. Unimoleculardissociation of HNO is sufficiently slow that its bimolecular reactions can beimportant, particularly at temperatures below 1500 K. Hydrogen abstractionsshould be especially important, as they will generally be exothermic. There isalso the possibility of radical addition across the double bond.

As a relatively stable molecule with an easily abstractable hydrogen atom,HNO offers the following catalytic chain termination sequence, where X and X′

Page 129: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 253

are atoms or radicals:

H+NO −→ HNO

HNO+X −→ HX+NO

(overall) H+X −→ HX .

Other species, such as HONO or HNO2, can function similarly. If we designatethese species as XR, then the catalytic mechanism is:

X+ R −→ XR

XR+X′ −→ XX′ + R

(overall) X′ +X −→ X′X .

(a) HNO + M −→ H + NO + M.

Hanson and Salimian (1984) suggest for M = Ar

k36a = 1.8× 1016 exp(−24500/T ) cm3mol−1s−1.

At 1500 K and 1 atm the effective first order rate coefficient is 1.2× 104 s−1,over four orders of magnitude slower than NNH dissociation. A radical at aconcentration of 1 nmol/cm3 (about 100 ppm at 1 atm) reacting with HNOwith a rate coefficient near 1× 1013 cm3mol−1s−1 would compete with HNOthermal dissociation.

(b) HNO + HNO −→ N2O + H2O.

This reaction was studied by He and Lin (1992), and a potential surface wasreported by Lin et al. (1992a). The initially formed dimer undergoes severalisomerizations prior to forming N2O + H2O. The barriers to isomerization arehigher than that of the reverse reaction, and thus one sees rate falloff at highertemperatures. The observed loss of HNO is best explained by adduct formationrather than as formation of final products. He and Lin report

k36b = 8.5× 108 exp(−1550/T ) cm3mol−1s−1

for formation of N2O + H2O over the temperature range 450 to 520 K. Therate coefficient is pressure dependent and was reported only at 1 atm. The Heand Lin expression is orders of magnitude smaller than that used by Miller andBowman.

(c) HNO + OH −→ H2O + NO.

The abstraction channel is 293 kJ/mol exothermic and expected to be fast.Transition state calculations by Soto et al. (1991) led to

k36c = 1.3× 107T 1.88 exp(+480/T ) cm3mol−1s−1,

Page 130: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

254 Anthony M. Dean, Joseph W. Bozzelli

in reasonable agreement with the limited data available. Soto et al. concludethat the abstraction channel is faster than the addition channel.

(d) HNO + H −→ Products.

This reaction was analyzed by Soto and Page (1993), who used variationaltransition state theory to calculate a rate coefficient for the abstraction channelgiven in cm3mol−1s−1 units by

H+HNO −→ H2 + NO, k36d1 = 4.5× 1011T 0.72 exp(−330/T )

over the range 200 to 3000 K. Their expression, about an order of magnitudehigher than earlier estimates, is supported by experiments on the reversereaction by Natarajan et al. (1994), who monitored H-atom profiles in H2–NO mixtures between 1760 and 2160 K. Their measurements imply k36d1 ≈5 × 1013 cm3mol−1s−1 at 1900 K, compared to the Soto and Page value of8.7× 1013 cm3mol−1s−1.In other work Page and Soto (1993) describe calculations which indicatethat abstraction is faster than addition at all temperatures. For the addition-elimination channel Cohen and Westberg (1991) report in cm3mol−1s−1 units

H+ HNO −→ O+NH2, k36d2 = 3.5× 1015T−0.3 exp(−14730/T ).

This expression gives rate coefficient values much lower than those of theabstraction channel, consistent with the Page-Soto analysis. As can be seenin Fig. 26, the lower expression for the addition channel is primarily due toenergetics. Not only is there a barrier in the entrance channel, but both O +NH2 and NH + OH are endothermic exit channels. We cautiously recommendthe rate coefficient calculated by Soto and Page for the abstraction channel. Theabstraction is direct and so no pressure dependence is anticipated. The Cohen-Westberg expression for the addition-elimination route is consistent with anexperimental measurement of the reverse reaction rate (cf. Section 6.11). Wesuggest adoption of the rate coefficient for the reverse reaction.

(e) HNO + O −→ OH + NO.

Campbell and Handy (1975) reported a lower limit of 1.1×1013 cm3mol−1s−1

for the rate coefficient of this reaction at 425 K, consistent with other exothermicradical + HNO abstractions; e.g., the Soto-Page expression for H + HNO gives1.6 × 1013 cm3mol−1s−1 at 425 K. We suggest use of the H + HNO ratecoefficient expression for this reaction also.

(f) HNO + NH2 −→ NO + NH3.

We are not aware of any measurements or calculations for this reaction. Hansonand Salimian estimate 5× 1011T 0.5 exp(−500/T ) cm3mol−1s−1, while Millerand Bowman use 2×1013 exp(−500/T ) cm3mol−1s−1. Again, from the nature

Page 131: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 255

of the species involved one expects fast exothermic abstraction. The DHT ratecoefficient expression is

k36f = 9.2× 105T 1.94 exp(+580/T ) cm3mol−1s−1.

This expression is 5 to 10 times less than the Hanson and Salimian and Millerand Bowman expressions. It is in much better agreement with the transitionstate calculations of Mebel et al. (1996). As with the other HNO reactions, oneexpects abstraction to dominate over addition and no pressure dependence ofthe rate coefficient.

(g) HNO + NO −→ N2O + OH.

Miller and Bowman use k = 2×1012 exp(−13100/T) cm3mol−1s−1 ,which isconsistent with the reaction proceeding by addition followed by OH eliminationto form the final products N2O + OH. The A-factor is consistent with an additionreaction, and one expects a high barrier, caused by ring strain, for hydrogentransfer from the nitrogen to the oxygen on the same nitrogen, followed byrapid scission to give the final products. This rate coefficient is much smallerthan that for H abstraction from HNO.

In a study of the NO–H2 system Diau et al. (1995) considered this reaction tobe rate limiting. The rates of NO decay and CO2 production from added COwere modeled to yield

k36g = 8.5× 1012 exp(−14890/T ) cm3mol−1s−1.

This expression was shown to be consistent with an addition–isomerization–decomposition pathway. An RRKM analysis of the unimolecular isomerizationstep, which is rate-limiting, with a master equation treatment of the collisionalstabilization process succeeded in matching the experimentally derived ratecoefficient. Their expression is also reasonably consistent with the one used byMiller and Bowman (1989) and we recommend it for combustion modeling.

(h) HNO + O2 −→ NO + HO2.

Because of the weak H–N bond in HNO this reaction is nearly thermoneutral.By analogy to other RH + O2 reactions we estimate

k36h = 2× 1013 exp(−8000/T ) cm3mol−1s−1.

This expression gives rate coefficient values greater than the one proposed byMiller et al. (1991).

(i) HNO + CH3 −→ NO + CH4.

To provide an interaction with a hydrocarbon radical we estimated an expressionfor abstraction by CH3, generally the most abundant radical in hydrocarbonflames, by the DHT method:

k36i = 8.2× 105 T 1.87 exp(−480/T ) cm3mol−1s−1.

Page 132: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

256 Anthony M. Dean, Joseph W. Bozzelli

2.7.12 Reactions of NH2O

In Sections 6.9 and 6.10 we indicated that NH2O can be an important intermediatein reactions of NH2 with O2 or HO2. For a radical, NH2O is unusually stable tothermal dissociation of the H atom, having an N–H bond strength of about 255kJ/mol, and so isomerization to the more stable isomer HNOH is more likely thandissociation. As a consequence of the relatively strong N–H bond, one expectsthat the rates of disproportionation reactions, such as NH2O + CH3 −→ HNO +CH4, would be somewhat smaller than usual since the exothermicity is less thanin the typical case. Radical recombination reactions should be important; two ofthem, NH2O + OH and NH2O + O, were introduced earlier as the reverses of NH2+ HO2 and NH2 + O2, respectively. (Cf. Sections 6.9 and 6.10.) One does notexpect hydrogen transfer reactions of the form NH2O + RH −→ NH2OH + R,since the strength of the O–H bond formed being only about 322 kJ/mol meansthat such reactions would usually be very endothermic.

(a) NH2O + M −→ HNO + H + M.

QRRK analysis predicts that the dissociation rate coefficient is close to its low-pressure limit for all conditions of interest in combustion, the rate coefficientexpression for M = N2 being given by

k37a = 2.8× 1024T−2.83 exp(−32700/T) cm3mol−1s−1.

At 2000 K and 1 atm pressure the effective first-order rate coefficient is about600 s−1, indicating that under combustion conditions dissociation will beunimportant compared to reactions with radicals.

(b) NH2O + M−→ HNOH + M.

QRRK calculations predict that the rate coefficient for isomerization is alsonear its low pressure limit under combustion conditions. Near 1 atm pressureof N2 the QRRK expression is

k37b = 1.1× 1029T−3.99 exp(−22100/T ) cm3mol−1s−1.

Isomerization is much faster than dissociation; at 2000 K and 1 atm N2,its effective first-order rate coefficient is about 7 × 105s−1, three orders ofmagnitude greater than the rate coefficient for dissociation.

(c) NH2O + H −→ H2NOH −→ NH2 + OH−→ HNO + H2.

The adduct formation pathway, exothermic by about 60 kJ/mol, is analogous toH + C2H5−→ 2 CH3. We estimate, based on the expression of Baulch et al.(1992) for the C2H5 analog,

k37c1 = 4.0× 1013 cm3mol−1s−1.

Page 133: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 257

The DHT estimate for the direct reaction producing HNO + H2 is

k37c2 = 4.8× 108 T 1.5 exp(−785/T ) cm3mol−1s−1.

The rate of the direct reaction approaches that of the adduct formation pathwayat high temperatures.

(d) NH2O + O −→ Products.

The adduct dissociates quickly to NH2 and O2, which has been considered inSection 6.9 as the reverse reaction. The DHT estimate for the channel producingHNO and OH is, in cm3mol−1s−1 units,

NH2O+ O −→ HNO+OH k37d = 3.3× 108 T 1.5 exp(−245/T ) .

(e) NH2O + OH −→ Products.

The channel forming HNO + H2O, exothermic by 250 kJ/mol, is blocked fromthe adduct, the 115 kJ/mol adduct well depth being overwhelmed by the 250kJ/mol exit barrier. (Cf. Fig. 25) The DHT estimate for the abstraction channelis, in cm3mol−1s−1 units,

NH2O+OH −→ HNO+H2O k37e = 2.4× 106 T 2 exp(+600/T ) .

(f) NH2O + CH3 −→ Products.

The channel forming NH2 and CH3O is exothermic by about 8 kJ/mol. Conse-quently, the NH2OCH3 adduct is expected to dissociate both to reactants and toproducts. We estimate the rate coefficient to be half that of the high-pressurelimit of the CH3 + CH3 reaction (Baulch et al. 1992)

NH2O+ CH3 −→ CH3O+NH2 k37f1 = 2× 1013 cm3mol−1s−1.

The DHT method gives in the same units

NH2O+ CH3 −→ HNO+ CH4 k37f2 = 1.6× 106 T 1.87 exp(−1490) .

The hydrogen transfer pathway to form NH2OH and 3CH2 is 130 kJ/molendothermic and can be neglected. The adduct pathway is substantially fasterthan the direct one for all combustion conditions. The reaction is thereforechain propagating rather than chain terminating.

(g) NH2O + NH2 −→ Products.

There are no accessible exothermic channels from the adduct, as the barrier toH2O formation is high. The DHT estimate for the HNO + NH3 channel is

k37g = 1.8× 106 T 1.94 exp(+580/T ) cm3mol−1s−1.

Page 134: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

258 Anthony M. Dean, Joseph W. Bozzelli

(h) NH2O + HO2 −→ Products.

Transfer of an H-atom can proceed in two ways

NH2O+ HO2 −→ HNO+H2O2 (37h1)

−→ NH2OH+O2 (37h2)

with comparable exothermicity. The DHT estimate for abstraction by HO2gives

k37h1 = 2.9× 104T 2.69 exp(+805/T ) cm3mol−1s−1 .

Since there is insufficient data to estimate the rate coefficient for abstraction byNH2O, we estimate k37h2 = k37h1 based on the comparable exothermicities andexpected similarity in A-factors.

2.7.13 Reactions of HNOH

Much of the discussion of NH2O reactions in the previous section also applies toNHOH, as its relatively high O–H bond dissociation energy, about 230 kJ/mol,has similar implications. Dissociation will be relatively slow under combustionconditions, and so reactions with radicals should be considered. As for NH2O,hydrogen abstraction reactions from stable molecules would be expected to beunimportant for HNOH, as the strength of the N–H bond formed would only beabout 350 kJ/mol.

(a) HNOH + M −→ HNO + H + M.

QRRK analysis shows that the rate coefficient of this reaction is near its lowpressure limit for typical combustion conditions, with

k38a = 2.0× 1024T−2.84 exp(−29700/T ) cm3mol−1s−1

for M = N2. While faster than NH2O dissociation, reaction 37a, it is still notfast enough to be important.

(b) HNOH + H −→ H2NOH −→ NH2 + OH−→ HNO + H2.

The adduct formation pathway is exothermic by about 80 kJ/mol and has nosignificant barrier to dissociation to products. We estimate

k38b = 4× 1013 cm3mol−1s−1.

from the Baulch et al. (1992) value for the rate coefficient of H + C2H5 −→ CH3+ CH3. The DHT estimate for the abstraction pathway is

k38b2 = 4.8× 108 T 1.5 exp(−190/T ) cm3mol−1s−1.

Page 135: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 259

(c) HNOH + O −→ HN(O)OH −→ HNO + OH−→ HNO + OH.

Adduct formation followed by exothermic (1r H = −200 kJ/mol) dissociationis expected to be fast. We suggest

k38c1 = 7× 1013 cm3mol−1s−1,

which is the Baulch et al. (1992) recommendation for the rate coefficient of theO + C2H5 reaction.For the direct hydrogen abstraction path giving the same products the DHTestimate is

k38c2 = 3.3× 108 T 1.5 exp(+180/T ) cm3mol−1s−1.

The two pathways have comparable rates at high temperatures.

(d) HNOH + OH −→ HNO + H2O.

We expect the adduct formation pathway to be slow in this case because of thehigh barrier and a low A-factor to the product channel. The DHT estimate fordirect reaction is

k38d = 2.4× 106 T 2 exp(+600/T ) cm3mol−1s−1.

(e) HNOH + CH3 −→ CH3NHOH −→ CH3NH + OH−→ HNO + CH4.

We estimatek38e1 = 2× 1013 cm3mol−1s−1

for the adduct formation pathway on the basis that its 17 kJ/mol exothermicitymakes it comparable to reaction 37f1. The DHT estimate for the abstractionpathway is

k38e2 = 1.6× 106 T 1.87 exp(−1055/T ) cm3mol−1s−1.

The adduct formation pathway is faster.

(f) HNOH + NH2 −→ HN(OH)NH2 −→ N2H3 + OH−→ H2NN + H2O

−→ HNO + NH3.

The adduct formation pathway forming water, analogous to H2NN + H2 for-mation in the NH2 + NH2 reaction, has a low barrier, H2NN being a singletbiradical. A barrier of 85 kJ/mol was assumed for its reverse reaction. QRRKanalysis gave in cm3mol−1s−1 units

HNOH+NH2 −→ N2H3 +OH, k38f1 = 6.7× 106 T 1.82 exp(−360/T )

−→ H2NN+H2O, k38f2 = 4.6× 1019 T−1.94 exp(−970/T ) .

Page 136: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

260 Anthony M. Dean, Joseph W. Bozzelli

The DHT estimate for the abstraction pathway is

k38f3 = 1.8× 106 T 1.94 exp(+580/T ) cm3mol−1s−1.

(g) HNOH + HO2 −→ HN(OH)OOH−→ HONHO + OH−→ HNO + H2O2

−→ NH2OH + O2.

The adduct formation pathway is exothermic by 70 kJ/mol and has no significantbarrier. We estimate

k38g1 = 4× 1013 cm3mol−1s−1.

The DHT estimate for the channel to HNO and H2O2 is

k38g2 = 2.9× 104 T 2.69 exp(+805/T ) cm3mol−1s−1.

In view of its comparable exothermicity and expected similar A-factor, weassume that the channel to NH2OH + O2 has the same rate.

k38g3 = 2.9× 104 T 2.69 exp(+805/T ) cm3mol−1s−1.

2.7.14 Reactions of 1HNOO

Formed in the reaction between 3NH and O2, this adduct can dissociate quiterapidly, (Figure 20) and reactions with other radicals can be neglected in com-bustion modeling. Reaction with O2 is also expected to be too slow to besignificant in combustion. The major reaction channels are thus dissociations.QRRK calculations using the parameters listed in Table 2.7 indicate that thedominant dissociation channel produces OH and NO, followed by dissociation tothe reactants 3NH and O2. Dissociation to H and NO2 is computed to be over 100times slower than dissociation back to OH and NO. For combustion conditionsthis reaction is at the low pressure limit. The QRRK estimate of its rate coefficientin N2 is in cm3mol−1s−1 units

1HNOO+M −→ OH+NO+M k39 = 1.5× 1036 T−6.18 exp(−15670/T) .

Page 137: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 261

2.7.15 Reactions of HONO

In Section 7.8 we discussed the formation of HONO from OH and NO. Becauseof the relative stability of HONO its reactions with radical species have to beconsidered. The H–O bond strength is only 330 kJ/mol, so H-atom abstractionfrom HONO would be generally quite exothermic. Abstraction rate coefficientsestimated by the DHT method are included in Table 19.

HONO thermal dissociation to form OH and NO was analyzed with the QRRKmethod. The dissociation rate coefficient for combustion conditions was foundto be near its low-pressure limit even up to pressures characteristic of internalcombustion engines and turbines. For N2 bath gas we estimate

k40a = 2.0× 1031T−4.56 exp(−25800/T ) cm3mol−1s−1.

Dissociation was found to dominate over isomerization to HNO2, as expected fromits much larger A-factor, i.e., much looser transition state. Rate measurements onthe reverse reaction OH + NO −→ HONO by Zabarnick (1993) indicate that thehigh pressure limit rate coefficient for this reaction is above 8×1012 cm3mol−1s−1,and the study by Forster et al. (1995) at pressures up to 150 bar gave k∞ =2×1013 cm3mol−1s−1 at 298 K. These values are consistent with the assumptionsmade for QRRK calculations. (Cf. Table 7).

Hsu et al. (1997) reported a molecular electronic structure study of the H +HONO reactions. Their calculations suggest that the adduct formation pathwaysare more important than direct abstraction, especially at lower temperatures. Theirvalue for the abstraction channel, in cm3mol−1s−1 units, is

HONO+H→ H2+NO2 , k40b1 = 2.0× 108 T 1.55 exp(−3330/T ) ,

which is lower than the DHT estimate at 300 K but within a factor of 2 of it at1000 and 2000 K. Their results for the channels resulting from adduct formationare, in cm3mol−1s−1 units

HONO+H −→ H2O+NO k40b2 = 8.1× 106 T 1.89 exp(−1935/T )

−→ OH+HNO k40b3 = 5.6× 1010 T 0.86 exp(−2500/T ) .

2.7.16 Reactions of HNO2

Figure 20 illustrates some routes for the production of HNO2. QRRK analysissuggests that the main unimolecular process is isomerization to HONO, whichis in the falloff pressure regime for combustion conditions. For N2 bath gas theQRRK result is in s−1 units

k41 = 7.1× 1027T−5.40 exp(−26400/T ) 0.1 atm

= 1.3× 1029T−5.47 exp(−26600/T ) 1.0 atm

= 2.0× 1030T−5.50 exp(−27000/T ) 10 atm.

Page 138: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

262 Anthony M. Dean, Joseph W. Bozzelli

Because of their similar thermochemistry one expects abstractions from HNO2to be comparable to those from HONO. Rate coefficient expressions derived bythe DHT method are listed in Table 19. A cause for concern is that the DHTexpression for OH + HNO2−→ H2O + NO2 gives a room temperature valueappreciably lower than the one reported by Atkinson et al. (1989b). Althoughthe room temperature values estimated by the DHT method for abstractions byH and O-atoms are larger by factors of 20 and 30 than those estimated by Tsangand Herron (1991), the differences decrease to about a factor of 3 at temperaturesabove 1000 K.

2.7.17 Reactions of HCN

Because the C–H bond in HCN is very strong, about 520 kJ/mol, thermaldissociation is unimportant in combustion. The reactions which are most likelyto be important are hydrogen abstraction by OH, addition to the triple bond, andisomerization to HNC

HCN −→ HNC 1r H = 54 kJ/mol. (42)

Isomerization was reported by Lin et al. (1992b) to be an important reaction ofHCN at high temperatures. The isomerization barrier of about 200 kJ/mol permitsisomerization to be a facile unimolecular reaction pathway for the relatively stableHCN molecule. Lin et al. computed the rate coefficient expression at infinitepressure to be

k∞42 = 3.5× 1013 exp(−23, 750/T ) s−1.

One expects the actual isomerization rate coefficient of such a small molecule tobe far from its high-pressure limit under combustion conditions. QRRK analysisshowed that the reaction should usually be in the low pressure limit at temperaturesabove 1000 K, the rate coefficient for N2 being given by the expression

k42 = 1.6× 1026T−3.23 exp(−24950/T ) cm3mol−1s−1.

At 300 K the reaction is in the falloff regime at pressures near 10 atm. Ratecoefficient expressions at 0.1, 1 and 10 atm are included in Table 19.

Wooldridge et al. (1995) reported that their experiments on OH + HCN kineticswere not affected by this isomerization. Our QRRK-derived falloff expression,however, suggests that the isomerization reaction proceeds on the same time scaleas their experiments, with half-times ranging from 10 to 100 µs.

(a) HCN + OH −→ CN + H2O.

Baulch et al. (1992) recommend an expression similar to ones found for H-atomabstraction from hydrocarbons, k = 9.0×1012 exp(−5400/T ) cm3mol−1s−1,

significantly greater than the measurements of Wooldridge et al. (1995). (Figure61) We recommend their expression

k42a = 3.9× 106 T 1.83 exp(−5180/T ) cm3mol−1s−1 .

Page 139: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 263

10

11

12

0.5 0.8 1.11000 K / T

log

( k

/ cm

3 mol

–1 s

–1)

Wooldridge et al. (1995) totalMiller-Melius+abstractionQRRK+abstractionWooldridge et al. (1995) CN+H2OBaulch et al. (1992) CN+H2O

HCN + OH → Products

FIGURE 2.61. Rate coefficients for OH + HCN −→ Products. The results for theabstraction channel forming CN + H2O reported by Wooldridge et al. (1995) are comparedto the recommendation of Baulch et al. (1992). The total rate coefficient of Wooldridgeet al. is also shown. The abstraction channel is dominant. The total is also comparedto the sum of the rate coefficient measured for the abstraction channel and calculations ofthe total addition rate coefficient for all 3 addition channels. The comparisons indicatethat both the Miller-Melius and the QRRK calculations give the correct magnitude for theaddition channel rate coefficients, but that the temperature dependence of the Miller-Meliusexpression is more consistent with the data.

(b) HCN + OH −→ Products.

A potential energy diagram for the addition step is shown in Fig. 51, where

Page 140: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

264 Anthony M. Dean, Joseph W. Bozzelli

it appears as one of the product channels for the CH2 + NO reaction. Theoriginally-formed NCHOH adduct (the higher energy HCNOH can be safelyignored) has several possible dissociation paths, for which Miller and Melius(1986) have reported rate coefficients. Fitting their results over the temperaturerange from 1000 to 2500 K leads to the expressions in cm3mol−1s−1 units

HCN + OH −→ HNCO + H k42b1 = 4.4× 103T 2.26 exp(−3220/T )−→ HOCN + H k42b2 = 1.1× 106T 2.03 exp(−6730/T )−→ CO + NH2 k42b3 = 1.6× 102T 2.56 exp(−4530/T ).

QRRK analysis shows that at 1 atm pressure and temperatures less than 1000 Kformation of the stabilized NCHOH adduct dominates, while at higher temper-atures the major channels yield HNCO + H and HOCN + H. (As discussed inSection 6.19, the lowest energy NH2 + CO channel is disfavored because of theisomerization step required to access it.) QRRK analysis for N2 bath gas led tothe rate coefficient expressions in cm3mol−1s−1 units

HCN + OH −→ HNCO + H k = 3.5× 1011T 0.02 exp(−3560/T )−→ HOCN + H k = 9.9× 1010T 0.31 exp(−5350/T )−→ CO + NH2 k = 1.1× 1011T−0.18 exp(−3590/T )

−→ NCHOH k42b4 = 1.7× 1029T−6.31 exp(−2580/T ) 0.1 atm−→ NCHOH k42b4 = 2.8× 1030T−6.37 exp(−2690/T ) 1.0 atm−→ NCHOH k42b4 = 1.1× 1032T−6.53 exp(−3140/T ) 10 atm.

QRRK analysis generally indicates faster rates near 1000 K and a more gradualtemperature dependence than calculated by Miller and Melius. The differencescould be due to small variations in the barrier heights. The predictions arecompared to the measurements of Wooldridge et al. (1995) in Fig. 61, whereit can be seen that the abstraction channel (42a) dominates and that the sum ofthe calculated rate coefficients for the three adduct formation pathways and themeasured rate coefficient for the abstraction channel agrees with the measuredtotal rate coefficient. We suggest use of the Miller-Melius expressions for thenon-adduct channels and the QRRK result for adduct formation.

(c) HCN + O −→ Products.

A potential energy diagram for this addition reaction is shown in Fig. 58, whereit appears as one of the product channels for the CH + NO reaction. The NCHOadduct can react in a variety of ways. Perry and Melius (1984) analyzed thisreaction and reported the rate coefficient expressions in cm3mol−1s−1 units

HCN + O −→ NH + CO k42c1 = 5.4× 108T 1.21 exp(−3770/T )−→ NCO + H k42c2 = 2.0× 108T 1.47 exp(−3820/T )−→ CN + OH k = 2.7× 109T 1.58 exp(−13400/T ).

Formation of stabilized adducts is relatively unimportant. The significantchannels are found by QRRK analysis to have the rate coefficient expressionsin cm3mol−1s−1 units

Page 141: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 265

HCN + O −→ NH + CO k = 4.0× 1014T−0.65 exp(−4285/T )−→ NCO + H k = 7.6× 1010T 0.48 exp(−3930/T )−→ CN + OH k42c3 = 4.2× 1010T 0.40 exp(−10400/T).

In general, these QRRK expressions give somewhat smaller rate coefficients,essentially because Perry and Melius decreased all barriers by 11 kJ/mol inorder to improve the fit to the high temperature data. A major difference isthat the QRRK results suggest a much lower rate for production of CN + OH.Since the Perry-Melius expression predicts rate coefficients for this channelabout a factor of 3 higher than the measured upper limit at 2000 K of Szekelyet al. (1984), we suggest use of the QRRK estimate for that channel.

2.7.18 Reactions of HNC

Although the strong H–N bond should make abstractions from HNC unimportant,radicals such as O and OH should react with it by other pathways. Lin et al. (1992b)obtained the rate coefficient expressions in cm3mol−1s−1 units

HNC+O −→ HNCO −→ NH+ CO k43a = 4.6× 1012 exp(−1100/T )

HNC+OH −→ HNCOH −→ HNCO+ H k43b = 2.8× 1013 exp(−1860/T )

by modeling CO formation during HCN oxidation by NO2 over the 623–773 Ktemperature range.

HNC might also be expected to add to O2. A QRRK analysis was doneconsidering two product channels. The first was bond scission to form HNCO +O. Although this channel is 38 kJ/mol lower in energy than the reactants, the barrierbetween it and the adduct is expected to be 10 to 15 kJ/mol above the entrancechannel (Melius 1993). The second pathway is formation of a 3-membered ringand subsequent formation of NH + CO2. This barrier is expected to be lowerthan the entrance channel (Carpenter 1993), but a lower A-factor is to be expectedbecause of the ring formation. The QRRK calculations led to the rate coefficientexpressions in cm3mol−1s−1 units

HNC+O2 −→ HNCO+O k43c1 = 1.5× 1012T 0.01 exp(−2070/T )

−→ NH+ CO2 k43c2 = 1.6× 1019T−2.25 exp(−890/T ).

The higher A-factor for dissociation of the adduct to HNCO + O causes this channelto be more important at combustion temperatures, while NH + CO2 dominates atlower temperatures.

2.7.19 Reactions of CN

The H–CN bond of hydrogen cyanide is so strong, about 520 kJ/mol, that oneexpects hydrogen transfer reactions to CN to be very important. Since CN itself

Page 142: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

266 Anthony M. Dean, Joseph W. Bozzelli

is extremely stable, about 750 kJ/mol more stable than C + N, it can also react byradical addition and recombination.

(a) CN + H2 −→ HCN + H.

The abstraction reaction from H2 was studied in shock tubes by Natarajanet al. (1988) and Wooldridge et al. (1996) and at lower temperatures byAtakan et al. (1989). Tsang (1992) has reviewed this reaction. We adopt hisrate coefficient recommendation, which is consistent with these experimentalresults:

k44a = 3.6× 108T 1.55 exp(−1510/T ) cm3mol−1s−1.

(b) CN + H2O −→ HCN + OH.

This is one of the few exothermic abstractions from water. Because of thelarge concentrations of water found in post-flame gases, it can be important ininterconverting CN and HCN. Its rate coefficient can be inferred from the ratecoefficient expression for the reverse reaction 42a.

(c) CN + O −→ OCN −→ CO + N.

This reaction is 322 kJ/mol exothermic overall. A likely mechanism is adductformation by recombination followed by dissociation to low energy products.It should be very fast, analogous to reaction (38b). This is consistent with thehigh temperature measurements of Davidson et al. (1991b), who report

k44c = 7.7× 1013 cm3mol−1s−1.

(d) CN + O2 −→ NCOO −→ NCO + O.

The overall reaction is exothermic by about 55 kJ/mol, suggesting that oneshould expect rate falloff at high temperatures as dissociation of the adductback to the entrance channel becomes significant; a higher A-factor for entrancechannel dissociation is expected because of the higher entropy for two diatomicproducts. Davidson et al. (1991b) reported a high temperature measurement of

k44d = 1.0× 1013 cm3mol−1s−1,

consistent with the Baulch et al. (1992) recommendation.

(e) CN + OH −→ HOCN −→ H + NCO.

This reaction is about 125 kJ/mol exothermic, suggesting that virtually all of theHOCN adduct dissociates to H + NCO. An estimate can be made by analogy tothe recombination rate coefficient of CH3 + OH. Using the Baulch et al. (1992)rate coefficient for CH3 + OH we infer for CN + OH

k44e = 4× 1013 cm3mol−1s−1,

Page 143: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 267

which is close to the value suggested by Tsang (1992) and identical to the valuemeasured by Wooldridge et al. (1996) over the temperature range 1250 to 1860 K.

(f) CN + HCN −→ HC(CN)N −→ NCCN + H.

This can be viewed as an addition-elimination reaction; overall it is exothermicby about 42 kJ/mol. One might again expect to see some falloff in rate athigh temperatures due to onset of the reverse reaction. This is not reflected,however, in the data summarized by Tsang (1992). We suggest adoption of hisrecommendation,

k44f = 1.5× 107T 1.71 exp(−770/T ) cm3mol−1s−1,

which is consistent with the results of Wooldridge et al. (1996).

(g) CN + N2O −→ Products.

CN might be expected to add at either end of the molecule, in a manneranalogous to H + N2O. (Sections 6.3 through 6.6) Addition to the O end wouldlead to NNOCN, which could dissociate to NCO + N2. This overall reaction isexothermic by about 360 kJ/mol. With such a low energy product channel onemight expect the rate-limiting step to be the adduct formation. Addition to the Nend would give NCNNO, which would have to isomerize first in order to access alow energy product channel. Wang et al. (1991) report for the temperature rangefrom 500 to 750 K k = 9.5×1011 exp(−3560/T ) cm3mol−1s−1.They proposethat the barrier to form NNOCN is sufficiently high that NCN + NO is the dom-inant reaction channel at their conditions. This analysis is consistent with themeasurements of Williams et al. (1995), who report that these are the dominantproducts from 400 to 870 K, with the NCO + N2yield being less than 10%. Therewas some pressure dependence, the measurement at 100 Torr being expressedas k = 6.0× 1013 exp(−7730/T ) + 1.9× 109 exp(−730/T ) cm3mol−1s−1,

about a factor of 3 smaller than the Wang et al. expression. Since thisexpression includes a contribution from the stabilization channel, we extractedan estimate for the NCN + NO channel from the low-pressure, high-temperaturemeasurements alone:

k44g = 4.2× 1011 exp(−3610/T ) cm3mol−1s−1.

At higher temperatures additional reaction to form NCO + N2 is to be expected.The total rate coefficient is nonetheless small enough that this reaction is notlikely to be important in combustion.

(h) CN + NO2 −→ NCO + NO.

The simplest way for this reaction to lead to these products is for CN to addto one of the oxygens to form NCONO, which would then dissociate to NCO+ NO. Measurements by Park and Herschberger (1993b) indicate that thischannel accounts for 87% of the total reaction. The rate coefficient has also

Page 144: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

268 Anthony M. Dean, Joseph W. Bozzelli

been measured at higher temperatures in shock tube experiments by Wooldridgeet al. (1994). Combining their data with that of Park and Herschberger and thatof Wang et al. (1989), Wooldridge et al. report

k44h = 6.2× 1015T−0.752 exp(−170/T ) cm3mol−1s−1

for temperatures from 300 to 1600 K. This very fast rate suggests that CNaddition to NO2 behaves as radical recombination, quite reasonable consideringthat NO2 is indeed a radical.

(i) CN + CH4 −→ HCN + CH3.

We include this reaction as representative of CN abstractions that can takeplace in a hydrocarbon environment. Methane should not be taken as fullyrepresentative of all hydrocarbons, however, and this rate coefficient would notbe appropriate for hydrogen abstraction from hydrocarbons having H–C bondstrengths different from that of methane. There have been two rate measure-ments, in reasonable agreement with one another over extended temperatureranges, by Atakan and Wolfrum (1991) and Balla et al. (1991). We suggestadoption of the Atakan and Wolfrum expression

k44i = 1.2× 105T 2.64 exp(+80/T ) cm3mol−1s−1.

(j) CN + NH3 −→ HCN + NH2.

Sims and Smith (1988) derived the rate coefficient expression

k44j = 9.2× 1012 exp(+180/T ) cm3mol−1s−1

from experimental measurements over the temperature range from 294 to 761 K.The dramatic difference in the temperature dependence of the reactions of CNwith methane and ammonia bears some discussion. Sims and Smith studied theammonia reaction over a wide temperature range, so the observed negativetemperature dependence cannot be dismissed as based upon a too-limitedinvestigation. Nonetheless, negative temperature dependence is very unusualfor an abstraction reaction. In fact, it is really only consistent with an additionmechanism. Calculations by Meads et al. (1993) support formation of anH3NCN complex and subsequent rearrangement to yield the same productsone would expect from a simple abstraction. Meads et al. computed arate coefficient and showed that it decreases with increasing temperature, inqualitative agreement with the experimental data.

Page 145: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 269

2.7.20 Reactions of H2CN

We have seen earlier that H2CN is predicted to be a product of several reactions,including CH3 + NO (Section 6.15). Dissociation to HCN + H is only 106kJ/mol endothermic, so rapid dissociation at combustion temperatures shouldoccur. At lower temperatures, one might expect H2CN to transfer a weakly-bonded H to radicals in a disproportionation reaction. Alternatively, radical–radical reactions could result in formation of an energized adduct and subsequentfragmentation. Most abstractions by H2CN would be endothermic and thereforerelatively unimportant, because the strength of the H–N bond formed in such areaction would be only about 370 kJ/mol.

(a) H2CN −→ HCN + H.

We are not aware of any measurements of the rate coefficient of this reaction.QRRK calculations indicate that it is in the falloff regime for combustionconditions, with rate coefficients for N2 bath gas, in s−1 units,

k45a = 1.3× 1029T−6.03 exp(−15000/T ) 0.1 atm

= 6.0× 1031T−6.46 exp(−16200/T ) 1.0 atm

= 3.5× 1029T−5.46 exp(−16400/T ) 10 atm.

(b) H2CN + HO2 −→ CH2NO + OH.

By analogy to hydrocarbon radical reactions with HO2 one expects a rapidaddition-dissociation reaction to break the weak O–O bond. On this basis weestimate

k45b = 3× 1013 cm3mol−1s−1.

which can be compared to the 2.4 × 1013 cm3mol−1s−1 value suggested byTsang and Hampson (1986) for the analogous channel of the C2H5 + HO2reaction. DHT estimates for the H transfer reactions are included in Table 19;these approach the adduct formation value at combustion temperatures.

(c) H2CN + O2 −→ CH2O + NO.

Reaction by an addition pathway would be expected to result, by analogy toC2H3 + O2, in these products, as formation of CH2NO + O is endothermicby about 170 kJ/mol and therefore unimportant. The initially formed adductshould be appreciably less stable than C2H3OO, which implies a larger barrierfor the cyclization to yield CH2O + NO. From these considerations we estimate

k45c = 3× 1012 exp(−3000/T ) cm3mol−1s−1.

(d) H2CN + CH3 −→ HCN + CH4.

This is an example of H2CN participating in radical disproportionation reac-tions. The DHT rate coefficient estimate is

k45d = 8.1× 105 T 1.87 exp(+560/T ) cm3mol−1s−1.

Page 146: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

270 Anthony M. Dean, Joseph W. Bozzelli

(e) H2CN + OH −→ HCN + H2O.

This reaction is analogous to C2H5 + OH. For the reaction with C2H5 one canrationalize the observed very fast rate in terms of radical recombination followedby elimination of H2O. A significant difference, however, is that the well depthof the H2CNOH complex is not as deep as that of CH3CH2OH; consequently, thebarrier for concerted H2O elimination is closer to the entrance channel. For thisreason one expects a falloff in the forward rate at higher temperatures as moreof the adduct dissociates through the entrance channel. QRRK calculations forH2CN reacting with OH in N2 bath gas yield

k45e = 2.1× 1017T−1.68 exp(−160/T ) cm3mol−1s−1 0.1 atm

= 1.5× 1019T−2.18 exp(−1090/T ) cm3mol−1s−1 1.0 atm

= 9.5× 1021T−2.91 exp(−2830/T ) cm3mol−1s−1 10 atm.

These expressions give appreciably smaller rate coefficients than observed forOH + C2H5 and there is noticeable falloff for combustion conditions. TheQRRK results also show an inverse pressure dependence, especially at lowtemperatures, where the adduct is in part stabilized by collisions.The same products can also be formed directly. The DHT estimate is

k45e2 = 1.2× 106 T 2 exp(+600/T ) cm3mol−1s−1.

At 2000 K direct pathway is calculated to be about 10 times faster than the adductformation pathway. Both should be considered important under combustionconditions.

(f) H2CN + N −→ H2CNN −→ CH2 + N2.

This reaction is very exothermic and thus expected to be fast. Nesbitt etal. (1990) measured the rate of reactant disappearance over the temperaturerange from 200 to 363 K; assuming that the reaction has only the channelshown gives

k45f = 6.0× 1013 exp(−200/T ) cm3mol−1s−1.

(g) H2CN + H −→ HCN + H2.

The DHT estimate is

k45g = 2.4× 108 T 1.5 exp(+450/T ) cm3mol−1s−1.

(h) H2CN + NH2−→ HCN + NH3.

The DHT estimate is

k45h = 9.2× 105 T 1.94 exp(+580/T ) cm3mol−1s−1.

Page 147: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 271

(i) H2CN + O −→ HCN + OH

The DHT estimate is

k45i = 1.7× 108 T 1.5 exp(+450/T ) cm3mol−1s−1.

Adduct formation leads to CH2NO. (Figure 51) QRRK analysis shows thedominant channels to be H atom loss

H2CN+O −→ HNCO+H, k45i2 = 6× 1013 cm3mol−1s−1

−→ HCNO+H, k45i2 = 2× 1013 cm3mol−1s−1

2.7.21 Reactions of HCNH

This species is formed in the CH3 + N reaction (Section 6.16). Dissociation byloss of the weakly bound N-hydrogen

HCNH −→ HCN+ H 1r H = +75 kJ/mol (46)

should have a rate coefficient intermediate between those for dissociation of NNH(Section 7.3) and HNO (Section 7.11). The QRRK estimate gave for N2 bath gas

k46 = 7.7× 1025T−5.20 exp(−11100/T ) s−1 0.1 atm

= 6.1× 1028T−5.69 exp(−12200/T ) s−1 1.0 atm

= 6.2× 1026T−4.77 exp(−12500/T ) s−1 10 atm.

Under typical combustion conditions dissociation will be very rapid and otherreactions of HCNH unimportant. However, as with NNH, important exceptionscould arise in flame fronts where the mole fractions of H, OH, and O are high,approaching 1%, or concentrations of about 60 nmol/cm3 at 1 atm. At suchconcentrations, a bimolecular rate coefficient for reaction with a flame radicalover 1013 cm3mol−1s−1 will imply a first order loss rate of HCNH above 106 s−1,approaching the thermal dissociation loss rate, which at 2000 K and 1 atm isestimated to be about 2 × 107 s−1. This is somewhat slower than for NNHdissociation, suggesting a greater chance for reactions of HCNH with flameradicals. We accordingly estimated these rate coefficients. Some of the estimatesproved to be large enough to suggest that such bimolecular reactions can occur.

(a) HCNH + H −→ Products.

Formation of H2CNH liberates about 400 kJ/mol, providing ample energy forproduction of either CH2N + H or HNC + H2. Although the barrier for H2elimination is probably lower than that for production of CH2N + H, we expectthe latter, with its higher A-factor,will be most important in the high temperatureregion where this reaction might occur. We estimate

HCNH+ H −→ CH2N+H k46a1 = 2× 1013 cm3mol−1s−1

Page 148: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

272 Anthony M. Dean, Joseph W. Bozzelli

from taking half the rate coefficient proposed by Baulch et al. (1992) for H +C2H5 −→ 2 CH3, which has similar energetics but would be expected to havea greater A-factor for adduct dissociation. The DHT estimate for the directreaction is, in cm3mol−1s−1 units

HCNH+H −→ HCN+ H2 k46a2 = 2.4× 108 T 1.5 exp(+450/T ) .

(b) HCNH + O −→ Products.

The adduct HC(O)NH can dissociate rapidly to HNCO + H, which is overall410 kJ/mol exothermic. We suggest that the rate coefficient be taken as theBaulch et al. (1992) recommendation for O + C2H5

HCNH+O −→ HNCO+H k46b1 = 7× 1013 cm3mol−1s−1.

The DHT estimate for the direct reaction is, in cm3mol−1s−1 units

HCNH+ O −→ HCN+ OH k46b2 = 1.7× 108 T 1.5 exp(+450/T ) .

(c) HCNH + OH −→ Products.

Formation of H2O from the HC(OH)NH adduct has a substantial barrier and alow A-factor. The same products can be formed by direct reaction, for whichthe DHT estimate is, in cm3mol−1s−1 units

HCNH+OH −→ HCN+ H2O k46c = 1.2× 106 T 2 exp(+600/T ) .

(d) HCNH + CH3 −→ HCN + CH4.

The DHT estimate is

k46d = 8.2× 105 T 1.87 exp(+560/T ) cm3mol−1s−1.

2.7.22 Reactions of HCNN

This is the stable adduct formed in the reaction between CH and N2 (Section6.20). It redissociates quickly to reactants under combustion conditions, witha rate coefficient for dissociation that can be inferred from the formation ratecoefficient. Because of the relatively weak C–N bond this channel will be so fastthat reactions with other radicals should be unimportant. The reaction with O2,however, should be considered. The adduct for this reaction can give two sets ofproducts

HCNN+ O2 −→ HCO2 +N2 1r H = −615 kJ/mol (47a)

−→ HCO+N2O 1r H = −330 kJ/mol. (47b)

The highly exothermic nature of these reactions suggests use of rate coefficientsnear that anticipated for formation of the initial adduct,

k47a = k47b = 4× 1012 cm3mol−1s−1.

HCO2 should rapidly dissociate to H + CO2.

Page 149: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 273

2.7.23 Reactions of H2CNH

This species can be formed by the reaction of N with CH3 (Section 6.16) and isalso expected to play a role in methylamine pyrolysis and oxidation. Hydrogenabstractions from it can leave two different radicals, i.e.,

H2CNH+X −→ H2CN+HX (48a)

−→ HCNH+HX. (48b)

The strength of the weaker H–N bond is about 370 kJ/mol. For estimating ratecoefficients for abstracting from it we used the DHT method and obtained the ratecoefficient expressions in cm3mol−1s−1 units:

H2CNH + H −→ H2CN + H2 k48a1 = 2.4× 108T 1.5 exp(−3685/T )+ O −→ H2CN + OH k48a2 = 1.7× 108T 1.5 exp(−2330/T )

+ OH −→ H2CN + H2O k48a3 = 1.2× 106T 2 exp(+45/T )+ CH3 −→ H2CN + CH4 k48a4 = 8.2× 105T 1.87 exp(−3585/T )+ NH2 −→ H2CN + NH3 k48a5 = 9.2× 105T 1.94 exp(−2235/T ).

For abstraction of an H attached to the C atom the bond strength is about 400kJ/mol, compared to 465 kJ/mol in ethylene, reflecting resonance stabilization ofthe radical. The DHT estimate was corrected accordingly. (Cf. Section 2.4) TheDHT estimates are, in cm3mol−1s−1 units,

H2CNH + H −→ HCNH + H2 k48b1 = 3.0× 108T 1.5 exp(−3085/T )+ O −→ HCNH + OH k48b2 = 2.2× 108T 1.5 exp(−2720/T )

+ OH −→ HCNH + H2 O k48b3 = 2.4× 106T 2 exp(−230/T )+ CH3 −→ HCNH + CH4 k48b4 = 5.3× 105T 1.87 exp(−4875/T )+ NH2 −→ HCNH + NH3 k48b5 = 1.8× 106T 1.94 exp(−3065/T ).

(c) H2CNH + O −→ CH2O + NH.

This reaction is exothermic by 100 kJ/mol and expected to be fast. We estimateit to have half of the addition rate of O-atoms to ethylene and derive from theBaulch et al. (1992) expression

k48c = 1.7× 106T 2.08 cm3mol−1s−1.

2.7.24 Reactions of CH3NH

The weak C–H bond suggests that dissociation

CH3NH −→ CH2NH+H 1r H = 125 kJ/mol

is the most important reaction of CH3NH under combustion conditions. TheQRRK method predicts for the temperature range from 600 to 2500 K in nitrogen

k49 = 1.6× 1036T−7.92 exp(−18300/T ) s−1 0.1 atm

= 1.3× 1042T−9.24 exp(−20800/T ) s−1 1.0 atm

= 2.3× 1044T−9.51 exp(−22800/T ) s−1 10 atm.

Page 150: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

274 Anthony M. Dean, Joseph W. Bozzelli

For combustion conditions this should be the only reaction of interest. We exploredthe possibility of addition to O2 by QRRK analysis. The combination of a shallowpotential well (estimated by analogy to NH2 + O2, Section 6.9, to be about 35kJ/mol) and a high A-factor for dissociation of the adduct to reactants results inlow rate coefficients for both product channels, i.e., to CH3NHO + O and H2CNH+ HO2. At 2000 K the computed rate coefficients are less than 109 cm3mol−1s−1,indicating pseudo-first order rate coefficients in air on the order of 103 s−1. Incontrast, the dissociation rate coefficient for these conditions is about 1×107 s−1.The shallow well also insures that reactions of the stabilized adduct can also beneglected. DHT estimates of the rate coefficients for direct hydrogen transfer toH, O, OH and CH3 have been included in Table 19.

2.7.25 Reactions of CH2NH2

The weak N–H bond means that CH2NH2 is another species that can be expectedto dissociate rapidly

CH2NH2 −→ CH2NH+ H 1r H = 160 kJ/mol

at combustion temperatures. QRRK analysis gives for the temperature range from600 to 2500 K and N2 bath gas the rate coefficient expressions

k50 = 1.1× 1045T−10.24 exp(−24100/T ) s−1 0.1 atm

= 2.4× 1048T−10.82 exp(−26200/T ) s−1 1.0 atm

= 3.2× 1046T−9.95 exp(−26900/T ) s−1 10 atm.

Abstractions to form CH3NH2 are unlikely to compete with thermal dissociationunder combustion conditions, as the strength of the C–H bond formed is only about390 kJ/mol, so most abstractions would be endothermic. For most conditionsradical concentrations are too small to allow reactions of CH2NH2 with radicals tocompete with dissociation. For completeness we include the following estimates.

(a) CH2NH2 + O2.

QRRK analysis of the addition of the carbon-centered radical to O2 indicatedthat the major reaction channel is intramolecular hydrogen transfer and lossof HO2, analogous to the C2H5 + O2 reaction, and that the reaction is nearits low-pressure limit for combustion conditions. The QRRK analysis gave incm3mol−1s−1 units

CH2NH2 +O2 −→ H2CNH+HO2 k50a = 1.0×1022T−3.09 exp(−3400/T ).

The QRRK estimate for the less important channel forming O + NH2CH2O isalso included in Table 19; because NH2CH2O should rapidly dissociate to NH2+ CH2O, these products are assumed to be produced directly.

Page 151: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 275

(b) CH2NH2 + H −→ Products.The adduct formation pathway leading to CH3 + NH2 is the reverse of reaction21b. The DHT estimate for the hydrogen transfer reaction forming H2CNH +H2 is

k50b = 4.8× 108 T 1.5 exp(+450/T ) cm3mol−1s−1.

(c) CH2NH2 + O −→ Products.

Adduct formation followed by dissociation to CH2O + NH2 is 320 kJ/molexothermic. We estimate its rate coefficient to be

k50c1 = 7× 1013 cm3mol−1s−1,

the Baulch et al. (1992) value for the rate coefficient of the association reactionof O-atoms with C2H5. The DHT estimate gives, in cm3mol−1s−1 units,

CH2NH2 + O −→ H2CNH+OH k50c2 = 3.3× 108 T 1.5 exp(+450/T ) .

(d) CH2NH2 + OH −→ Products.

The adduct formation pathway has a 92 kJ/mol exit channel producing CH2OH+ NH2. We suggest

k50d1 = 4× 1013 cm3mol−1s−1,

the Baulch et al. (1992) rate coefficient for the CH3 + OH reaction. The DHTestimate for the direct reaction producing H2CNH + H2O is

k50d2 = 2.4× 106 T 2 exp(+600/T ) cm3mol−1s−1.

(e) CH2NH2 + CH3 −→ Products.

Assuming that the adduct formation rate coefficient is 3× 1013 cm3mol−1s−1,

the Baulch et al. (1992) recommendation for the high pressure limiting ratecoefficient of CH3 + C2H5, and that the A-factors for dissociation into the twopossible exit channels are equal, leads to

k50e1 = 2× 1013 exp(−1360/T ) cm3mol−1s−1

for the 12 kJ/mol endothermic reaction forming C2H5 + NH2. At lowertemperatures and higher pressures some degree of adduct stabilization is tobe expected. The DHT estimate for the direct reaction producing H2CNH +CH4 is

k50e2 = 1.6× 106 T 1.87 exp(+315/T ) cm3mol−1s−1.

Page 152: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

276 Anthony M. Dean, Joseph W. Bozzelli

2.7.26 Reactions of CH3NH2

This recombination product of CH3 and NH2 (Section 6.17) can undergo hydrogentransfer reactions as well as dissociation. The dissociation rate coefficient can becomputed from the previously estimated recombination rate coefficient (Section6.17). For flame front radical concentrations approaching 5 × 10−8 mol/cm3,abstraction rate coefficients above 1012 cm3mol−1s−1 will imply abstraction ratesthat compete with thermal dissociation.

Hydrogen abstraction forms two products

CH3NH2 +X −→ CH2NH2 +HX

−→ CH3NH+HX

The site of radical attack is affected by the C–H and N–H bond dissociationenergies as follows. The C–H bond dissociation energy of about 395 kJ/mol issome 30 kJ/mol less than typical primary C–H bond dissociation energies becauseof stabilization of the radical center formed by the adjacent nitrogen atom. TheN–H bond dissociation energy is about 423 kJ/mol. Using these values and theDHT estimation method leads to the expressions listed in Table 19. (The Table 3coefficients were adjusted to account for the partial resonance stabilization thataccounts for the weakening of the C–H bond.) At 2000 K and 1 atm, with radicalconcentrations approaching 1%, these expressions for H, O and OH imply thatboth abstraction rates are faster than thermal dissociation. Abstractions by NH2are about as fast as thermal dissociation while abstractions by CH3 are somewhatslower.

2.7.27 Reactions of NCCN

Because the C–C bond in cyanogen is very strong, with 1r H ◦ = 560 kJ/molfor dissociation, its dissociation rate is expected to be slow, as was confirmed inshock tube experiments by Natarajan et al. (1986). Their low pressure limit ratecoefficient in argon test gas is given by

k52a = 1.1× 1034 T−4.32 exp(−65465/T ) cm3mol−1s−1.

At 2000 K and 1 atm the corresponding first order rate coefficient is 2 s−1, far tooslow to be relevant in combustion. The most likely reactions would be addition-elimination sequences such as

NCCN + H −→ NCHCN −→ HCN + CN 1r H = +42 kJ/mol (52b)+ O −→ NCOCN −→ NCO + CN 1r H = +8 kJ/mol (52c)

+ OH −→ NCOHCN −→ HOCN + CN 1r H = +75 kJ/mol (52d)The reaction with H atoms is the reverse of reaction 44f. For the reaction with Oatoms, Louge and Hanson (1984a, b) reported k52c = 5× 1011 cm3mol−1s−1 at2000 K, which is consistent with the Miller and Bowman (1986) expression

k52c = 4.6× 1012 exp(−4470/T ) cm3mol−1s−1.

Page 153: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 277

We are not aware of any high temperature studies of reaction 52d. Millerand Bowman use an expression taken from low-temperature measurements that isappropriate for formation of the adduct but not for formation of the products HOCN+ CN, which would not be accessible in the low-temperature experiments. Becauseof the relatively high endothermicity of reaction 52d it should be unimportant andcan safely be omitted from combustion models. We estimate

k52d = 2× 1012 exp(−9500/T ) cm3mol−1s−1

by analogy to QRRK results with other OH addition reactions.

2.7.28 Reactions of NCO

Reactions of NCO are important when cyanuric acid is used as an NOx abatementagent in the RAPRENOX process (Perry and Siebers 1986), wherein HNCOproduced from thermal decomposition of cyanuric acid loses an H-atom byabstraction. It can also be formed by the addition of O to HCN (cf. Section7.17) or in reactions of OH or O2 with CN (cf. Section 7.19). NCO is a relativelystable radical, about 230 kJ/mol more stable than N + CO, and thus has a relativelylong lifetime. It can react with other radicals, with a high probability of breakingthe N–C bond to form CO. Since the H–N bond in HNCO is so strong (about 470kJ/mol), one expects that NCO can abstract hydrogen atoms from most species toproduce HNCO.

(a) NCO + NO −→ Products.

The formation of N2 from reaction of NCO with NO is one of the reactionsby which addition of cyanuric acid to combustion systems can remove nitricoxides. Other channels are also possible for this reaction, namely

NCO + NO −→ CO2 + N2 1r H = −620 kJ/mol (53a1)−→ CO + N2O 1r H = −250 kJ/mol (53a2)−→ CO + N2 + O 1r H = −84 kJ/mol (53a3)

Mertens et al. (1992) combined their shock tube measurements of the overallrate coefficient with lower temperature data and reported

ktotal = 1.4× 1018T−1.73 exp(−384/T ) cm3mol−1s−1

for the temperature range from 294 to 2660 K. The rate coefficient decreasesby about an order of magnitude over this temperature range, consistent withintermediate formation of a chemically-activated adduct.Product distributions were measured at room temperature by Cooper and Her-schberger (1992) and Becker et al. (1992b). These groups report that the N2Ochannel is 33 to 35% of the total; Cooper and Herschberger also report that theCO2 channel is 44% of the total and the CO + O + N2 is 22%. Measurements

Page 154: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

278 Anthony M. Dean, Joseph W. Bozzelli

by Cooper et al. (1993) extended the temperature range for branching ratiomeasurements to 623 K and revised their 1992 measurements such that the CO2+ N2 pathway accounts for 56±7% and N2O + CO for the remaining 44±7%,with no observed temperature dependence of the branching ratio. A theoreticalanalysis by Lin et al. (1993) can be made consistent with the data of Cooper etal. by a reduction in the barrier height for dissociation of the OCNNO complexto N2O + CO. We suggest use of the above overall rate coefficient expressionof Mertens et al together with the temperature-independent branching ratiosmeasured by Cooper et al. Measurements by Flatness and Kramlich (1996)gave a branching ratio for the N2O channel in the 30 to 50% range from 1100 to1400 K range, which is consistent with the lower temperature measurements.

(b) NCO + M −→ N + CO + M.

Baulch et al. (1992) recommend k = 1.0×1015 exp(−23500/T) cm3mol−1s−1

for M = Ar. At 1500 K and 1 atm this corresponds to a pseudo-first-order ratecoefficient of 0.5 s−1. Using the enthalpy of formation value for NCO proposedby East and Allen (1993), however, implies that this reaction is 230 kJ/molendothermic. The activation energy of Baulch et al. is thus probably too low.We recommend the expression of Mertens et al. (1996) for dissociation in Arbath gas

k53b = 2.2× 1014 exp(−27200/T ) cm3mol−1s−1

for the temperature range from 2370 to 3050 K. This expression is moreconsistent with the East and Allen thermochemistry. This reaction is too slowto be important in combustion except at very high temperatures.

(c) NCO + H2 −→ HNCO + H.

Miller and Melius (1992b) calculated the transition state theory expression

k53c = 7.6× 102T 3 exp(−2000/T ) cm3mol−1s−1

which is in reasonable agreement with the limited data available.

(d) NCO + O −→ ONCO −→ NO + CO.

This reaction is expected to proceed by adduct formation followed by breakingof the weak N–C bond. Baulch et al. (1992) recommend

k53d1 = 4.2× 1013 cm3mol−1s−1,

which is consistent with this pathway. O atoms can also add to the C–O bond,and the adduct can dissociate exothermically to N and CO2. We estimate therate coefficient for this reaction to be half of that for O addition to ethylene

k53d2 = 8× 1012 exp(−1300/T ) cm3mol−1s−1 .

Page 155: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 279

(e) NCO + H −→ HNCO −→ NH + CO

Baulch et al. (1992) recommend

k53e = 5.2× 1013 cm3mol−1s−1,

which is also consistent with a adduct formation-elimination mechanism.

(f) NCO + N −→ NNCO −→ N2 + CO.

Brownsword et al. (1997) reported a room temperature rate coefficient

k53f = 3.3× 1013 cm3mol−1s−1.

While this reaction is spin-forbidden, it is also highly exothermic, by 710kJ/mol, which explains the large rate coefficient value.

(g) NCO + OH −→ Products.

Tsang (1992) considered this reaction to be an exothermic abstraction to formO + HNCO and recommended, based on the reverse reaction, the expression incm3mol−1s−1 units

NCO+OH −→ O+HNCO k53g1 = 7.8× 104T 2.27 exp(+495/T ).

These radicals can also recombine to form HONCO and then dissociate via twochannels, one to HON + CO and the other to H + NO + CO. QRRK analysisgave the expressions in cm3mol−1s−1 units

NCO + OH−→ CO + HON k53g2 = 5.3× 1012T−0.07 exp(−2600/T )−→ H + CO + NO k53g3 = 8.3× 1012T−0.05 exp(−9080/T ).

(h) NCO + NO2 −→ Products.This reaction is exothermic by 500 kJ/mol if the products are N2O and CO2.Other pathways, such as CO + N2 + O2, or CO + 2 NO, are also thermodynam-ically accessible. Park and Herschberger (1993a) report a total rate coefficientand a product distribution over the temperature range from 298 to 500 K. Wesuggest using their expressions, which are in cm3mol−1s−1 units

NCO+ NO2 −→ CO2 +N2O k53h1 = 2.3× 1012 exp(+440/T )

−→ CO+ 2 NO k53h2 = 2.1× 1011 exp(+440/T ).

At 1250 K these expressions yield 3.6× 1012 cm3mol−1s−1 for the total ratecoefficient, in good agreement with the 4.5×1012 cm3mol−1s−1 value measuredby Wooldridge et al. (1994). The total rate coefficient into both channelswas also measured by Juang et al. (1995) from 294 to 774 K, who reportan expression about 60% higher than the Park and Herschberger result, with

Page 156: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

280 Anthony M. Dean, Joseph W. Bozzelli

an indication of non-Arrhenius behavior, that extrapolates smoothly to theWooldridge et al. value.

(i) NCO + CH4 −→ HNCO + CH3.

We include this reaction to represent abstractions from hydrocarbons. (Asnoted in Section 7.19 in connection with the CN reaction, methane is not arepresentative hydrocarbon; for other hydrocarbons this expression must bemodified to account for different C–H bond strengths.) This reaction has beenstudied in a pump-probe experiment over the temperature range from 512 to1113 K by Schuck et al. (1994), who report

k53i = 9.8× 1012 exp(−4090/T ) cm3mol−1s−1.

Studies of NCO reactions with other hydrocarbons have been reported (Parkand Herschberger 1994; Becker et al. 1995a, 1995b).

(j) NCO + NH3 −→ HNCO + NH2.

This reaction is approximately thermoneutral, while abstraction of an H atomby NCO from CH4 is about 13 kJ/mol exothermic, and it has fewer H atoms toabstract than does CH4. Nonetheless, its rate coefficient is reported by Beckeret al. (1997) to be faster than that of the corresponding abstraction (53i) frommethane.

k53j = 2.8× 104 T 2.48 exp(−495/T ) cm3mol−1s−1.

2.7.29 Reactions of HCNO

This species has been reported to be the major product of the CH2 + NOreaction (Section 6.19). Thus one needs to account for its subsequent reactions.Dissociation to HCN + O is endothermic by 210 kJ/mol, so this pathway will berelatively slow. QRRK analysis indicates that this reaction is in the falloff regimefor combustion conditions and has the rate coefficient expressions in N2 bath gas

k54 = 2.0× 1030T−6.03 exp(−30600/T ) s−1 0.1 atm

= 4.2× 1031T−6.12 exp(−30800/T ) s−1 1.0 atm

= 5.9× 1031T−5.85 exp(−31200/T ) s−1 10 atm.

Abstraction reactions by HCNO should be unimportant, because the strength ofthe C–H bond formed would only be about 220 kJ/mol. Recombination reactionscould be important. QRRK calculations with the parameters shown in Table 15gave, in cm3mol−1s−1 units,

HCNO+H −→ HNCO+ H k54a = 2.1× 1015T 0.69 exp(−1435/T )

−→ HCN+ OH k54b = 2.7× 1011T 0.18 exp(−1065/T )

−→ NH2 + CO k54c = 1.7× 1014T−0.75 exp(−1455/T )

−→ HOCN+ H k54d = 1.4× 1011T−0.19 exp(−1250/T ).

Page 157: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 281

Reaction 54a is the most important channel. Stabilization channels are notimportant under combustion conditions. Other radical recombinations, followedby dissociation, include

HCNO + O −→ HC(O)NO −→ HCO + NO 1r H = −290 kJ/mol (54e)HCNO + OH −→ HC(OH)NO −→ HCOH + NO 1r H = −30 kJ/mol. (54f)

We expect these rate coefficients to reflect the recombination rates, as in theCH2NH2 reactions discussed in Section 7.25:

k54e = 7× 1013 cm3mol−1s−1

k54f = 4× 1013 cm3mol−1s−1.

The reaction with OH has a more exothermic exit channel, to CH2O + NO, butQRRK analysis confirms that isomerization of HCOH to CH2O is slow, due to thelow A-factor caused by loss in entropy in forming a 3-member cyclic transitionstate. and the high barrier. Decomposition of the HC(OH)NO adduct to HCOHand NO is expected to be the only output channel.

2.7.30 Reactions of HOCN

HOCN can be formed by the addition of OH to HCN, as discussed in Section 7.17.It can form NCO by loss of the hydrogen; the relatively weak, about 360 kJ/mol,O–H bond strength means that such abstractions would be exothermic and fast.Radicals could also add across the C–N bond. One such reaction is the additionof hydrogen atoms to form the adduct NCHOH (cf. Fig. 51). QRRK analysispredicts for this adduct formation pathway, in cm3mol−1s−1 units,

HOCN+ H −→ HCN+ OH k55a = 2.0× 1013T−0.04 exp(−1075/T )

−→ HNCO+ H k55b = 3.1× 108T 0.84 exp(−965/T )

−→ NH2+ CO k55c = 1.2× 108T 0.61 exp(−1045/T ).

The first channel is the fastest. It is the reverse of reaction (42b2), and the samerate coefficient can also be inferred from the rate coefficient expression for k42b2;accordingly, combustion modeling programs that automatically include reversereactions should not have reactions (42b2) and (55a) both present.

Absent experimental data for the abstraction reactions, we estimated their directpathway rate coefficients by the DHT method. The results are listed in Table 19.Abstractions by H, O and OH compete with the H atom addition channel atcombustion temperatures.

2.7.31 Reactions of HNCO

HNCO is the initial decomposition product of cyanuric acid, the selective non-catalytic reducing agent of the Raprenox process for NOx removal. (Perry and

Page 158: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

282 Anthony M. Dean, Joseph W. Bozzelli

Siebers 1986) It is also predicted to be the major channel for the CH2 + NOreaction (Section 6.19). Spin-forbidden thermal dissociation to the lowest energyproducts

HNCO+M −→ 3NH+ CO+M 1r H56a = 350kJ/mol

has been studied in shock tubes by Mertens et al. (1989) and Wu et al. (1990)with concordant results. We suggest adoption of the Wu et al. expression

k56a = 8.4× 1015 exp(−42640/T) cm3mol−1s−1

valid for argon bath gas over the temperature range from 1830 to 3340 K. Themeasured activation energy suggests that the barrier for reverse reaction is small,which implies that there is no significant energy barrier on the surface for spinreversal.

HNCO participates in radical addition and abstraction reactions. QRRK analy-sis of hydrogen addition, which includes formation of either NCHOH or HNCHO(cf. Fig. 51), predicts the dominant channel to be

HNCO+ H −→ NH2 + CO (56b)

and a rate coefficient expression comparable to literature expressions, slightlyhigher at 1000 K than the expression suggested by Tsang (1992), but lower thanhis suggestion at 2000 K; and somewhat higher than the measurements of Mertenset al. (1991b) and the calculations of Miller and Melius (1992b). The Millerand Melius analysis confirms that addition dominates abstraction. Nguyen et al.(1996) computed a potential surface for reaction (56b) and from it QRRK ratecoefficient expressions with tunneling corrections. We suggest adoption of theirexpression

k56b = 3.6× 104 T 2.49 exp(−1180/T ) cm3mol−1s−1.

Reaction with O has two exothermic adduct formation pathways:

HNCO+O −→ HN(O)CO −→ HNO+ CO 1r H (56c1)= –160 kJ/mol

−→ HNC(O)O −→ NH+ CO2 1r H (56c2)= –190 kJ/mol

Assuming that adduct formation is rate limiting and that the rate coefficient is thesame for oxygen atom addition to either N or C, we estimate

k56c1 = k56c2 = 1.7× 106T 2.08 cm3mol−1s−1

by taking half of the expression recommended for O–atom addition to ethyleneby Baulch et al. (1922). This expression gives values significantly higher than theexperimental results for k56c2 reported by Mertens et al. (1992a).

Page 159: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 283

OH addition to the carbon leads to HNC(OH)O, which can form NH2 + CO2after a hydrogen shift. Similarly, addition to nitrogen yields HN(OH)CO, whichcan form NCO + H2O. QRRK analysis suggests the rate coefficient expressionsin cm3mol−1s−1 units

HNCO+OH −→ NH2 + CO2 k56d1 = 6.3× 1010T−0.06 exp(−5860/T )

−→ NCO+H2O k56d2 = 5.2× 1010T−0.03 exp(−8840/T ).

These expressions are much lower than those for O–atom addition, reflecting theisomerizations required to reach the low energy exit channels in these reactions.

Direct abstractions from HNCO should be similar to abstractions from NH3,except for there being but one abstractable hydrogen atom, as the bond strengthsare the same. The reactions with H, O, CH3 and NH2 have appeared already asreverse reactions in Section 7.28. Wooldridge et al. (1996) report an expressionfor the OH abstraction rate in cm3mol−1s−1:

HNCO+OH −→ NCO+H2O k56g = 3.6× 107 T 1.5 exp(−1810/T ).

The abstraction products are the same as for one of the OH addition pathways(56d2). Direct abstraction is faster.

2.7.32 Reactions of CH2NO

The endothermicity for loss of an H-atom in direct dissociation of CH2NO

CH2NO −→ HCNO+H 1r H = + 222 kJ/mol

is comparable to that for the corresponding process in HNOH (cf. Section 7.13),and so we expect similar kinetics. In the CH2NO case, however, there are otherpossibilities, as illustrated in Fig. 51. From our earlier discussion of CH2 + NO(Section 6.19) we expect

CH2NO −→ HNCO+H 1r H = −60kJ/mol

to be important even though several isomerizations precede final bond-breaking.The long thermal dissociation lifetime of CH2NO suggests that radical-radicalreactions with it may be important.

(a) CH2NO −→ Products.

QRRK analysis indicates that CH2NO thermal dissociation is in the falloffpressure regime under combustion conditions and that HNCO + H is thedominant product channel. A good fit to the rate coefficient is obtained byrestricting the temperature range to 600 to 2500 K, for which the analysis gavein cm3mol−1s−1 units for N2 bath gas

k57a = 6.9× 1041T−9.30 exp(−26000/T ) 0.1 atm

= 2.3× 1042T−9.11 exp(−27100/T ) 1.0 atm

= 1.7× 1038T−7.64 exp(−27000/T ) 10 atm.

Page 160: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

284 Anthony M. Dean, Joseph W. Bozzelli

(b) CH2NO + O2 −→ CH2O + NO2.

This case is analogous to allyl radical addition to O2 in that an O atom can addat the NO double bond and lead ultimately to CH2O + NO2. In the absence ofexperimental information we suggest using our QRRK expressions (Bozzelliand Dean 1993) for the allyl analog, in cm3mol−1s−1 units

k57b = 2.9× 1012T−0.31 exp(−8900/T ) 300 to 1000 K

= 1.2× 1015T−1.01 exp(−10100/T) 1000 to 2500 K.

(c) CH2NO + H −→ CH3 + NO.

Adduct formation followed by breaking the C–N bond is very exothermic,suggesting that the rate coefficient be taken as that of adduct formation,

k57c1 = 4× 1013 cm3mol−1s−1,

using the Baulch et al. (1992) value for the H + C2H5 association reaction. TheDHT method gives for the direct pathway

k57c2 = 4.8× 108 T 1.5 exp(+450/T ) cm3mol−1s−1.

(d) CH2NO + O −→ Products.

Dissociation of the adduct to formaldehyde and NO is very exothermic, sug-gesting that the rate coefficient be taken as that of adduct formation,

k57d1 = 7× 1013 cm3mol−1s−1,

the Baulch et al. (1992) value for the rate coefficient of the O + C2H5 reaction.Direct reaction forming HCNO + OH is also very exothermic; the DHT estimateis

k57d2 = 3.3× 108 T 1.5 exp(+450/T ) cm3mol−1s−1.

(e) CH2NO + OH −→ Products.

Formation of H2COH + NO from the adduct is 130 kJ/mol exothermic, sug-gesting that its rate coefficient be taken as that of adduct formation, for whichwe estimate

k57e1 = 4× 1013 cm3mol−1s−1

based on the Baulch et al. recommendation for OH + CH3. The DHT ratecoefficient estimate for direct hydrogen abstraction to form HCNO + H2O is

k57e2 = 2.4× 106 T 2 exp(+600/T ) cm3mol−1s−1.

Page 161: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 285

(f) CH2NO + CH3 −→ Products.

Formation of C2H5 + NO, which is 105 kJ/mol exothermic, is the only likelypathway for the adduct to decompose. The rate coefficient is expected to bethat for adduct formation,

k57f1 = 3× 1013 cm3mol−1s−1,

the Baulch et al. (1992) recommendation for the rate coefficient of the CH3+ C2H5 reaction. The DHT estimate for the rate coefficient for hydrogenabstraction to form HCNO + CH4 is

k57f2 = 1.6× 106 T 1.87 exp(+560/T ) cm3mol−1s−1.

(g) CH2NO + NH2 −→ Products.

Decomposition of the adduct to form CH2NH2 and NO is 115 kJ/mol exother-mic; we suggest taking the same rate coefficient as for the reaction with CH3

k57g1 = 3× 1013 cm3mol−1s−1.

The DHT estimate for the rate coefficient of the direct hydrogen transfer reactionforming HCNO + NH3 is

k57g2 = 1.8× 106 T 1.94 exp(+580/T ) cm3mol−1s−1.

2.7.33 Reactions of CH3NO

CH3NO is an important intermediate in the reaction between CH3 and NO atlower temperatures, as discussed in Section 6.15. Although rapid dissociationof this adduct is likely at higher temperatures, there is also the possibility ofradical attack, because the C–H bond dissociation energy is only 315 kJ/mol.Rate coefficients for hydrogen abstraction (reactions 58a through 58e) computedby the DHT method are listed in Table 19.

Radicals can also add to CH3NO and then eliminate CH3. We estimated thecorresponding rate coefficients as half of the corresponding rate coefficients forC2H4 (Baulch et al. 1992; Lightfoot and Pilling 1987: Tsang and Hampson 1986).The results in cm3mol−1s−1 are

CH3NO + H −→ CH3 + HNO k58f = 1.8× 1013 exp(−1400/T )+ O −→ CH3 + NO2 k58g = 1.7× 106T 2.08

+ OH −→ CH3 + HONO k58h = 2.5× 1012 exp(−500/T ).

Page 162: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

286 Anthony M. Dean, Joseph W. Bozzelli

2.7.34 Reactions of HON

HON appeared in Section 7.28 as a product of the recombination of OH withNCO. For completeness, we consider possible further reactions of this species.HON can dissociate to H + NO or isomerize to HNO. Melius (1993) estimatedthe barrier for isomerization to HNO to be about 85 kJ/mol, while the barrier fordissociation to H + NO should be only 5 to 10 kJ/mol above the endothermicity,i.e., about 60 kJ/mol. The A-factor for dissociation should also be considerablyhigher, suggesting that dissociation is the dominant path. This conclusion wasconfirmed by QRRK analysis, which suggested that the dissociation channel willbe at the low pressure limit for conditions of interest in combustion, with the ratecoefficient expression in cm3mol−1s−1 units for N2 bath gas

HON+M −→ H+NO+M, k59a = 5.1× 1019T−1.73 exp(−8075/T ) .

At 2000 K and 1 atm N2, the effective first-order rate coefficient is about 1 ×107 s−1, similar to that for HCNH (Section 7.21). Thus one should consider thesame types of reactions for HON—recombination with the flame radicals H, O,and OH, and addition to O2. The adduct formation pathways all have exothermicdissociation channels. We expect them to have rate coefficients equal to those ofadduct formation and estimate from the Baulch et al. (1992) recommendationsfor the analogous association reactions with C2H5, in cm3mol−1s−1 units,

HON+ H −→ HNO+H k59b1 = 2× 1013

+ H −→ OH+NH k59b2 = 2× 1013

+ O −→ OH+NO k59c = 7× 1013

+ OH −→ HONO+H k59d = 4× 1013.

The reaction with O2 should be more important for HON, since the adductcan easily isomerize and then dissociate. However, the well depth of HONO2 isexpected to be quite shallow, so even the relatively low barrier for isomerizationmight be above the entrance channel. The adduct has an alternative low energyexit channel in breaking the O–O bond. We expect the rate coefficient to be thatfor adduct formation,

1HON+O2 −→ O+HONO, k59e = 1.0×1012 exp(−2500/T ) cm3mol−1s−1.

2.7.35 Reactions of HCOH

HCOH can be produced in the reaction between HCNO and OH, reaction (54f). Itsmost likely reaction is isomerization to CH2O, which is more stable by about 205kJ/mol. We are not aware of any experimental or theoretical information availablefor this reaction. We estimate its rate coefficient from that of the analogous

Page 163: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 287

reaction in which CCH2 isomerizes to HCCH. In both cases there is an exothermichydrogen shift from a carbene-type carbon atom. The barrier for CCH2 is about20 kJ/mol; we estimate a more conservative value of 40 kJ/mol for HCOH. Usingthis value in a QRRK analysis gave for N2 bath gas the expressions in s−1 units

k60 = 3.5× 1017 T−2.86 exp(−4470/T ) 0.1 atm

= 2.1× 1019 T−3.07 exp(−4800/T ) 1.0 atm

= 1.8× 1021 T−3.32 exp(−5460/T ) 10 atm.

This isomerization rate is comparable to that of NNH dissociation, and it isunlikely that bimolecular reactions would be sufficiently rapid to compete. Weconsidered the reaction with O2, but the energies of the various intermediates andthe barriers for the various rearrangements were too uncertain to allow meaningfulconclusions. The results suggest that reactions with O2 may be about 20 timesslower than isomerization for typical combustion conditions.

2.7.36 Reactions of NH2OH

NH2OH can be formed in the reaction between NH2 and OH as well as in reactionsof HO2 with HNOH and NH2O. Dissociation to NH2 + OH, the reverse ofreaction (16a), discussed in Section 6.12, is sufficiently slow that NH2OH canalso undergo radical attack and hydrogen abstraction at both ends of the molecule.Rate coefficient expressions derived by the DHT method are included in Table 19for the direct hydrogen abstraction reactions with H, O, OH, CH3, NH2 and HO2.

2.7.37 Reactions of NH2NO

NH2NO is an adduct formed in the NH2 + NO reaction and a product of theH2NN + OH reaction. It dissociates mostly to N2 and H2O. (Dissociation to NH2and NO has appeared as the reverse of reaction 18c.) QRRK analysis indicatesthat most reaction proceeds via isomerization to HNNOH, the rest is direct. Forcombustion conditions the reaction is in the falloff pressure regime. The QRRKrate coefficient expressions for dissociation to N2 and H2O, derived using the datafrom Table 10, are, in cm3mol−1s−1 units,

k62a = 4.1× 1033 T−7.18 exp(−17700/T ) 0.1 atm

= 3.1× 1034 T−7.11 exp(−18260/T ) 1.0 atm

= 2.9× 1031 T−5.91 exp(−18205/T ) 10 atm.

Rate coefficient expressions derived by the DHT method are included in Table 19for the direct hydrogen abstraction reactions with H, O, OH, CH3, NH2 and HO2.

2.7.38 Reactions of H2NNHO

H2NNHO is formed in the reaction between N2H3 and HO2 (Section 7.6). AQRRK analysis gave for its most likely dissociation pathway to NH2 and HNO

Page 164: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

288 Anthony M. Dean, Joseph W. Bozzelli

the rate coefficient expressions for reaction in N2, in cm3mol−1s−1 units

k63a = 2.7× 1039 T−8.74 exp(−20945/T ) 0.1 atm

= 2.4× 1040 T−8.73 exp(−20940/T ) 1.0 atm

= 1.2× 1041 T−8.64 exp(−20920/T ) 10 atm.

The reaction is close to the low pressure limit at temperatures over 1000 K andpressures below 10 atm. A channel forming NH2NO + H is also possible, butproved to be much slower and can be neglected.

Rate coefficient expressions derived by the DHT method are included in Table 19for the direct hydrogen abstraction reactions with H, O, OH, CH3, NH2 and HO2.

2.7.39 Reactions of ClNO

Chemical bonds of chlorine are usually quite weak, the Cl–H bond in HClbeing the most important exception. For this reason, chlorine atoms may bethe radical species present at highest concentration in combustion and pyrolysissystems where chlorinated hydrocarbons or other sources of chlorine are presentin significant quantities. Such combustion processes are usually operated underfuel lean conditions, where hydrogen needed to form HCl is in limited supply.Because the O–H bonds in OH and H2O are stronger than the H–Cl bond, hydrogenis generally not found to be bonded to chlorine under these conditions. Fuel leanoperation favors high levels of Cl atoms, or at low temperatures Cl2 or other Cladducts. Thermochemical properties of the chlorine-containing species of interestin this section are shown in Table 18.

Several reactions involving chlorine have to be considered in connection withNOx chemistry, specifically the addition reaction with NO and subsequent reac-tions of the ClNO adduct. Because the adduct has a relatively weak Cl–NO bond,160 kJ/mol, NO can react with Cl much as with H and OH, as discussed previously,to provide a catalyst for combination reactions

Cl+ NO −→ ClNO

ClNO+X −→ XCl+NO

(overall) Cl+X −→ XCl

where X is an atom or radical present in the system. The rates of dissociation andaddition reactions of triatomic molecules like ClNO are in or near the low pressurelimit under combustion conditions. Even for weakly bonded molecules like ClNO,however, the low-pressure dissociation rate at typical combustion conditions isslow enough that transient concentrations become significant. Thus ClNO can bean important species both in hazardous waste disposal processes and in general-purpose industrial or municipal waste incinerators. We consider several of thereactions of ClNO with radicals expected to be in the highest concentrations.

Page 165: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

290 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.18. Thermochemical data for chlorine species.

Species1 f H ◦298

kJ/molS◦298

J mol−1 K−1C◦P(T )/J mol−1 K−1 at T/K =

300 400 600 800 1000 1500

Cl 120.9 165.3 21.8 22.3 22.6 22.4 22.2 22.0Cl2 0.0 223.0 33.9 35.1 36.6 37.3 37.6 38.1HCl –92.3 186.8 29.1 29.1 29.6 30.5 31.6 33.9ClO 101.3 226.4 31.4 33.1 35.3 36.4 36.9 37.7HOCl –74.5 236.4 37.3 40.0 43.9 46.6 48.5 51.9ClNO 51.7 261.5 44.6 47.0 50.7 53.1 54.9 57.3

All values refer to the ideal gas state.

(a) ClNO + M −→ NO + Cl + M.We suggest using the Baulch et al. (1981) recommendation, which covers thetemperature range 800 to 1500 K and is valid for Ar bath gas:

k64a = 1.3× 1015 exp(−16100/T ) cm3mol−1s−1.

(b) O + ClNO −→ ClO + NO.We suggest use of the expression of Abbatt et al. (1989) but caution that theirdata were collected at low temperatures (220–450 K).

k64b = 5× 1012 exp(−1520/T ) cm3mol−1s−1.

(c) OH + ClNO.This reaction has two exothermic pathways, one being abstraction of Cl, theother being addition of OH to the unsaturated bond in ClN=O followed bydissociation of the weaker Cl–N bond to release Cl. The addition-dissociationchannel has, as one would expect, a slower rate than the abstraction channel.We suggest adoption of the expressions, in cm3mol−1s−1 units, of Abbatt et al.(1989) for both the abstraction channel

OH+ ClNO −→ HOCl+NO, k64c1 = 5.4× 1012 exp(−1130/T )

and the addition channel

OH+ ClNO −→ HONO+ Cl, k64c2 = 5.5× 1010 exp(+240/T ) .

(d) Cl + ClNO −→ Cl2 + NO.We suggest adoption of the expression of Abbatt et al. (1989)

k64d = 4.0× 1013 exp(+128/T ) cm3mol−1s−1.

(e) H + ClNO −→ HCl + NO.The temperature dependence has been measured at low temperatures (240–460 K) by Wagner et al. (1976), who reported

k64e = 4.6× 1013 exp(−457/T ) cm3mol−1s−1.

This expression is consistent with other measurements at room temperature.(Mallard et al. 1993).

Page 166: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 291

2.8 ILLUSTRATIVE MODELING RESULTS

To illustrate the kinetics of nitrogen chemistry for representative combustionconditions, we assembled the elementary reactions discussed in the precedingsections into the full mechanism given as Table 19 at the end of this chapter. Itis not complete for combustion modeling, however, because even if one restrictsone’s interest to ammonia oxidation it is still necessary to add additional reactionsto describe the H–O chemistry that runs in parallel to the H–O–N chemistry.Exploring the combustion chemistry consequences of interactions between hydro-carbon radicals and nitrogen species requires one to include elementary reactionsdescribing all of the appropriate hydrocarbon oxidation chemistry.

To facilitate a number of comparisons with the well-tested earlier nitrogenchemistry model of Miller and Bowman (1989), we combined the Table 19reactions with the hydrogen and hydrocarbon reactions of the Miller and Bowmanmechanism to compose a full C–H–O–N mechanism that we will designate asDBN/MBC. It comprises 522 reactions of 87 species (151 reactions from Millerand Bowman, with their third-body efficiencies, and the 371 reactions in Table19). All elementary reactions were considered to be reversible, the reverse reactionrate coefficient expressions being derived from equilibrium constants computedusing the thermodynamic data listed in Tables 1 and 2. With this mechanismwe computed product distributions, reaction rates, and sensitivity spectra forrepresentative reaction conditions that we found to be suitable for exploring themain kinetic features of the combustion chemistry of nitrogen.†

We also present results computed using the complete Miller-Bowman mech-anism, i.e., using their mechanism for both the hydrocarbon and the nitrogenchemistry, in order to compare predictions made using our nitrogen chemistry tothose obtained with the mechanism most commonly used in recent years. Wecall this mechanism MB; it includes in addition to the Miller and Bowman (1989)reactions the updates described by Miller et al. (1991) and the NH3 + M −→NH2 + H + M reaction. Comparing MB profiles to DBN/MBC profiles allows usto uncover the differences that can result using alternate descriptions of nitrogenchemistry. Full discussion of the causes of these differences is beyond the scope ofthis chapter, but just showing them illustrates some important issues confrontinghow one models nitrogen chemistry in combustion today. Detailed study of thesedifferences will reveal important topics for further research in this area.

† A number of important rate coefficients have been measured under conditionswhere Ar was the dominant collision partner. In the construction of Table 19 thesewere all converted to expressions valid for N2 as the principal collision partnerusing the assumption that the ratio of Ar to N2 collision efficiency is 0.7. Othercollision efficiencies relative to N2 were assigned to be H2O = 7.0, CO2 = 2.0 andCH4 = 2.0. See Section 9.4 of Chapter 1 for further discussion of this topic.

Page 167: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

FIGURE 2.62. Overview of reactions connecting various species in the N/H/O system.The bimolecular partners for the major reactions are shown. An X designates that thepathway involves a series of hydrogen transfer reactions and sometimes unimoleculardissociation as well.

2.8.1 Ammonia oxidation

Figure 62 shows the reaction paths connecting the species of the H–O–N system.An interesting aspect of this diagram is that most species can be converted readilyeither into NO or into N2. Small shifts in conditions can affect the balance ofreactive flux in alternative directions. A significant difference between the Fig. 62description of the H–O–N system and earlier ones is the absence of a direct pathwayfrom NH2 to HNO via O2. As discussed in Section 6.9, a better description isthat this route to HNO proceeds via NH2O and HNOH. Other differences are therole of NNH in NO production, as discussed in Section 7.3, and the participationof H2NN. Figure 62 also shows the potential involvement of HONO as a routefor NO loss. The reaction pathways shown in Fig. 62 make clear that radical-radical reactions play important roles in nitrogen combustion chemistry—the largedissociation barriers of nitrogen-containing radicals, relative to their hydrocarboncounterparts, imply much longer radical lifetimes.

Kinetic model predictions for ammonia oxidation are shown in Figs. 63–68. Acomparison with the experimental data of Dean et al. (1982) at 1280 K and 1.2atm is given in Fig. 63. While there was experimental evidence that the observedinduction time was influenced by surface reactions in the quartz reactor, other

Page 168: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 293

Ammonia Oxidation at 1280 K

–7

–6

–5

–4

–3

100 200 300t / ms

log

(mol

e fr

actio

n)

NH3 observed NH3 DBN/MBCNH3 MBNH3 Hennig et al.NO observedNO DBN/MBCNO MBNO Hennig et al.

FIGURE 2.63. Comparison of predictions for NH3 decay and NO formation to the dataof Dean et al. (1982) for the oxidation of 900 ppm NH3 by 4% O2 with 1% H2O inhelium at 1279 K and 1.2 atm. The induction time was observed to be influenced by thereactor surface, suggesting heterogeneous initiation. Since this was not considered in themodel, one expects to see somewhat longer induction times in the model predictions thanwere observed in the experiment. Taking this offset in induction time into account, themechanism based on the rate coefficients recommended in this chapter (DBN/MBC) isseen to describe much of the observed kinetics. The Miller-Bowman mechanism (MB)predicts much faster decay of ammonia than observed as well as too much NO production.Predictions using the Hennig et al. (1995) rate coefficient for NH2 + O2 in the DBN/MBCmechanism are much too slow.

aspects of the kinetics appeared to be unaffected. The amount of NO producedwas much less than the amount of NH3 reacted—8 ppm of NO were produced byloss of about 500 ppm NH3. Most dynamic features are qualitatively captured bythe DBN/MBC model. Given the heterogeneous character of the initiation, theearly time mismatch for NO is to be expected. Aside from this offset, the modelcorrectly predicts the “burst” of NO at early time, and the computed NO yield is

Page 169: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

294 Anthony M. Dean, Joseph W. Bozzelli

somewhat larger than observed. The predicted rate of NH3 decay is faster thanobserved.

It is encouraging that these features are reasonably reproduced with elementaryreaction rate coefficient expressions assigned from an assessment of the exper-imental measurements and theoretical plausibility, with no adjustment to fit theDean et al. data. The burst of NO, for example,was accounted for without invokingdirect conversionof NH2 to HNO as was done in the original analysis. (Cf. Section6.9) For these conditions the MB mechanism substantially overpredicts the NH3decay rate and the NO yield. It can also be seen in Fig. 63 that the Hennig etal. (1995) set of rate coefficients leads to predicted reaction rates that are muchslower than found experimentally. The essential reason for this is that Hennig etal. considered the major channel of the NH2 + O2 reaction to be formation of NOand H2O, which depletes the radical pool too rapidly.

In the DBN/MBC reaction mechanism, the species NH2O, NHOH and H2NNplay important roles. These species have not been considered in most previouslyproposed ammonia oxidation mechanisms; indeed, HNOH and H2NN have notbeen considered before at all. Reaction flux analysis shows that most of the NH2is formed by

NH3 + OH −→ NH2 +H2O. (7)

Some of the NH2 reacts with HO2 to form NH2O,

NH2 +HO2 −→ NH2O+ OH (14b)

which isomerizes rapidlyNH2O −→ NHOH. (37b)

The NHOH product reacts mostly with NH2

HNOH+NH2 −→ H2NN+H2O, (38f2)

and the H2NN produced mostly reacts with O2 to regenerate NH2 and form NO2,

H2NN+O2 −→ NH2 +NO2 (30b)

which then react with one another in

NH2 +NO2 −→ NH2O+NO (34d2)

and then cycle through the 37b–34d2 sequence.Figure 64 shows the evolution in time of the logarithmic sensitivity spectrum

for NH3 consumption under the ammonia oxidation conditions of Fig. 63. Thetwo most sensitive reactions, with opposing effects, are seen to be the two majorchannels of the NH2 + NO reaction. Reactions involving NH2O, H2NN andHNOH are also important.

Page 170: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 295

NH3 Oxidation at 1280 K

–2

–1

0

1

100 200 300 t / ms

Nor

mal

ized

NH

3 S

ensi

tivity

NH2 + NO → NNH + OH

NH2 + NO → N2 + H2O

O + OH → O2 + H

NH2 + O2 → NH2O + O

HNOH + NH2 → H2NN + H2O

NH2 + HO2 → NH2O + OH

NH + O2 → NO + OH

FIGURE 2.64. Normalized sensitivity coefficients computed by the SENKIN program(Lutz et al. 1991) for the reactions with the largest effects on ammonia concentrationsduring ammonia oxidation for the conditions of Fig. 63. The clearly important reactionsinvolving H2NN, HNOH and NH2O have not been included in previous ammonia oxidationmechanisms.

The rate of NH3 decay is sensitive to several rate coefficients for which there islittle experimental data. In particular, the rate coefficients of

NH2 +O2 −→ NH2O+O (13b)

HNOH+NH2 −→ H2NN+H2O (38f2)

NH2 +HO2 −→ NH2O+OH, (14b)

all of which are important for NH3 decay (Fig. 64). The fit to the NH3 decayprofile (Fig. 63) could have been improved by adjusting one or more of theseexpressions, but experimental and theoretical investigations are to be preferred.The fact that the computed NH3 profile is close to the observed one suggests thatthe mechanism and estimated rate coefficient expressions are reasonable, but even

Page 171: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

296 Anthony M. Dean, Joseph W. Bozzelli

Ammonia Oxidation at 2000 K

–5

–4

–3

300 600 900

t / µs

log

(mol

e fr

actio

n)NH3 DBN/MBCNH3 MBNO DBN/MBCNO MBN2O DBN/MBCN2O MBN2 DBN/MBCN2 MB

FIGURE 2.65. Comparison of predictions using a mechanism comprising the reactionsand rate coefficient expressions recommended in this chapter (DBN/MBC) to those of theMiller-Bowman mechanism (MB) for oxidation of 900 ppm NH3 by 4% O2 at 2000 Kand 1 atm. The rates and selectivities to major products computed by the two mechanismsare essentially the same. Ammonia is almost completely converted to NO at this hightemperature.

that suggestion needs to be tested by further studies of the relevant elementaryreactions themselves.

The effects of temperature can be seen in Figure 65, which shows computedconcentration profiles for 900 ppm NH3 and 4% O2 in Ar reacting at 2000 K. Atthis temperature the two models are in better agreement with one another. TheNH3 decay rate and the major product yields are similar for the two mechanisms.The DBN/MBC model predicts slightly more NO production, with less N2 andN2O. Unlike at lower temperatures, NH3 is almost completely converted into NO.

To explore the reasons for the shift in selectivity toward NO at high temperatures

Page 172: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 297

–15

–10

–5

0

5

10

100 200 t /ms

rate

/ nm

ol c

m –3

s–1

total rate

NH2 + NO → N2 + H2O

HNO + O2 → NO + HO2 NO2 + NH2 → NH2O + NO

NH2 + NO → NNH + OH

NH + O2→ NO + OH

HO2 + NO → NO2 + OH

.

NO production at 1280 K

FIGURE 2.66. Major routes of NO production and decay in the DBN/MBC model for theconditions of Fig. 63. The rapid burst of NO predicted near 100 ms is due to HNO + O2−→ NO + HO2, with NO2 + NH2−→ NH2O + NO arriving slightly later. The rate thendecays to essentially zero as NO is consumed, mainly by reaction with NH2.

we have to identify the major reactions affecting NO production at the twotemperatures. The DBN/MBC reaction rates at 1280 and 2000 K shown inFigs. 66 and 67 suggest that the NO burst at low temperature is due mostly to

HNO+O2 −→ NO+HO2 (36h)

and NO2 +NH2 −→ NH2O+NO (34d2)

After this burst NO decays mainly through reaction with NH2, with the result thatthere is virtually no further increase of NO. In contrast, at 2000 K (Figure 67)none of the important reactions consume NO. The major source of NO is thermaldissociation of HNO, which is much faster at the higher temperature. The absenceof reactions that consume NO at 2000 K can be attributed to the increased NH2concentration being offset by the decrease in the overall rate coefficient for NH2 +

Page 173: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

298 Anthony M. Dean, Joseph W. Bozzelli

0

10

20

30

300 600 900 t / µs

mol

cm

µra

te /

–3 s

–1

total rate

HNO + M → H + NO + M

NH + O → NO + H

HNO + O → OH + NO

HNO + O2 → NO + HO2

HNO + OH → NO + H2O

H + N2O → NH + NO

.

NO production at 2000 K

FIGURE 2.67. Major routes of NO production in the DBN/MBC model for the hightemperature conditions of Fig. 65. The higher temperature causes HNO + M−→ H + NO+ M to dominate NO production. In contrast to the lower temperature regime, reduction ofNO by reaction with NH2 is unimportant, and one sees almost quantitative conversion ofammonia to NO.

NO so that this rate is comparable to that at 1280 K, about 20 nmolcm−3s−1, whilethe NO production rate at 2000 K peaks near 30 µmolcm−3s−1, overwhelmingall NO consumption rates. (Cf. Figures 34 and 66)

HNO is the major precursor of NO at both temperatures. At 1280 K, 5 reactionscontribute to producing HNO. (Figure 68) The long-lived radicals NH2O andHNOH (which are also involved in the alternate path to NO going through thereaction sequence 37b–34d2) are prominent in this chemistry, primarily throughlosing an H-atom in reactions with other radicals. Some HNO is also produced by

NH+O2 −→ HNO+O (12g)

NH2 + O −→ HNO+H (15c)

At 2000 K reaction 15c accounts for most of the HNO production, essentiallybecause at this temperature the concentration of O atoms is much greater.

Page 174: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 299

–10

–5

0

5

100 200 t / ms

mol

cm

nra

te /

–3 s

–1

total rate

HNO + O2 → NO + HO2 NH + O2 → HNO + O

NH2O + NH2 → HNO + NH3

NH2 + O → HNO + H

NH2O + OH → HNO + H2O

HNOH + NH2 → HNO + NH3

.

HNO production at 1280 K

FIGURE 2.68. Rate analysis for the most important reactions contributing to HNOproduction for the conditions of Fig. 63. Most of them involve hydrogen transfer betweenradicals.

2.8.2 Kinetics of selective noncatalytic reduction of NO

The elementary reactions involved in selective noncatalytic reduction of NO(SNCR), whereby NH2 sources are injected into the post-combustion zone, havealready been discussed in various sections of this chapter. The earliest SNCRrealization, the Thermal DeNOx process invented by Lyon in 1975, uses NH3 asthe NH2 source; it has been installed in over a hundred commercial furnaces andboilers. (Lyon and Hardy 1986) The kinetics common to all SNCR processesare interesting in that each has a temperature window—for NH3 addition between1100 and 1400 K—where some of the NO is reduced to N2. At lower temperaturesthe kinetics are too slow and at higher temperatures more NO is produced thanconsumed. The representative laboratory data (Lyon 1987) shown in Fig. 69

Page 175: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

300 Anthony M. Dean, Joseph W. Bozzelli

Thermal DeNOx

100

200

1100 1200 1300 1400 1500

T / K

NO

mol

e fr

actio

n / p

pm

experiment (Lyon 1987)

MB

DBN/MBC

FIGURE 2.69. Comparison of predictions for NO removal at various temperatures tothe experimental data of Lyon for typical Thermal DeNOx conditions: 380 ppm NH3,230 ppm NO, 4% O2, 10% H2O in helium at 1.1 atm with a residence time of 0.2 s.The Miller-Bowman mechanism (MB) is in good agreement with the data. Although theDBN/MBC mechanism requires higher temperatures for NO reduction than observed, theoverall agreement is especially satisfying in that no rate coefficient adjustments were madeto attain it.

illustrate the effectiveness of SNCR: At the center of the temperature window,near 1250 K, addition of 380 ppm NH3 reduces a starting NO concentration of230 ppm to about 35 ppm.

Many models describing SNCR chemistry have been advanced and testedagainst the available experimental data. Glarborg et al. (1994) summarize theearlier studies and discuss the remaining problems, in particular two troublesomeaspects of the MB mechanism. The first is their use of a larger branching ratiofor OH production in the NH2 + NO reaction. (cf. Section 6.14). Glarborg etal. reduced the branching ratio from the 0.51 MB value to values near 0.3 for theSNCR temperature range. The second one is MB’s use of an NNH lifetime much

Page 176: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 301

longer than expected from theory. (Cf. Section 7.3). Glarborg et al. reducedthe assumed dissociation lifetime of NNH from the Miller and Bowman value of100 µs to 1 µs and found that the combination of these two changes significantlydiminished the agreement between model and experiment.

Figure 69 shows model predictions for typical SNCR conditions. The MBmechanism captures much of the detail of the Lyon and Hardy data. Use of theDBN/MBC mechanism leads to slower chemistry, with the result that the SNCRwindow is offset to higher temperatures. Further calculations showed that a slightincrease in the branching ratio of the NH2 + NO reaction (Section 6.14), from 0.28to 0.32 at 1200 K,1 suffices to shift the minimum of the DBN/MBC temperaturewindow from 1300 to 1250 K. Thus the BDN/MBC model, despite the imperfectmatch with the data, captures the main features of the SNCR process. This isespecially encouraging in that the model correctly addresses both of the problemsnoted by Glarborg et al. (1994), i.e., the branching ratio of the NH2 + NO reactionand the lifetime of NNH.

Figure 70 illustrates the effect of H2O on SNCR kinetics. In both the experi-ments and the DBN/MBC model the SNCR window of the dry system shifts tolower temperature by about 25 to 50 K.

The overall agreement with experiment found for SNCR conditions using theDBN/MBC mechanism is encouraging. We feel that because of the complexityof the model it would not be appropriate to adjust the rate coefficient expressionsof reactions 18a and 18b to achieve agreement with the experimental SNCR data.Further experiments offer a better alternative to deepen our understanding ofnitrogen chemistry under these conditions.

2.8.3 Fuel-rich ammonia flames

Dean et al. (1984) concluded that it was necessary to include a number ofreactions of various NHi species that form N–N bonds in order to satisfactorilyaccount for species profiles that they had measured in fuel-rich ammonia flames.We have compared the predictions of the DBN/MBC mechanism to the profilesmeasured in that work as a further test of our rate coefficient assignments. Suchcomparisons are stringent tests of the mechanism in that concentrations of thereactive intermediates NH, NH2, and OH were measured in addition to NH3 andNO in the flame experiments.

1 For modified branching ratio calculations we adjusted the expressions listed inTable 19 for the two NH2 + NO reactions, k18a and k18b, so as to keep their sumconstant but their ratio greater by 0.04 over the temperature range of the SNCRwindow, i.e., from 1000 to 1300 K.

Votsmeier et al. (1998) reported a direct measurement of the branching ratiobetween 1350 and 1750 K. Extrapolation of the formula they derive to lowertemperature gives a branching ratio of 0.35 at 1250 K, similar to the value suggestedby the above modeling exercise.

Page 177: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

302 Anthony M. Dean, Joseph W. Bozzelli

Thermal DeNOx

100

200

1200 1300

T/K

NO

mol

e fr

actio

n / p

pm

experiment (Lyon 1987)

DBN/MBC

experiment (dry) (Lyon 1987)

DBN/MBC (dry)

FIGURE 2.70. Effect of H2O on NO reduction in the Thermal DeNOx process. Conditionsas in Fig. 69 except that no H2O was added for the dry case. The predicted shift is similarto the one observed.

Figures 71 and 72 compare the predictions of the DBN/MBC and MB mecha-nisms to the observed profiles of NO and NH for an atmospheric pressure NH3–O2–N2 flame with an equivalence ratio of 1.81. The observed peak flame temperaturewas 1875 K at a point 1.6 mm above the burner surface. Much of the NOinitially produced in the flame front region is destroyed in the hot post-flameenvironment. The peak NH concentration is lower than that for NO, and its decayis more gradual. For both of these species, the DBN/MBC mechanism provides abetter description than the MB mechanism. The extent of the agreement with theobserved profiles is good, especially considering that no normalization was usedand that no adjustments were made to the DBN/MBC mechanism.

Profiles were also computed using the Hennig et al. (1995) values for NH2 +O2 rate coefficients (which include NO + H2O as the major product channel, cf.Section 6.9) in the DBN/MBC mechanism. The results are similar to the effectseen above for lower temperature ammonia oxidation (cf. Fig. 63). Inclusion ofthis pathway not only overpredicts the NO concentration, but the chain termination

Page 178: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 303

Phi = 1.81 ammonia flame

–4

–3

–2

2 4 6z / mm

log

( N

O m

ole

frac

tion

)

observed

DBN/MBC

MB

Hennig et al. (1995)

FIGURE 2.71. Comparison of observed and predicted NO profiles in an ammonia flamewith an equivalence ratio of 1.81 (Dean et al. 1984). The predictions using the mechanismdeveloped in this work (DBN/MBC) are closer to the experimental results than thoseobtained using the Miller/Bowman mechanism (MB). Use of the rate coefficients of Henniget al. (1995) for NH2 + O2 with the DBN/MBC mechanism substantially overpredict theamount of NO formed and move the peak to larger distances from the burner surface.

nature of this reaction makes the chemistry much slower, such that the predictedpeak NO concentration is too far above the burner surface. The NH predictionsalso show the peak much too high above the burner surface. In this case, thepredicted peak concentration is lower than that observed.

Figures 73 and 74 compare the rates of NO production as a function of heightabove burner for the two mechanisms. The peak production rate is slightlylower with the DBN/MBC mechanism (Fig. 73) than with the MB mechanism(Fig. 74). In both, the dominant pathway for NO production is HNO dissociation,just as was found earlier for lean ammonia oxidation at 2000 K. However, therethe similarities with lean ammonia oxidation cease; in the rich flame there aresignificant pathways for NO destruction, with the result that in the fuel-rich flamelittle of the NH3 is converted to NO, whereas NH3 is almost completely convertedto NO under fuel-lean conditions.

Page 179: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

304 Anthony M. Dean, Joseph W. Bozzelli

Phi = 1.81 ammonia flame

–5

–4

2 4 6z / mm

log

( m

ole

frac

tion

NH

)

observed

DBN/MBC

MB

Hennig et al. (1995)

FIGURE 2.72. Comparison of observed and predicted NH profiles in an ammonia flamewith an equivalence ratio of 1.81 (Dean et al. 1984). The predictions using the mechanismdeveloped in this work (DBN/MBC) are somewhat closer to the experimental results thanthose obtained using the Miller/Bowman mechanism (MB). Use of the rate coefficientsof Hennig et al. (1995) for NH2 + O2 with the DBN/MBC mechanism underpredict theamount of NH formed and move the peak to much larger distances from the burner surface.

Comparing the predictions of the MB and DBN/MBC models, we see that themajor sink for NO is different. In the MB mechanism the dominant channel is

NNH+NO −→ N2+HNO, (28h)

while in the DBN/MBC mechanism it is

NH2 +NO −→ N2+H2O. (18a)

Much of this difference can be attributed to the use of a much lower rate coefficientfor reaction 28h in DBN/MBC. There are other substantial differences; forexample, only 3 of the fastest 6 reaction rates in the DBN/MBC model are amongthe fastest 6 in the MB model.

Page 180: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 305

–200

0

200

0.5 1.0 z / mm

mol

cm

µra

te /

–3 s

–1

total rate

HNO + M → H + NO + M

NH2 + NO → N2 + H2O

HNO + H → H2 + NO

H + N2O → NH + NO

NH2 + NO → NNH + OH

HNO + O2 → NO + HO2

.

NO production at phi = 1.81 (DBN/MBC)

FIGURE 2.73. Rate analysis based upon the DBN/MBC mechanism for the most importantreactions contributing to NO production and/or decay in an ammonia flame with anequivalence ratio of 1.81. Most of the NO is produced by reactions of HNO, whilethe most important sinks involve reactions with NH2 and NH.

Figures 75 and 76 compare the rates of NH production. Again there aresignificant differences. The dominant source of NH in the DBN/MBC model is

2 NH2 −→ NH3 +NH, (17e)

while this reaction is not included in the MB model. In both models the mainreaction consuming NH is

NH2 +NH −→ N2H2 +H, (27b1)

which is important in accounting for the low conversion of NH3 to NO in thisrich flame—it consumes NHi species while forming an N-N bond. Thus, theNHi species are no longer available to be oxidized to NO and much of the N2H2goes on to form N2. A second loss route for NH in the DBN/MBC mechanism

2 NH −→ N2+ 2 H (27a)

also converts NHi to N2.

Page 181: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

306 Anthony M. Dean, Joseph W. Bozzelli

–300

0

300

1 2z / mm

mol

cm

µra

te /

–3 s

–1

total rate

HNO+M → H+NO+M

NNH+NO → N2 +HNO

HNO+NH2 → NH3 +NO

NH2 +NO → NNH+OH

NH2 +NO → N2 +H2O

N+O2 → NO+O

.

NO production at phi = 1.81 (MB)

FIGURE 2.74. Rate analysis based upon the MB mechanism for the most importantreactions contributing to NO production and/or decay in the ammonia flame with anequivalence ratio of 1.81. The production of NO from HNO dissociation is approximatelytwice as large as for the DBN/MBC mechanism, and the major sink with the MB mechanismis NNH + NO, which is much less important in the DBN/MBC mechanism.

2.8.4 Implications of the O + NNH reaction

We presented an overview of NOx production mechanisms in Section 2. Inthe absence of hydrocarbons, NO production is governed by the concentrationof oxygen atoms, as shown in Fig. 62. In the course of our investigation of thecombustion chemistry of nitrogen we identified a new pathway for NOx formation,the reaction

NNH+O −→ NH+ NO, (−11g)

which leads directly to one molecule of NO and indirectly, through subsequentreactions of NH, to still more. As discussed in Section 7.3, NNH can rapidlybecome equilibrated with H and N2 in flames, such that bimolecular reactions of

Page 182: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 307

–1

0

1

1 2z / mm

mol

cm

mra

te /

–3 s

–1

total rate

2 NH2 → NH3 + NH

NH2 + NH → N2H2 + H

NH2 + H → NH + H2

2 NH → N2 + 2 H

NH2 + OH → NH + H2O

NH + O2 → HNO + O

.

NH production at phi = 1.81 (DBN/MBC)

FIGURE 2.75. Rate analysis based upon the DBN/MBC mechanism for the most importantreactions contributing to NH production and/or decay in a φ = 1.81 ammonia flame. Mostof the NH is produced and consumed by reactions of NH2.

NNH become important. This equilibrium has consequences for NOx chemistry,for example in hydrogen/air flames. Figure 77 compares the predictions of theDBN/MBC and MB models for a stoichiometric mixture of hydrogen and air at1800 and 2400 K. At 1800 K the DBN/MBC model predicts much larger amountsof NO than the MB model, while at 2400 K, where most of the NO arises fromthe Zeldovich mechanism, the differences between the two models are small.

Harrington et al. (1995, 1996) measured NO production rates in low pressure,low temperature (near 1200 K) hydrogen–air flames. For these conditions, NOother NO production are computed to be negligible. The amount of NO observedproved to be consistent with that predicted to be formed via the O + NNH channel,which the authors felt to be the only explanation for NO levels they measured.

Figure 78 shows the major pathways for NO production and decay computedusing the DBN/MBC model for H2 burning in air at 1800 K. The two mostimportant reactions for NO production are

Page 183: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

308 Anthony M. Dean, Joseph W. Bozzelli

–1

0

1

1 2z / mm

mol

cm

mra

te /

–3 s

–1

total rate

NH2 + H → NH + H2

NH2 + NH → N2H2 + H

NH + H → N + H2

NH2 + OH → NH + H2O

NH + O2 → HNO + O

NH + NO → N2O + H

.

NH production at phi = 1.81 (MB)

FIGURE 2.76. Rate analysis based upon the MB mechanism for the most importantreactions contributing to NH production and/or decay in the ammonia flame with anequivalence ratio of 1.81. In contrast to the DBN/MBC mechanism, the major sourceof NH is H + NH2, since the reaction NH2 + NH2−→ NH3 + NH is not considered in thismechanism. The major sink is the same as in the DBN/MBC mechanism.

NNH+ O −→ NH+NO and (−11g)

HNO+ H −→ H2 +NO . (36d1)

This clarifies the difference between the DBN/MBC and MB predictions; the MBmodel does not include reaction 11g, and its rate coefficient for reaction 36d1 isover 10 times slower than the DBN/MBC expression (Section 7.11).

As discussed in Section 7.3, the flame front concentration of NNH establishedby equilibration with H + N2 is high enough that NNH reacts there with the high

Page 184: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 309

NO Production in H2/air

–7

–6

–5

–4

–3

100 200 t / µs

log

( N

O m

ole

frac

tion

)

DBN/MBC 2400 K

MB 2400 K

DBN/MBC 1800 K

MB 1800 K

FIGURE 2.77. Predicted effect of temperature on NO production in a stoichiometrichydrogen/air mixture using the DBN/MBC and MB models. Results are shown for a plugflow model at constant temperatures of 1800 K and 2400 K and 1 atm.

(over 1% mole fraction) flame front concentration of atomic oxygen. The NNH−→ N2 + H reaction is effectively in equilibrium at the flame front even if alower value for NNH dissociation rate coefficient is assumed, i.e., if the tunnelingcontribution (cf. Section 7.3) is omitted. However, use of the Miller-Bowmanvalue of 104s−1, five orders of magnitude smaller, does completely suppress thisNO production route. With the MB rate coefficient, the (reverse) reaction betweenH and N2 is not rapid enough and there is insufficient NNH to react with O-atoms.

The theoretical evidence (Section 7.3) supports a dissociation rate coefficient forNNH much larger than the value assumed by Miller and Bowman, and thereforea much larger formation rate coefficient. One has to consider in this context thereliability of the equilibrium constant for this reaction. Our calculations used theNNH thermochemistry of Melius (1993), which is consistent with that impliedby the potential energy surface of Koizumi et al. (1991). The uncertainty in theentropy of NNH can be estimated to be about 8 kJ mol−1K−1, and the heat offormation uncertainty to be about 12 kJ/mol. A worst case combination of theseerrors would shift the equilibrium constant by about a factor of 6 at 2000 K; amore realistic uncertainty estimate would be a factor of 3. Such a change in

Page 185: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

310 Anthony M. Dean, Joseph W. Bozzelli

0

10

20

50 100 150

t / µs

mol

cm

µra

te /

–3 s

–1

total rate

NH + NO → NNH + O

HNO + M → H + NO + M

HNO + H → H2 + NO

N + OH → NO + H

NH + O → NO + H

HNO + O→ OH + NO.

NO production at 1800 K in H2/air

FIGURE 2.78. Rate analysis based on the DBN/MBC mechanism for NO production anddecay in a stoichiometric H2-air mixture at 1800 K and 1 atm. Most of the NO is formedby reactions of NNH, HNO and NH. The major sink is conversion back to HNO.

the equilibrium constant (and thus the concentration of NNH) translates into acomparable shift in the predicted rate of NO production and would also affectthe time scale required for the reaction to become partially equilibrated. Sincean order of magnitude increase in the time required to attain equilibrium did notappreciably affect the results, the main thermochemical influence would be toaffect the rate of NO production; the NNH −→ N2 + H reaction will still bepartially equilibrated in flame fronts.

Comparison of the NNH production rate in Fig. 79 to that for NO productionin Fig. 78 shows that they are comparable to one another even though theconcentration of NNH is orders of magnitude less than the concentration of NO.NNH is produced by both the collisional activation and the tunneling pathways.Even though the rate of NNH + O −→ NH + NO is smaller than that for NNH +H−→ N2+ H2, rapid resupply of NNH by the H + N2 reaction maintains a near-equilibrium NNH concentration, and NO production continues to be significantuntil the concentrations of H and O begin to decay.

Page 186: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 311

–10

0

10

20

50 100 150 t / µs

mol

cm

µra

te /

–3 s

–1

total rate

NNH + M → N2 + H + M

NNH + H → N2 + H2 NH + NO → NNH + O

NNH → N2 + H

NNH + O → N2 + OH

H + N2O → NNH + O

.

NNH production at 1800 K in H2/air

FIGURE 2.79. Rate analysis based on the DBN/MBC mechanism for NNH productionand decay in a stoichiometric H2–air mixture at 1800 K and 1 atm. NNH is produced byH addition to N2 and mainly consumed in reactions with H and O. The reaction with Hrecycles NNH to N2.

2.8.5 Nitrogen chemistry in hydrocarbon–air flames

Nitrogen chemistry is much more complex in the presence of hydrocarbons,specifically hydrocarbon radicals, as illustrated in Fig. 80. Nitrogen-containingspecies can react with CHi species to form a wide variety of compounds that mayultimately cascade back to nitrogen species, not necessarily those from whencethey came. NO can be formed from N2, via attack by CH (Prompt NO), orconverted to N2 by hydrocarbon radical attack, the essence of NOx reburningstrategies. Reliable prediction of the kinetics of these processes requires anaccurate description of both the hydrocarbonand nitrogen chemistry. For purposesof this review, we can illustrate some of the interactions that occur in flames usinga plug flow model. This neglect of diffusion does not introduce major errors atpressures of 1 atm or higher, and the calculation is appreciably simpler.

Figure 81 illustrates the effect of temperature on NO production during thecombustion of a stoichiometric mixture of methane and air as predicted bythe DBN/MBC and MB models. The results are for a pressure of 1 atm andtemperature held fixed at either 1800 or 2400 K. As with the H2/air case considered

Page 187: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

312 Anthony M. Dean, Joseph W. Bozzelli

NH2

N

N2O

N2

NO

HONO

HNO

NO2

CH3NH2

CH3NH

CH2NH2

H2C=NH

HCNN

CH3NO

CH2NO

HCNH

H2CN

HCN

CN

HNCO

NCO

HCNO

HNC

HOCN

NCCN

NH2

NH

N

N2

NO NH

FIGURE 2.80. Overview of the reactions connecting various species in the C/N/H/Osystem. The N-containing species in the left column can react with the hydrocarbonradicals CH, CH2 and CH3 to produce the various CHx NyOz species that undergo extensiverearrangements while cascading back to the N-containing species in the right column. Noattempt was made to denote the co-reactants for the transformations. The N/H/O submodelis outlined in Fig. 62.

in the previous section, the results of the two models are close at 2400 K,while at 1800 K the DBN/MBC model predicts an order of magnitude higherfinal concentration of NO than the MB model. The DBN/MBC results indicatethat the increase in temperature from 1800 to 2400 K results in an order ofmagnitude increase in NO formed. The NO profile increases much more steeplyat 2400 K, following the more rapid increase in oxygen atom concentration at thistemperature. The difference in the NO profiles is greater than the difference in theO-atom profiles because the rate coefficient for NO production (Reaction 1) has ahigh activation energy.

The reasons for the difference in NO production at 1800 K can be seen in theanalysis of the NO production rates in the two models, as seen in Figs. 82 and 83.

Page 188: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 313

–7

–6

–5

–4

–3

100 200 300

t / µs

log

(NO

mol

e fr

actio

n)

DBN/MBC 2400 K

MB 2400 K

DBN/MBC 1800 K

MB 1800 K

NO production from CH4/air

FIGURE 2.81. Predicted effect of temperature on NO production in a stoichiometricmethane/air mixture using the DBN/MBC and MB models. Results are shown for a plugflow model at 1 atm and constant temperatures of 1800 K and 2400 K.

The major contributors to NO production in the DBN/MBC model are

O+NNH −→ NH+NO (−11g)

NH+ O −→ NO+H (27e1)

HNO+ H −→ H2+ NO (36d1)

In the DBN/MBC model, most NO production at 1800 K is non-Zeldovich NO.In contrast, the MB model attributes most NO production to the Zeldovich pair

N+ OH −→ NO+H (26c)

N+O2 −→ NO+O (26a)

At 2400 K essentially the same amount of NO is predicted for stoichiometricmethane and hydrogen flames (cf. Figs. 77 and 81), while at 1800 K the NOconcentration predicted for a hydrogen flame is about twice that predicted for amethane flame. Comparison of Figs. 78 and 82 reveals that the reactions producingmost of the NO are the same in both flames: The early sources of NO are reactions−11g and 27e1, while reaction 36d1 generates NO at later times. The rates of allthree reactions are lower in the methane flame.

Page 189: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

314 Anthony M. Dean, Joseph W. Bozzelli

0

2

4

100 200 t / µs

mol

cm

µra

te /

–3 s

–1

total rate

NH + NO → NNH + O

HNO + M → H + NO + M

NH + O → NO + H

HNO + H → H2 + NO

N + OH → NO + H

HNO + O → OH + NO

.

NO production at 1800 K in CH4/air

FIGURE 2.82. Rate analysis based on the DBN/MBC mechanism for NO production anddecay in a stoichiometric CH4–air mixture at 1800 K and 1 atm. The major sources of NOare NNH, HNO and NH as in the hydrogen/air system shown in Fig. 78. HNO is the majorsink, as in the hydrogen system. The peak rate of NO production, however, is about fivetimes lower for methane than for hydrogen.

2.8.6 General conclusions from modeling tests

The modeling results described in the foregoing sections show that the Table 19mechanism captures the essential features of several high-temperature nitrogenchemistry systems. The agreement found in the instances where predictions canbe compared to data is especially gratifying in that the rate coefficient assignmentswere not adjusted to fit data taken in such complex chemical situations. Animportant aspect of these tests is the emergence of hitherto unfamiliar species andreaction pathways that appear to play central roles in high-temperature nitrogenchemistry. A number of such systems were shown to warrant further study.

A generalization emerging from the results of these modeling tests is thatradical–radical reactions are important, essentially because nitrogen radicals usu-ally have slower unimolecular decay rates than their hydrocarbon analogs. Thisis traceable to the underlying thermochemistry. With hydrocarbon radicals, beta-scission reactions usually have barriers ranging from 120 to 170 kJ/mol. Mostof the energy needed to break a C–C or C–H single bond is compensated byenergy gained in formation of a carbon–carbon or carbon–oxygen double bond.

Page 190: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 315

0

300

600

100 200 300 400t / µs

mol

cm

nra

te /

–3 s

–1

total rate N+OH → NO+H N+O2 → NO+O HNO+M → H+NO+M HNO+OH → NO+H2O NH+NO → N2O + H N + NO → N2 + O

.

NO production at 1800 K in CH4/air (MB)

FIGURE 2.83. Rate analysis based on the MB mechanism for NO production and decay ina stoichiometric methane–air mixture at 1800 K and 1 atm. N-atoms are the major sourceof NO and HNO the major sink. The peak rate of NO production is about seven timeslower than predicted using the DBN/MBC mechanism (Fig. 82).

With nitrogen radicals, the energy compensation by double bond formation issubstantially lower, leading to higher barriers for beta scission. Further refinementof rate coefficient estimation procedures, e.g., our DHT procedure for radical–radical hydrogen transfer rate coefficients, could lead to improved understandingof the effects of these radical-radical reactions.

The literature contains many flame modeling studies which suggest that im-portant aspects of hydrocarbon-nitrogen interactions remain to be discovered.Examples of such studies include those by Etzkorn et al. (1992) and Williams etal. (1994) of NO-doped methane flames,of Garo et al. (1992) on production of NOand HCN in flames, and the analysis of reburn chemistry by Staph and Leuckel(1996). Reinvestigation of the relevant reaction pathways using the Table 19mechanism, together with an appropriate hydrocarbon oxidation mechanism, forthe conditions studied by these authors would be a useful undertaking.

Page 191: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

316 Anthony M. Dean, Joseph W. Bozzelli

2.9 SUMMARY

Table 19 summarizes our recommended rate coefficient expressions for the el-ementary reactions considered in this chapter. They are based upon measure-ments as far as possible, and consistency with thermochemical and theoreticalexpectations has been enforced. For important elementary reactions for whichexperimental data are not yet available, QRRK analysis was used to derive interimrate coefficient expressions and to compute the effects of temperature and pressureon rate coefficients that are expected to be pressure-dependent.

An important caveat about our rate coefficient recommendations is that usingthem as part of a dynamic model will not resolve all basic ambiguities. There arequestions in C–H–O–N chemistry for which experimental answers are still neededbefore rate coefficients and/or product channels can be assigned.

We urge acceptance of a modeling philosophy in which one makes maximumuse of the best available expressions for the elementary reaction rate coefficients,preferably obtained from studies where the reaction of interest is well isolated, asthe basis set for the model. For reactions for which such information is unavailable,we suggest adoption of techniques like those used in this chapter to obtain plausibletheoretical estimates. If using such models leads to predictions significantlydifferent from the observations, the most likely explanation is that importantelementary reactions have been overlooked. Sensitivity analysis will suggestwhere to look for missing chemistry, where to adjust estimates of reaction barriers,and so on. It is generally more productive, when faced with such circumstances,to improve one’s understanding of the chemistry than to adjust rate coefficientexpressions to improve a fit to data. Such adjustments may only serve to obscurea significant scientific problem.

We hope that our analysis has clarified which C–H–O–N reactions have ratecoefficient expressions based on reliable measurements and shown how plausibleestimates can be made for the rate coefficients of reactions lacking such measure-ments. We hope that combustion researchers will use our material extensively foranalysis of experimental data on the combustion chemistry of nitrogen. Whereour mechanism does not satisfactorily rationalize new data, we hope that it willhelp reveal the C–H–O–N combustion chemistry that remains to be discovered.

2.10 ACKNOWLEDGMENTS

We are grateful to C.F. Melius for providing results of BAC-MP4 calculations ofstandard enthalpies of formation for many species and transition state energies formany reactions. We also thank the many researchers who gave us pre-publicationcopies of their manuscripts.

Page 192: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 317

2.11 REFERENCES

Abbatt, J.P.D., Toohey, B.W., Fenter, F.F., Stevens, P.S., Brune, W.H. & Anderson,J.G. (1989). J. Phys. Chem., 93, 1022.

Adamson, J.D., Farhat, S.K., Morter, C.L., Glass, G.P., Curl, R.F. & Phillips, L.F.(1994). J. Phys. Chem., 98, 5665–5669.

Allara, D.L., & Shaw, R. (1980). J. Chem. Phys. Ref. Data, 9, 523–559.

Allen, M.T., Yetter, R.A., & Dryer, F.L. (1994). The Decomposition of NitrousOxide at Elevated Pressures, Spring Western States Section Meeting of theCombustion Institute, Paper No. 94–012.

Allen, M.T., Yetter, R.A. & Dryer, F.L. (1995). Int. J. Chem. Kinet., 27, 883.

Anastasi, C., & Maw, P.R. (1982). J. Chem. Soc. Faraday Trans. I, 78, 2423.

Atakan, B., Jacobs, A., Wahl, M., Weller, R., & Wolfrum, J. (1989). Chem. Phys.Lett., 155, 609–615.

Atakan, B., & Wolfrum, J. (1991). Chem. Phys. Lett., 186, 547–552.

Atakan, B., Kocis, D., Wolfrum, J., & Nelson, P. (1992). 24th Symposium(International) on Combustion, pp. 691–699.

Atkinson, R. (1989). Kinetics and Mechanisms of Gas-Phase Reactions of theHydroxyl Radical with Organic Compounds., J. Phys. Chem. Ref. DataMonograph No. 1.

Atkinson, R., Baulch, D.L., Cox, R.A., Hampson, R.F., Jr., Kerr, J.A., & Troe, J.(1989). J. Phys. Chem. Ref. Data, 18, 881–1097.

Atkinson, R., Baulch, D.L., Cox, R.A., Hampson, R.F., Jr., Kerr, J.A. & Troe, J.(1992). J. Phys. Chem. Ref. Data, 21, 1125–1568.

Back, R.A. & Yokota, T. (1973). Int. J. Chem. Kinet., 5, 1039.

Baldwin, A.C., & Golden, D.M. (1978). Chem. Phys. Lett., 55, 350–352.

Balla, R.J., Casleton, K.H., Adams, J.S., & Pasternack, L. (1991). J. Phys.Chem., 95, 8694–8701.

Bauerle, S., Klatt, M. & Wagner, H.Gg. (1995). Ber. Bunsenges. Phys. Chem.,99, 97–104.

Baulch, D.L., Cobos, C.J., Cox, R. A., Esser, C., Frank, P., Just, T., Kerr, J.A.,Pilling, M.J., Troe, J., Walker, R.W., & Warnatz, J. (1992). J. Phys. Chem.Ref. Data, 21, 411–734.

Baulch, D.L., Cobos, C.J., Cox, R.A., Frank, P., Hayman, G., Just, T., Kerr,J.A., Murrells, T., Pilling, M.J., Troe, J., Walker, R.W. & Warnatz, J. (1994).Combust. Flame, 98, 59–79.

Page 193: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

318 Anthony M. Dean, Joseph W. Bozzelli

Baulch, D.L., Duxbury, J., Grant, S.J. & Montague, D.C. (1981) J. Phys. Chem.Ref. Data, 10, Suppl. 1

Becker, K.H., Engelhardt, B., Geiger, H., Kurtenbach, R., Schrey, G., & Wiesen,P. (1992a). Chem. Phys. Lett., 195, 322–328.

Becker, K.H., Engelhardt, B., Geiger, H., Kurtenbach, R. & Wiesen, P. (1993).Chem. Phys. Lett., 210, 135–140.

Becker, K.H., Engelhardt, B., Wiesen, P., & Bayes, K.D. (1989). Chem. Phys.Lett., 154, 342–348.

Becker, K.H., Geiger, H. & Wiesen, P. (1996). Int. J. Chem. Kinet., 28, 115–123.

Becker, K.H., Kurtenbach, R., Schmidt, F. & Wiesen, P. (1995a). Chem. Phys.Lett., 235, 230–234.

Becker, K.H., Kurtenbach, R., Schmidt, F. & Wiesen, P. (1997). Ber. Bunsenges.,101, 128–133.

Becker, K.H., Kurtenbach, R., & Wiesen, P. (1992b). Chem. Phys. Lett., 198,424–428.

Becker, K.H., Kurtenbach, R. & Wiesen, P. (1995b). J. Phys. Chem., 99, 5986–5991.

Benson, S.W. (1976). Thermochemical Kinetics., New York, John Wiley & Sons.

Benson, S.W. (1985). J. Phys. Chem., 89, 4366–4369

Berman, M.R., Fleming, J.W., Harvey, A.B., & Lin, M.C. (1982). 19th Sympo-sium (International) on Combustion, pp. 73–79.

Berman, M.R., & Lin, M C. (1983). J. Phys. Chem., 87, 3933–3942.

Bian, J., Vandooren, J., & Van Tiggelen, P.J. (1990). 23rd Symposium (Interna-tional) on Combustion, pp. 379–386.

Biggs, P., Canosa-Mas, C.E., Fracheboud, J.-M., Parr, A.D., Shallcross, D.E.,Wayne, R.P. & Caralp, F. (1993). J. Chem. Soc. Faraday Trans., 89, 4163–4169.

Blauwens, J., Smets, B., & Peeters, J. (1978). 16th Symposium (International) onCombustion, pp. 1055–1064.

Bohland, T., Temps, F. & Wagner, H.Gg. (1984). Ber. Bunsenges. Phys. Chem.,88, 455–458.

Boullart, W., Nguyen, M.T. & Peeters, J. (1994). J. Phys. Chem., 98, 8036–8043.

Bowman, C.T. (1992). 24th Symposium (International) on Combustion, pp. 859–878.

Bozzelli, J.W., Chang, A.Y. & Dean, A.M. (1997). Int. J. Chem. Kinet., 29,161–170.

Bozzelli, J.W., & Dean, A.M. (1989). J. Phys. Chem., 93, 1058–1065.

Bozzelli, J.W., & Dean, A.M. (1990). J. Phys. Chem., 94, 3313–3317.

Bozzelli, J.W., & Dean, A.M. (1993). J. Phys. Chem., 97, 4427–4441.

Page 194: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 319

Bozzelli, J.W., & Dean, A.M. (1995). Int. J. Chem. Kinet., 27, 1097–1110.

Bozzelli, J.W., & Ritter, E.R. (1993). Radicalc---Computer code to calculateentropy & heat capacity contributions to transition states & radical speciesfrom changes in vibrational frequencies, barriers, moments of inertia, andinternal rotations. Proceedings of the Eastern States Section of the CombustionInstitute, 103, 459.

Bradley, K.S., McCabe, P., Schatz, G.C. & Walch, S.P. (1995). J. Chem. Phys.,102, 6696–6705.

Branch, M.C., Miller, J.A., & Kee, R.J. (1982). Combust. Sci. Technol., 29,147–165.

Brownsword, R.A., Gatenby, S.D., Herert, L.B., Smith, I.W.M., Stewart, D.W.A.& Symonds, A. C.(1996). J. Chem. Soc. Faraday Trans., 92, 723–727.

Brownsword, R.A., Hancock, G., & Heard, D.E. (1997). J. Chem. Soc. Farad.Trans., 93, 2473–2475.

Bulatov, V.P., Ioffe, A.A., Lozovsky, V.A., & Sarkisov, O.M. (1989). Chem. Phys.Lett., 161, 141–146.

Butler, J.E., Fleming, J.W., Goss, L.P., & Lin, M.C. (1981). Chem. Phys., 56,355–365.

Campbell, I.M., & Handy, B.J. (1975). J. Chem. Soc. Faraday Trans. I, 71,2097.

Carpenter, B.K. (1993). J. Am. Chem. Soc., 115, 9806–9807.

Cattolica, R.J., Smooke, M.D., & Dean, A.M. (1982). A Hydrogen--NitrousOxide Flame Study, Sandia National Laboratories Report SAND82–8776.

Cheskis, S.G., & Sarkisov, O.M. (1979). Chem. Phys. Lett., 67, 72–76.

Cohen, N. (1987). Int. J. Chem. Kinet., 19, 319–362.

Cohen, N. (1990). A Revised Model for Transition State Theory Calculationsfor Rate Coefficients of OH with Alkanes, Aerospace Corporation ReportATR–90(7179)–1.

Cohen, N., & Westberg, K.R. (1991). J. Phys. Chem. Ref. Data, 20, 1211–1311.

Cohen, N. (1992). Chemical Kinetic Data Sheets for High--Temperature ChemicalReactions, Vol. III., Aerospace Corporation Report ATR–91(7189)–2.

Cooper, W.F., & Hershberger, J.F. (1992). J. Phys. Chem., 96, 771–775.

Cooper, W.F., Park, J., & Hershberger, J.F. (1993). J. Phys. Chem., 97, 3283–3290.

Darwin, D.C., Young, A.T., Johnston, H.S., & Moore, C.B. (1989). J. Phys.Chem., 93, 1074–1078.

Davidson, D.F., & Hanson, R.K. (1990a). Int. J. Chem. Kinet., 22, 843–861.

Davidson, D.F., & Hanson, R.K. (1990b). 23rd Symposium (International) onCombustion, pp. 267–273.

Page 195: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

320 Anthony M. Dean, Joseph W. Bozzelli

Davidson, D.F., Kohse–H�oinghaus, K., Chang, A.Y., & Hanson, R.K. (1990). Int.J. Chem. Kinetics, 22, 513–535.

Davidson, D.F., DiRosa, M.D., Chang, A.Y., & Hanson, R.K. (1991a). 18thInternational Symposium on Shock Waves, Sendai, pp. 813–818.

Davidson, D.F., Dean, A.J., DiRosa, M.D., & Hanson, R.K. (1991b). Int. J.Chem. Kinet., 23, 1035–1050.

Davies, J.W., Green, N.J.B., & Pilling, M.J. (1991). J. Chem. Soc. FaradayTrans., 87, 2317–2324.

Dean, A.J., Hanson, R.K., & Bowman, C.T. (1990). 23rd Symposium (Interna-tional) on Combustion, pp. 259–265.

Dean, A.J., Hanson, R.K., & Bowman, C.T. (1991). J. Phys. Chem., 95,3180–3189.

Dean, A.M. (1985). J. Phys. Chem., 89, 4600–4608.

Dean, A.M., Bozzelli, J.W., & Ritter, E.R. (1991). Combust. Sci. Tech., 80,63–85.

Dean, A.M., Chou, M.S., & Stern, D. (1984). Int. J. Chem. Kinet., 16, 633– 653.

Dean, A.M., Hardy, J.E., & Lyon, R.K. (1982). 19th Symposium (International)on Combustion, pp. 97–105.

Dean, A.M., Johnson, R.L., & Steiner, D.C. (1980). Combust. Flame, 37, 41–62.

Dean, A.M., & Steiner, D.C. (1977). J. Chem. Phys., 66, 598–604.

Dean, A.M., & Westmoreland, P.R. (1987). Int. J. Chem. Kinet., 19, 207–228.

Diau, E.W., Halbgewachs, M.J., Smith, A.R. & Lin, M.C. (1995). Int. J. Chem.Kinet., 27, 867–881.

Diau, E.W.G. & Smith, S.C. (1996). J. Phys. Chem., 100, 12349–12354.

Diau, E.W.-G., Tso, T.-Y., & Lee, Y.-P. (1990). J. Phys. Chem., 94, 5261–5265.

Diau, E.W., Yu, T., Wagner, M.A.G., & Lin, M.C. (1994). J. Phys. Chem., 98,4034–4042.

Dolson, D.A. (1986). J. Phys. Chem, 90, 6714–6718.

Dorthe, G., Caubet, P., Vias, T., Barrere, B., & Marchais, J. (1991) J. Phys.Chem., 95, 5109–5116

Drake, M.C., & Blint, R.J. (1991a). Combust. Sci. Tech., 75, 261–285.

Drake, M.C., & Blint, R.J. (1991b). Combust. Flame, 83, 185–203.

Drake, M.C., Ratcliffe, J.W., Blint, R.J., Carter, C.D., & Laurendeau, N.M. (1990).23rd Symposium (International) on Combustion, pp. 387–395.

Dransfeld, P., Hack, W., Kurzke, H., Temps, F., & Wagner, H.Gg. (1984). 20thSymposium (International) on Combustion, pp. 655–663.

Dransfeld, P., & Wagner, H.Gg. (1987). Z. Phys. Chem. (Munich), 153, 89.

Duan, X. & Page, M. (1995). J. Chem. Phys., 102, 6121–6127.

Page 196: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 321

Duffy, B.L. & Nelson, P.F. (1996). 26th Symposium (International) on Combus-tion, pp. 2099–2108.

Duran, R.P., Amorebieta, V.T. & Colussi, A.J. (1988). J. Phys. Chem., 92, 636.

Durant, J.L. (1994). J. Phys. Chem., 98, 518–521.

East, A.L.L., & Allen, W.D. (1993). J. Chem. Phys., 99, 4638–4650.

Etzkorn, T., Muris, S., Wolfrum, J., Dembny, C., Bockhorn, H., Nelson, P.F.,Attia–Shanin, A., & Warnatz, J. (1992). 24th Symposium (International) onCombustion, pp. 925–932.

Felder, W., & Young, R.A. (1972). J. Chem. Phys., 57, 572.

Fenimore, C.P. (1970). 13th Symposium (International) on Combustion, pp. 373–380.

Fenimore, C.P. (1976). Comb. Flame, 26, 249–256.

Fenimore, C.P. (1980). Comb. Flame, 37, 245.

Flatness, S.A. & Kramlich, J.C. (1996). 26th Symposium (International) onCombustion, pp. 567–573.

Foresman, J.B., & Frisch, �, (1996) Exploring Chemistry with Electronic Struc-ture Methods, Second Edition,, Gaussian, Pittsburgh.

Forst, W. (1972). Theory of Unimolecular Reactions, Academic Press, NewYork.

Forster, R., Frost, M., Fulle, D., Hamann, H.F., Hippler, H., Schlepegrell, A. &Troe, J. (1995). J. Chem. Phys., 103, 2949-2957.

Frank, P., & Just, T. (1985). Ber. Bunsenges. Phys. Chem., 89, 181–185.

Frerichs, H., Koch, R., Schliephake, V., Tappe, M., & Wagner, H.Gg. (1990). Z.Phys. Chem., 166, 145–155.

Fujii, N., Miyama, H., Koshi, M., & Asaba, T. (1981). 18th Symposium(International) on Combustion, pp. 873–883.

Fujii, N., Chiba, K., Uchida, S., & Miyama, H. (1986). Chem. Phys. Lett., 127,141–144.

Fujii, N., Sagawai, S., Sato, T., Nosaka, Y., & Miyama, H. (1989). J. Phys. Chem.,93, 5474–5478.

Gardiner, W.C., Jr. (1977) Accts. Chem. Res., 10, 326.

Gardiner, W.C., In The 12th IMACS World Congress on Scientific Computation,Vichnevetsky, R., Borne, P., & Vignes, J., Eds., IMACS, Paris, 1988, p. 58.

Garo, A., Hilaire, C., & Puechberty, D. (1992). Combust. Sci. Tech., 86, 87–103.

Gehring, V.M., Hoyermann, K., Wagner, H.Gg., & Wolfrum, J. (1971). Ber.Bunsenges. Phys. Chem., 75, 1287.

Gilbert, R.G., Luther, K., & Troe, J. (1983) Ber. Bunsenges. Phys. Chem., 87,169–177.

Gilbert, R.G. & Smith, S.C. (1990). Theory of Unimolecular and RecombinationReactions., Oxford, Blackwell Scientific Publications.

Page 197: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

322 Anthony M. Dean, Joseph W. Bozzelli

Glarborg, P., Dam-Johansen, K., Miller, J.A., Kee, R.J., & Coltrin, M.E. (1994).Int. J. Chem. Kinet., 26, 421-436

Glarborg, P., Dam-Johansen, K. & Miller, J.A. (1995). Int. J. Chem. Kinet., 27,1207–1220.

Golden, D.M. & Larson, C.W. (1985) 20th Symposium (International) on Com-bustion, pp. 595–603.

Gordon, M.S., Truong, T.N. & Pople, J.A. (1986). Chem. Phys. Lett., 130,245–248.

Grussdorf, J., Temps, F. & Wagner, H.Gg. (1997). Ber. Bunsenges. Physik.Chemie, 101, 134–138.

Hack, W., & Kurzke, H. (1985a). Ber. Bunsenges. Phys. Chem., 89, 86–93.

Hack, W., Kurzke, H., & Wagner, H.Gg. (1985b). J. Chem. Soc. Faraday Trans.2, 81, 949–961.

Hack, W., Rouverirolles, P., & Wagner, H.Gg. (1986). J. Phys. Chem., 90,2505–2511.

Hack, W., Wagner, H.Gg., & Zasypkin, A. (1994). Ber. Bunsenges. Phys. Chem.,98, 156–164.

Halbgewachs, M.J., Diau, E.W.G., Mebel, A.M., Lin, M.C. & Melius, C. F. (1996).26th Symposium (International) on Combustion, pp. 2109–2115.

Hall, J.L., Zeitz, D., Stephens, J.W., Kasper, J.V.V., Glass, G.P., Curl, R.F., &Tittel, F.K. (1986). J. Phys. Chem., 90, 2501–2505.

Hanson, R.K., & Salimian, S. (1984). In Combustion Chemistry, W.C. Gardiner,Jr., Ed., Springer–Verlag, New York, 361–421.

Harrington, J.E., Noble, A.R., Smith, G.P., Jeffries, J.B. & Crosley, D.R. (1995).Evidence for a New NO Production Mechanism in Flames. Fall Meeting ofthe Western States Section of the Combustion Institute, Stanford, California,Paper 95F-193.

Harrington, J.E., Smith, G.P., Berg, P.A., Noble, A.R., Jeffries, J.B. & Crosley,D.R. (1996). 26th Symposium (International) on Combustion, pp. 2133–2138.

Harrison, J.A., Whyte, A.R., & Phillips, L.F. (1986). Chem. Phys. Lett., 129,346–352.

Harrison, J.A., & Maclagan, R.G.A.R. (1990). J. Chem. Soc. Faraday Trans., 86,3519–3523.

Hayhurst, A.N., & Vince, I.M. (1980). Prog. Energy Combust. Sci., 6, 35–51.

Haynes, B.S. (1978). Astronaut. Aeronaut., 62, 359–394.

He, Y., & Lin, M.C. (1992). Int. J. Chem. Kinet., 24, 743–760.

Hennig, G., Klatt, M., Spindler, B. & ~Wagner, H.Gg. (1995). Ber. Bunsenges.Phys. Chem., 99, 651–657.

Page 198: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 323

Hennig, G., R�ohrig, M. & Wagner, H.Gg. (1993). Ber. Bunsenges. Phys. Chem.,97, 830–832.

Hennig, G., & Wagner, H.Gg. (1994). Ber. Bunsenges. Phys. Chem., 98,749–753.

Hennig, G. & Wagner, H.Gg. (1995). Ber. Bunsenges. Phys. Chem., 99,863–869.

Herron, J.T. (1988). J. Phys. Chem. Ref. Data, 17, 967–1026.

Hidaka, Y., Taskuma, H., & Suga, M. (1985). Bull. Chem. Soc. Jpn., 58,2911–2916.

Higashihara, T., Gardiner, W.C., Jr., & Hwang, S.M. (1987). J. Phys. Chem., 91,1900–1905.

Hoffmann, A., Wagner, H.Gg., Wolff, T., & Hwang, S.M. (1990). Ber. Bunsenges.Phys. Chem., 94, 1407–1411.

Howard, C.J. (1980). J. Amer. Chem. Soc., 102, 6937.

Hsu, C.C., Lin, M.C., Mebel, A.M. & Melius, C.F. (1997). J. Phys. Chem. A,101, 60–66.

Hwang, S.M., Higashihara, T., Shin, K.S., & Gardiner, W.C., Jr. (1990) J. Phys.Chem., 94, 2883–2889.

Jeffries, J.B., & Smith, G.P. (1986). J. Phys. Chem., 90, 487.

Jemi–Alade, A.A., & Thrush, B.A. (1990). J. Chem. Soc. Faraday Trans., 86,3355–3363.

Jodkowski, J.T., Ratajczak, E., Sillesen, A., & Pagsberg, P. (1993). Chem. Phys.Lett., 203, 490–496.

Johnsson, J.E., Glarborg, P., & Dam–Johansen, K. (1992). 24th Symposium(International) on Combustion, pp. 917–923.

Juang, D.Y., Lee, J.-S. & Wang, N.S. (1995). Int. J. Chem. Kinet., 27, 1111–1120.

Just, T. (1994). 25th Symposium (International) on Combustion, pp. 687–704.

Kaiser, E.W., & Japar, S.M. (1978). J. Phys. Chem., 82, 2753.

Kaiser, E.W. (1993). J. Phys. Chem., 97, 11681–11688.

Kaiser, E.W. (1995). J. Phys. Chem., 99, 707–711.

Kazakov, A., Wang, H., & Frenklach, M. (1994). J. Phys. Chem., 98, 10598.

Kee, R.J., Grcar, J.F., Smooke, M.D. & Miller, J.A. (1985). A Fortran Programfor Modeling Steady Laminar One-Dimensional Premixed Flames. SandiaNational Laboratory Report SAND85–8240 UC401.

Kee, R.J., & Miller, J.A. (1987). The Chemkin Thermodynamic Data Base.Sandia National Laboratory Report SAND87–8215 UC4.

Kee, R.J., Rupley, F.M. & Miller, J.A. (1989). Chemkin-II. A Fortran ChemicalKinetics Package for the Analysis of Gas-Phase Chemical Kinetics. SandiaNational Laboratory Report SAND89–8009 UC-401.

Page 199: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

324 Anthony M. Dean, Joseph W. Bozzelli

Kerr, J.A., & Moss, S.J., Ed. (1981). Handbook of bimolecular and termoleculargas phase reactions. CRC Press, Boca Raton.

Khe, P.V., Soulignac, J.C. & Lesclaux, R. (1977). J. Phys. Chem., 81, 210.

Kimball–Linne, M. A., & Hanson, R. K. (1986). Combust. Flame, 64, 337–351.

Knispel, R., Koch, R., Siese, M., & Zetzsch, C. (1990). Ber. Bunsenges. Phys.Chem., 94, 1375.

Ko, T., Marshall, P., & Fontijn, A. (1990). J. Phys. Chem., 94, 1401–1404.

Ko, T., & Fontijn, A. (1991). J. Phys. Chem., 95, 3984–3987.

Koizumi, H., Schatz, G.C., & Walch, S.P. (1991). J. Chem. Phys., 95, 4130–4135.

Koshi, M., Bando, S., Saito, M., & Asaba, T. (1978). 17th Symposium (Interna-tional) on Combustion, pp. 553–562.

Koshi, M., Yoshimura, M., Fukuda, K., Matsui, H., Saito, K., Watanabe, M.,Imamura, A., & Chen, C. (1990). J. Chem. Phys., 93, 8703–8708.

Lambrecht, R., & Herschberger, J. F. (1994). J. Phys. Chem., 98, 8406–8410.

Lang, V.I. (1992). J. Phys. Chem., 96, 3047–3050.

Larson, C.W., Patrick, R., & Golden, D.M. (1984). Combust. Flame, 58, 229.

Laufer, A., & Bass, A.M. (1974). J. Phys. Chem., 78, 1344–1348.

Laufer, A.H., Gardner, E.P., Kwok, T.L. & Yung, Y.L. (1983). Icarus, 56, 560.

Lesclaux, R. (1984). Rev. Chem. Intermed., 5, 347–392.

Lichtin, D.A., Berman, M.R., & Lin, M.C. (1984). Chem. Phys. Lett., 108,18–24.

Lifshitz, A., Bidani, M., Carroll, H.F., Hwang, S.M., Fu, P.-Y., Shin, K.S., &Gardiner, W.C., Jr. (1991). Comb. Flame, 86, 229–236.

Lifshitz, A., Tamburu, C., Frank, P., & Just, T. (1993). J. Phys. Chem., 97,4085–4090.

Lightfoot, P.D., & Pilling, M.J. (1987.) J. Phys. Chem., 91, 3373.

Lillich, H., Schuck, A., Volpp, H.-R. & Wolfrum, J. (1994). 25th Symposium(International) on Combustion, pp. 993–1001.

Lin, M.C., He, Y., & Melius, C.F. (1992a). Int. J. Chem. Kinet., 24, 489–516.

Lin, M.C., He, Y., & Melius, C.F. (1992b). Int. J. Chem. Kinet., 24, 1103–1107.

Lin, M.C., He, Y., & Melius, C.F. (1993). J. Phys. Chem., 97, 9124–9128.

Lindackers, D., Burmeister, M., & Roth, P. (1990). 23rd Symposium (Interna-tional) on Combustion, pp. 251–257.

Linder, D.P., Duan, X. & Page, M. (1995). J. Phys. Chem., 99, 11458–11463.

Lindholm, N. & Hershberger, J.F. (1997). J. Phys. Chem. A, 101, 4991–4995.

Lindstedt, R.P., Lockwood, F.C. & Selim, M.A. (1995). Combust. Sci. Technol.,108, 231–254.

Louge, M.Y., & Hanson, R.K. (1984a). Combust. Flame, 58, 291–300.

Page 200: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 325

Louge, M.Y., & Hanson, R.K. (1984b). Int. J. Chem. Kinet., 16, 231–250.

Lutz, A.E., Kee, R.J. & Miller, J.A. (1991) SENKIN: A Fortran Program forPredicting Homogeneous Gas Phase Chemical Kinetics with SensitivityAnalysis. Sandia National Laboratory Report SAND87-8248.

Lyon, R. K. (1975). U. S. Patent No. 3,900,554.

Lyon, R.K. (1987). Div. Fuel Chem., 32, 433.

Lyon, R.K. & Hardy, J.E. (1986). Ind. Eng. Fundam., 25, 19–24.

Mackie, J.C., Colket, M.B., & Nelson, P.F. (1990). J. Phys. Chem, 94, 4099–4106.

Mallard, W.G., Westley, F., Herron, J.T., Hampson, R.F., & Frizzell, D.H. (1993).NIST Chemical Kinetics Data Base: Version 5.0. National Institute ofStandards and Technology, Gaithersburg.

Malte, P.C., & Pratt, D.T. (1974). Combust. Sci. Tech., 9, 221.

Manaa, M.R., & Yarkony, D.R. (1991). J. Chem. Phys., 95, 1808–1816.

Manaa, M.R., & Yarkony, D.R. (1992). Chem. Phys. Lett., 188, 352–358.

Marshall, P., & Fontijn, A. (1986). J. Chem. Phys., 85, 2637–2643.

Marshall, P., Fontijn, A., & Melius, C. F. (1987). J. Chem. Phys., 86, 5540–5549.

Marshall, P., Ko, T., & Fontijn, A. (1989). J. Phys. Chem., 93, 1922–1927.

Marston, G., Nesbitt, F.L., & Stief, L.J. (1989a). J. Chem. Phys., 91, 3483–3491.

Marston, G., Nesbitt, F.L., Nava, D.F., Payne, W.A., & Stief, L.J. (1989b). J.Phys. Chem., 93, 5769–5774.

Martin, J.M.L., & Taylor, P.R. (1993). Chem. Phys. Lett., 209, 143–150.

Matsui, Y., & Yuuki, A. (1985). Japan J. Appl. Phys, 24, 598–603.

Meads, R.F., Machigan, G.A.R., & Phillips, L.F. (1993). J. Phys. Chem., 97,3257–3265.

Meagher, N.E., Anderson, W.R., Goumri, A. & Fontijn, A. (1997). Kinetics ofthe O(3P) + N2O Reaction at Intermediate Temperatures., 4th InternationalConference on Chemical Kinetics, Gaithersburg, Maryland.

Mebel, A.M., Diau, E.W.G., Lin, M.C. & Morokuma, K. (1996). J. Phys. Chem.,100, 7517–1725.

Mebel, A.M., Lin, M.C., Morokuma, K. & Melius, C.F. (1996). Int. J. Chem.Kinet., 28, 693–703.

Medhurst, L.J., Garland, N.L., & Nelson, H.H. (1993). J. Phys. Chem., 97,12275–12281.

Melius, C.F. (1988). BAC--MP4 Heats of Formation & Free Energies. Privatelycirculated report.

Melius, C.F. (1993). BAC--MP4 Heats of Formation & Free Energies. Privatelycirculated report.

Page 201: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

326 Anthony M. Dean, Joseph W. Bozzelli

Melius, C.F., & Binkley, J.S. (1984a). In The Chemistry of Combustion Processes,T.M. Sloane , Ed., ACS Symposium Series, 103–115.

Melius, C.F., & Binkley, J.S. (1984b). 20th Symposium (International) onCombustion, pp. 575–583.

Mertens, J.D., Chang, A.Y., Hanson, R.K., & Bowman, C.T. (1989). Int. J. Chem.Kinet., 21, 1049–1067.

Mertens, J.D., Chang, A.Y., Hanson, R.K., & Bowman, C.T. (1991a). Int. J.Chem. Kinet., 23, 173–196.

Mertens, J.D., Chang, A.Y., Hanson, R.K., & Bowman, C.T. (1992a). Int. J.Chem. Kinet., 24, 279–295.

Mertens, J.D., Dean, A.J., Hanson, R.K., & Bowman, C.T. (1992b). 24thSymposium (International) on Combustion, pp. 701–710.

Mertens, J.D. & Hanson, R.K. (1996). 26th Symposium (International) onCombustion, pp. 551–558.

Mertens, J.D., Kohse–H�oinghaus, K., Hanson, R.K., & Bowman, C.T. (1991b).Int. J. Chem. Kinet., 23, 655–668.

Messing, I., Filseth, S.V., Sadowski, C.M., & Carrington, T. (1981). J. Chem.Phys., 74, 3874.

Meunier, H., Pagsberg, P. & Sillesen, A. (1996). Chem. Phys. Lett., 261,277–282.

Michael, J.V., Klemm, R.B., Brobst, W.D., Bosco, S.R., & Nava, D.F. (1985). J.Phys. Chem., 89, 3335–3337.

Michael, J.V., & Lim, K.P. (1992). J. Chem. Phys., 97, 3228–3234.

Michael, J.V., Sutherland, J.W., & Klemm, R.B. (1986). J. Phys. Chem., 90,497–500.

Michaud, M.G., Westmoreland, P.R., & Feitelberg, A.S. (1992). 24th Symposium(International) on Combustion, pp. 879–887.

Miller, J.A., & Bowman, C.T. (1989). Prog. Energy Combust. Sci., 15, 287–338.

Miller, J A., & Bowman, C.T. (1991). Int. J. Chem. Kinet., 23, 289–313.

Miller, J.A., Branch, M.C., McLean, W.J., Chandler, D.W., Smooke, M.D., & Kee,R.J. (1984). 20th Symposium (International) on Combustion, pp. 673–684.

Miller, J.A., & Melius, C.F. (1986). 21st Symposium (International) on Combus-tion, pp. 919–927.

Miller, J.A., & Melius, C.F. (1992a). 24th Symposium (International) on Com-bustion, pp. 719–726.

Miller, J.A., & Melius, C.F. (1992b). Int. J. Chem. Kinet., 24, 421–432.

Miller, J.A., Melius, C.F., & Glarborg, P. 30, 223–228, 1998

Miller, J.A., Smooke, M.D., Green, R.M., & Kee, R.J. (1983). Combust. Sci.Technol., 34, 149–176.

Morley, C. (1980). 18th Symposium (International) on Combustion, pp. 23–32.

Page 202: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 327

Myerson, A.L. (1974). 14th Symposium (International) on Combustion, pp. 219–228.

Natarajan, K., Mick, H.J., Woiki, D. & Roth, P. (1994). Combust. Flame, 99,610.

Natarajan, K., & Roth, P. (1988). 21st Symposium (International) on Combustion,pp. 729–737.

Natarajan, K., Thielen, K., Hermanns, H.D. & Roth, P. (1986). Ber. Bunsenges.Phys. Chem., 90, 533.

Nesbitt, F.L., Marston, G., & Stief, L.J. (1990). J. Phys. Chem., 94, 4946.

Nguyen, M.T., Boullart, W. & Peeters, J. (1994). J. Phys. Chem., 98, 8030–8035.

Nguyen, M.T., Sengupta, D. & Vanquickenborne, L.G. (1996). J. Phys. Chem.,100, 10956–10966.

Nguyen, M.T., Sengupta, D., Vereecken, L., Peeters, J. & Vanquickenborne, L.G.(1996). J. Phys. Chem., 100, 1615–1621.

Okada, S., Tezaki, A., Miyoshi, A., & Matsui, H. (1994). J. Chem. Phys, 101,9582–9588.

Okada, S., Yamasaki, K., Matsui, H., Saito, K., & Okada, K. (1993). Bull. Chem.Soc. Japan, 66, 1004–1011.

Oref, I. (1989). J. Phys. Chem., 87, 3465.

Page, M., & Soto, M.R. (1993). J. Chem. Phys., 99, 7709–7717.

Pagsberg, P., Sztuba, B., Ratajczak, E., & Sillesen, A. (1991). Acta Chem.Scandinavica, 45, 329–334.

Park, J., & Hershberger, J.F. (1993a). J. Phys. Chem., 97, 13647–13652.

Park, J., & Hershberger, J.F. (1993b). J. Chem. Phys., 99, 3488–3493.

Park, J. & Hershberger, J.F. (1994). Chem. Phys. Lett., 218, 537–543.

Park, J. & Lin, M.C. (1996). J. Phys. Chem., 100, 3317–3319.

Park, J. & Lin, M.C. (1997). J. Phys. Chem. A, 101, 2643–2647.

Patel–Misra, D., & Dagdigian, P.J. (1991). Chem. Phys. Lett., 185, 387–392.

Patrick, R., & Golden, D.M. (1984). J. Phys. Chem., 88, 491–495.

Pawlowska, Z., Gardiner, W.C., & Oref, I. (1993). J. Phys. Chem., 97, 5024.

Peeters, J., Van Look, H. & Ceursters, B. (1996). J. Phys. Chem., 100, 15124–15129.

Perry, R.A. (1984). Chem. Phys. Lett., 106, 223–228.

Perry, R.A., & Melius, C.F. (1984). 20th Symposium (International) on Combus-tion, pp. 639–646.

Perry, R.A., & Siebers, D.L. (1986). Nature, 324, 657.

Pouchan, C., Lam, B., & Bishop, D.M. (1987). J. Phys. Chem., 91, 4809–4813.

Prezhdo, O. (1995). J. Phys. Chem., 99, 8633–8637.

Page 203: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

328 Anthony M. Dean, Joseph W. Bozzelli

Quandt, R.W. & Hershberger, J.F. (1996). J. Phys. Chem., 100, 9407–9411.

Qin, Z., Yang, H., Lissianski, V., Gardiner, W.C., & Shin, K.S. (1997). Chem.Phys. Lett.,, 110-113.

Reid, R.C., Prausnitz, J.M., & Poling, B.E. (1987). Properties of Gases andLiquids, McGraw Hill, New York.

Ritter, E. R. (1989). Ph.D. Thesis, New Jersey Institute of Technology.

Ritter, E.R., & Bozzelli, J.W. (1991). Int. J. Chem. Kinet., 23, 767–778.

Robinson, P.J., & Holbrook, K.A. (1972). Unimolecular Reactions, WileyInterscience, London.

Rodgers, A.S. & Smith, G.P. (1996). Chem. Phys. Lett., 253, 313–321.

R�ohrig, M., Petersen, E.L., Davidson, D.F. & Hanson, R.K. (1996). Int. J. Chem.Kin., 28, 599–608.

R�ohrig, M., Petersen, E.L., Davidson, D.F. & Hanson, R.K. (1997). Int. J. Chem.Kinet., 29, 483–493.

R�ohrig, M., R�omming, H.-J., & Wagner, H.Gg. (1994). Ber. Bunsenges. Phys.Chem., 98, 1332–1334.

R�omming, H.-J. & Wagner, H.G. (1996). 26th Symposium (International) onCombustion, pp. 559–566.

Roose, T.R. (1981). Ph.D. Thesis, Stanford University.

Ross, S.K., Sutherland, J.W., Kuo, S.-C. & Klemm, R.B. (1997). J. Phys. Chem.A, 101, 1104–1116.

Roth, P., & Just, T. (1984). 20th Symposium (International) on Combustion, pp.807–818.

Sarkisov, O.M., Cheskis, S.G., Nadtochenko, V.A., Sviridenkov, E.A. & Vedeneev,V.I. (1984). Arch. Combust., 4, 111.

Sausa, R.C., Anderson, W.R., Dayton, D.C., Faust, C.M., & Howard, S. L. (1993).Combust. Flame, 94, 407–425.

Schranz, H.W., Nordholm, S., & Hamer, N.D. (1982). Int. J. Chem. Kinet., 14,543–564.

Schuck, A., Volpp, H.-R. & Wolfrum, J. (1994). Combust. Flame, 99, 491.

Seideman, T. (1994). J. Chem. Phys., 101, 3662–3670.

Seideman, T. & Walch, S.P. (1994). J. Chem. Phys., 101, 3656–3661.

Seidler, V., Temps, F., Wagner, H.Gg., & Wolf, M. (1989). J. Phys. Chem., 93,1070–1073.

Sengupta, D. & Chandra, A.K. (1994). J. Chem. Phys., 101, 3906-3915.

Silver, J.A., & Kolb, C.E. (1987). J. Phys. Chem., 91, 3713–3714.

Sims, I.R., & Smith, I.W.M. (1988). J. Chem. Soc. Faraday Trans. 2, 84, 527.

Slagle, I.R., Sarzynski, D., & Gutman, D. (1987). J. Phys. Chem., 91, 4375.

Page 204: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 329

Soto, M.R., Page, M.J., & McKee, M.L. (1991). Chem. Phys. Lett., 153,415–426.

Soto, M.R., & Page, M. (1993). J. Chem. Phys., 97, 7287–7296.

Stapf, D. & Leuckel, W. (1996). 26th Symposium (International) on Combustion,pp. 2083–2090.

Stephens, J.W., Morter, C.L., Farhat, S.K., Glass, G.P., & Curl, R.F. (1993). J.Phys. Chem., 97, 8944–8951.

Stewart, P.H., Larson, C.W. & Golden, D.M. (1989). Combust. Flame, 75, 25-31.

Stief, L.J., Marston, G., Nava, D.F., Payne, W.A., & Nesbitt, F.L. (1988). Chem.Phys. Lett., 147, 570–574.

Stothard, N., Humpfer, R. & Grotheer, H.-H. (1995). Chem. Phys. Lett., 240,474–480.

Sutherland, J.W., & Klemm, R.B. (1987). 16th International Symposium on ShockTubes and Waves, p. 395.

Sutherland, J.W., Patterson, P.M., & Klemm, R.B. (1990). J. Phys. Chem., 94,2471–2475.

Szekely, A., Hanson, R.K., & Bowman, C.T. (1984). 20th Symposium (Interna-tional) on Combustion, pp. 647–654.

Thielen, K., & Roth, P. (1984). 20th Symposium (International) on Combustion,pp. 685–693.

Troe, J. (1979). J. Phys. Chem., 83, 114–126

Tsang, W. (1991). J. Phys. Chem. Ref. Data, 20, 221–273.

Tsang, W. (1992). J. Phys. Chem. Ref. Data, 21, 753–791.

Tsang, W., & Hampson, R.F. (1986). J. Phys. Chem. Ref. Data, 15, 1087–1279.

Tsang, W., & Herron, J.T. (1991). J. Phys. Chem. Ref. Data, 20, 609–663.

Tyndall, G.S., Orlando, J.J., Nickerson, K.E., Cantrell, C.A., & Calvert, J. G.(1991). J. Geophys. Res., 96, 20761–20768.

Unfried, K.G., Glass, G.P., & Curl, R.F. (1990). Chem. Phys. Lett., 173,337–341.

Vaghjiani, G.L. (1996). J. Chem. Phys., 104, 5479–5489.

Vakhtin, A.B., & Petrov, A.K. (1990). Spectrochimica Acta., 46A, 603–606.

Vandooren, O.M., Sarkisov, O.M., Balakhnin, V.P., & Van Tiggelen, P.J. (1991).Chem. Phys. Lett., 184, 294–300.

Venkatesh, P.K., Chang, A.Y., Dean, A.M., Cohen, M.H. & Carr, R.W. (1997).A.I.Ch.E. Journal, 43, 1331–1340.

Vinckier, C., & Debruyn, W. (1979). J. Phys. Chem., 83, 2057–2062.

Votsmeier, M., Song, S., Hanson, R.K., & Bowman, C.T. (1998). SensitiveDetection of NH2 in Shock Tube Kinetics Experiments Using FrequencyModulation Spectroscopy, Work-in-Progress Poster Session W4F06, 27thSymposium (International) on Combustion, Boulder, to be published.

Page 205: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

330 Anthony M. Dean, Joseph W. Bozzelli

Wagal, S.S., Carrington, T., Filseth, S.V., & Sadowski, C. M. (1982). Chem.Phys., 69, 61–70.

Wagner, A.F., Slagle, I.R., Sarzynski, D., & Gutman, D. (1990). J. Phys. Chem.,94, 1853–1868.

Wagner, H.Gg., Welzbacher, U.. & Zellner, R. (1976). Ber. Bunsenges. Phys.Chem., 80, 1023

Walch, S.P., Duchovic, R.J., & Rohlfing, C.M. (1989). J. Chem. Phys., 90,3230–3240.

Walch, S.P. (1993a). J. Chem. Phys., 98, 1170–1177.

Walch, S.P. (1993b). J. Chem. Phys., 99, 5295–5300.

Walch, S.P. (1993c). Chem. Phys. Lett., 208, 214–218.

Walia, N. & Kakker, R. (1990). Indian J. Chem., 29B, 1007.

Walia, N. & Kakker, R. (1991). Indian J. Chem., 30A, 308.

Wallington, T.J., Maricq, M.M., Ellermann, T., & Nielsen, O.J. (1992). J. Phys.Chem., 96, 982–986.

Wang, H., & Frenklach, M. (1993). Chem. Phys. Lett., 205, 271.

Wang, N.S., Yang, D.L., & Lin, M.C. (1989). Chem. Phys. Lett., 163, 480.

Wang, N.S., Yang, D.L., & Lin, M.C. (1991). Int. J. Chem. Kinet., 23, 151.

Warnatz, J. (1983). Combust. Sci. Technol., 34, 177.

Warnatz, J. (1984). In Combustion Chemistry, W.C. Gardiner., Jr , Ed., Springer–Verlag, New York.

Washida, N. (1980). J. Chem. Phys., 73, 1665–1672.

Westmoreland, P.R., Howard, J.B., Longwell, J.P., & Dean, A.M. (1986). AIChEJournal, 32, 1971–79.

Weston, R.E. (1986). Int. J. Chem. Kinet., 18, 1259–1276.

Whyte, A.R., & Phillips, L.F. (1983a). Chem. Phys. Lett., 102, 451.

Whyte, A.R., & Phillips, L.F. (1983b). Bull. Soc. Chim. Belg., 92, 635.

Williams, B.A., & Fleming, J.W. (1994). Combust. Flame, 98, 93–106.

Williams, B.A., Papas, P. & Nelson, H.H. (1995). J. Phys. Chem., 99, 13471–13475.

Wolf, M., Yang, D.L. & Durant, J.L. (1994). J. Photochem. Photobiol. A Chem.,80, 85-93.

Wolff, T., & Wagner, H.Gg. (1988). Ber. Bunsenges. Phys. Chem., 92, 678–686.

Wolfrum, J. (1972). Chemie Ingenieur Technik, 44, 656.

Wooldridge, S.T., Hanson, R.K. & Bowman, C.T. (1995). Int. J. Chem. Kinet.,27, 1075–1087.

Wooldridge, M.S., Hanson, R.K. & Bowman, C.T. (1996a). Int. J. Chem. Kinet.,28, 361–372.

Page 206: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 331

Wooldridge, S.T., Hanson, R.K. & Bowman, C.T. (1996b). Int. J. Chem. Kinet.,28, 245–258.

Wooldridge, S.T., Mertens, J.D., Hanson, R.K. & Bowman, C.T. (1994). 25thSymposium (International) on Combustion, pp. 983–991.

Wray, K.L., & Teare, J.D. (1962). J. Chem. Phys., 36, 2582–2596.

Wu, C.H., Wang, H.T., Lin, M.C. & Fifer, R.A. (1990). J. Phys. Chem., 94,3344.

Xu, Z.-F., Fang, D.-C. & Fu, X.-Y. (1995). J. Phys. Chem., 99, 5889–5893.

Yamasaki, K., Okada, S., Koshi, M., & Matsui, H. (1991). J. Chem. Phys., 95,5087–5096.

Yang, H.–Y., Lissianski, V., Okoroanyanwu, J.U., Gardiner, W.C., & Shin, K.S.(1993). J. Phys. Chem., 97, 10042–10046.

Yarkony, D.R. (1993). J. Phys. Chem., 97, 4407–4412.

Yokoyama, K., Sakane, Y., & Fueno, T. (1991a). Bull. Chem. Soc. Jpn., 64,1738–1742.

Yokoyama, K., Takane, S., & Fueno, T. (1991b). Bull. Chem. Soc. Jpn., 64,2230–2235.

Yoshimine, M.J. (1989) J. Chem. Phys., 90, 378

Yumura, M., & Asaba, T. (1980). 18th Symposium (International)on Combustion,pp. 863–872.

Zabarnick, S., & Lin, M.C. (1989). Chem. Phys., 134, 185.

Zabarnick, S. (1993). Chem. Phys., 171, 265–273.

Zabielski, M.F., & Seery, D.J. (1985). Int. J. Chem. Kinet., 17, 1191–1199.

Zabielski, M.F., & Seery, D.J. (1989). In Fossil Fuels Combustion Symposium,Vol. 25, American Society of Mechanical Engineers, 91–94.

Zeldovich, Y.B. (1946). Acta Physicochim. URSS, 21, 577.

Zellner, R. (1984). Bimolecular Reaction Rate Coefficients. In CombustionChemistry, W.C. Gardiner, Jr. , Ed., Springer-Verlag, New York, 127–172.

Zuev, A.P., & Starikovskii, A.Y. (1991a). Khim. Fiz., 10, 52–63.

Zuev, A.P., & Starikovskii, A.Y. (1991b). Khim. Fiz., 10, 179–189.

Zuev, A P., & Starikovskii, A.Y. (1991c). Khim. Fiz., 10, 190–199.

Page 207: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

332 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.19(a) Summary of Rate Coefficient Parameters1 Part I.

Reaction A (Note 1) m Ea /R Notes

1 O + N2 −→N + NO 2.0×1014 38660

2 NO + M−→N + O + M 1.4×1015 747003 N2O + M−→N2 + O + M 5.7×1014 28230

4a N2O + O−→N2 + O2 1.4×1012 54404b N2O + O−→ 2NO 2.9×1013 11650

5 NH3 + M−→NH2 + H + M 3.6×1016 47200

6 NH3+ H−→NH2 + H2 5.4×105 2.40 49907 NH3+ OH−→NH2 + H2O 5.0×107 1.60 480

8 NH3+ O−→NH2 + OH 9.4×106 1.94 3250

9 NH2 + H−→NH + H2 4.8×108 1.50 399510 HO2 + NO−→NO2 + OH 2.2×1012 −240

11a N2O + H−→HNNO 1.2×1024 −4.46 5385 0.1 atm

11a 1.3×1025 −4.48 5420 1.0 atm

11a 3.2×1026 −4.58 5650 10 atm

11a N2O+H+M−→HNNO + M 1.1×1027 −3.48 5420 Note 2.

11b N2O + H−→N2 + OH 2.2×1014 8430 Note 3.

11c H + N2O−→NH + NO 8.5×1020 −1.62 17800

11d H + N2O−→NNH + O 2.4×1019 −1.26 2370011e NH + NO−→N2 + OH 1.4×1017 −1.49 660

11f NH + NO−→N2O + H 3.0×1018 −1.65 720 Note 4.

11g NH + NO−→NNH + O 1.7×1014 −0.20 614012a NH + O2 −→HNOO 3.5×1023 −5.00 1145 0.1 atm

12a 3.7×1024 −5.00 1155 1.0 atm

12a 5.4×1025 −5.05 1235 10 atm

12a NH+O2+M−→HNOO + M 3.0×1026 −4.00 1155 Note 5.

12d NH + O2 −→NO + OH 7.6×1010 770

12e NH + O2 −→H + NO2 2.3×1010 1250

12g NH + O2 −→HNO + O 4.6×105 2.00 3270

13b NH2 + O2 −→NH2O + O 2.5×1011 0.48 1489013d NH2 + O2 −→HNO + OH 6.2×107 1.23 17665

14b NH2 + HO2 −→NH2O + OH 2.5×1013

14d NH2 + HO2 −→NH3 + O2 9.2×105 1.94 −58015c NH2 + O−→HNO + H 4.6×1013 0

15d1 NH2 + O−→NH + OH 7.0×1012 015d2 NH2 + O−→NH + OH 3.3×108 1.50 2555

16a NH2 + OH−→NH2OH 1.8×1032 −6.91 2070 0.1 atm

16a 3.9×1033 −7.00 2235 1.0 atm

16a 5.6×1034 −7.02 2700 10 atm

16b NH2 + OH−→NH + H2O 2.4×106 2.00 25

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) P = 10 atm,T > 1000 K or P = 1 atm, T > 300 K. (3) For 1000 to 2000 K. (4) Reverse of reaction 11c.(5) Slight falloff at T = 300 K, P = 10 atm.

Page 208: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 333

TABLE 2.19(b) Summary of Rate Coefficient Parameters1 Part II.

Reaction A (Note 1) m Ea /R Notes

17a NH2 + NH2 −→N2H4 2.0×1046 −10.93 5030 Note 2.

17a 5.6×1048 −11.30 5980 Note 3.

17a 3.2×1049 −11.18 7040 Note 4.

17b NH2 + NH2−→H2NN + H2 2.4×1020 −2.91 1075 0.1 atm

17b 1.2×1021 −3.08 1695 1.0 atm

17b 2.3×1019 −2.54 2105 10 atm

17d NH2 + NH2 −→N2H3+ H 9.2×1011 −0.01 5040 0.1 atm

17d 1.2×1012 −0.03 5075 1.0 atm

17d 4.7×1012 −0.20 5345 10 atm

17e NH2 + NH2 −→NH3 + NH 5.0×1013 5000

18a NH2 + NO−→N2 + H2O 4.7×1012 −0.25 −605

18b NH2 + NO−→NNH + OH 3.5×1010 0.34 −38518c NH2 + NO−→NH2NO 1.9×1030 −6.67 1760 0.1 atm

18c 3.5×1031 −6.75 1875 1.0 atm

18c 1.7×1033 −6.92 2320 10 atm

19a CH3+NO(+N2)−→CH3NO (+N2) 1.0×1013 0 k∞1.9×1018 0 k0

SRI fallo� coe�cients: a=0.03, b={790, c=1.0 Note 5.

19a CH3+NO −→CH3NO 3.6×1035 −8.25 2420 0.1 atm

19a 1.0×1037 −8.38 2630 1.0 atm

19a 4.6×1041 −9.39 4160 10 atm

19d CH3 + NO−→H2CN + OH 2.2×109 0.75 590019e CH3 + NO−→HCN + H2O 4.9×108 0.46 6240

20a CH3 + N−→H2CN + H 6.1×1014 −0.31 145

20b CH3 + N−→HCN + H2 3.7×1012 0.15 −4520f CH3 + N−→HCNH + H 1.2×1011 0.52 −185

21a CH3 + NH2 −→CH3NH2 1.3×1054 −12.72 7855 Note 2.

21a 5.1×1052 −11.99 8450 Note 3.

21a 1.6×1047 −10.15 7895 Note 4.

21b CH3 + NH2 −→CH2NH2 + H 1.1×1013 −0.13 4985 0.1 atm

21b 1.4×1014 −0.43 5590 1.0 atm

21b 7.4×1012 6075 10 atm

21c CH3 + NH2 −→CH3NH + H 1.2×1013 −0.15 8125 0.1 atm

21c 4.4×1013 −0.31 8375 1.0 atm

21c 1.4×1014 −0.42 8990 10 atm

21d CH3 + NH2 −→H2C=NH + H2 2.1×1011 −0.10 9610 0.1 atm

21d 4.8×1011 −0.20 9765 1.0 atm

21d 2.9×1012 −0.40 10320 10 atm

21e CH3 + NH2 −→CH4 + NH 2.8×106 1.94 4635

21f CH3 + NH2 −→CH2 + NH3 1.6×106 1.87 3810

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) Valid for0.1 atm, 600 to 2500 K. (3) Valid for 1 atm, 600 to 2500 K. (4) Valid for 10 atm, 600 to2500 K. (5) Stewart et al. (1989). The k◦ expression of Kaiser (1993) has been increasedby a factor of 1.4 to account for the greater collision efficiency of N2. These expressionsare valid for the temperature range 300 to 500 K. The following expressions are suitablefor temperatures over 500 K. (See text Section 6.15.)

Page 209: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

334 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.19(c) Summary of Rate Coefficient Parameters1 Part III.

Reaction A (Note 1) m Ea /R Notes

22a CH2 + N2−→CH2NN 9.3×1030 −7.01 9935 0.1 atm

22a 1.6×1032 −7.07 10050 1.0 atm

22a 4.3×1033 −7.18 10500 10 atm

22d CH2 + N2−→HCN + NH 1.0×1013 37240

236 CH2 + NO−→Products 1.0×1013 −190 Note 2.

23g CH2 + NO−→HCNO + H 3.8×1013 −0.36 29023i CH2 + NO−→HCN + OH 2.9×1014 −0.69 380

23j CH2 + NO−→HNCO + H 3.1×1017 −1.38 64023k CH2 + NO−→NH2 + CO 2.3×1016 −1.43 670

23l CH2 + NO−→H2CN + O 8.1×107 1.42 207024a CH + N2−→HCNN 2.3×1027 −5.78 1230 0.1 atm

24a 3.6×1028 −5.84 1320 1.0 atm

24a 1.8×1030 −6.02 1735 10 atm

24c CH + N2−→HCN + N 4.4×1012 11060

25e CH + NO−→HCN + O 5.3×1013 025f CH + NO−→H + NCO 2.0×1013 025g CH + NO−→N + HCO 2.9×1013 025h CH + NO−→NH+CO 5.5×1012 025i CH + NO−→OH+CN 3.3×1012 026a N + O2 −→NO + O 9.0×109 1.00 327026b N + OH−→NH + O 6.4×1012 0.10 10700 Note 3.

26c N + OH−→NO + H 1.1×1014 56526d CH + N−→CN + H 1.7×1014 −0.09 026e CH2 + N−→HCN + H 5.0×1013 0

26f NH + N−→N2 + H 1.5×1013 026g NH2 + N−→N2 + 2H 7.1×1013 0

26h CN + N−→C + N2 2.4×1013 −28027a NH + NH−→N2 + 2H 5.1×1013 0

27b1 NH2 + NH−→N2H2 + H 1.5×1015 −0.50 0

27b2 NH2 + NH−→NH3 + N 9.2×105 1.94 1230

27c1 NH + OH−→HNO + H 2.0×1013 027c2 NH + OH−→N + H2O 1.2×106 2.00 −245

27d NH + H−→N + H2 3.5×1013 870

27e1 NH + O−→NO + H 6.0×1013 027e2 NH + O−→N + OH 1.7×108 1.50 169527f1 NH + CH3−→H2C=NH + H 4.0×1013 0

27f2 NH + CH3−→CH4 + N 8.2×105 1.87 2945

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) Total ratecoefficient for all channels of CH2 + NO. (3) Reverse of reaction 27e2.

Page 210: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 335

TABLE 2.19(d) Summary of Rate Coefficient Parameters1 Part IV.

Reaction A (Note 1) m Ea /R Notes

28a NNH−→N2 + H

3.0×108+ 1.0×1013

[M] 0.5 154028b1 NNH + O2 −→N2 + HO2 1.2×1012 −0.34 7528b2 NNH + O2 −→N2O + OH 2.9×1011 −0.34 75

28c NNH + H−→N2 + H2 2.4×108 1.50 −45028d1 NNH + OH−→N2 + H2O 2.4×1022 −2.88 1235

28d2 NNH + OH−→N2 + H2O 1.2×106 2.00 −60028e1 NNH + O−→N2 + OH 1.7×1016 −1.23 250

28e2 NNH + O−→N2 + OH 1.7×108 1.50 −450

28f NNH + NH2−→N2 + NH3 9.2×105 1.94 −58028g1 NNH + HO2−→N2 + H2O2 1.4×104 2.69 −805

28g2 NNH + HO2−→HNNO + OH 2.4×1013 855

28h NNH + NO−→N2 + HNO 1.2×106 2.00 −60029a1 N2H2 −→NNH + H 5.6×1036 −7.75 35355 Note 2.

29a1 1.8×1040 −8.41 36935 Note 3.

29a1 3.1×1041 −8.42 38270 Note 4.

29a2 N2H2 −→NNH + H 1.6×1037 −7.94 35610 Note 2.

29a2 2.6×1040 −8.53 36700 Note 3.

29a2 1.3×1044 −9.22 38790 Note 4.

29b1 N2H2 −→H2NN 9.2×1038 −9.01 34085 Note 2.

29b1 2.0×1041 −9.38 34450 Note 3.

29b1 1.3×1045 −10.13 35610 Note 4.

29c1 N2H2 + H−→NNH + H2 4.8×108 1.50 795

29c2 N2H2 + O−→NNH + OH 3.3×108 1.50 250

29c3 N2H2 + OH−→NNH + H2O 2.4×106 2.00 −60029c4 N2H2 + NH2 −→NH3 + NNH 1.8×106 1.94 −580

29c5 N2H2 + CH3 −→NNH + CH4 1.6×106 1.87 149529d N2H2 + NH−→NNH + NH2 2.4×106 2.00 −600

29e N2H2 + NO−→N2O + NH2 4.0×1012 600030a1 H2NN −→NNH + H 5.9×1032 −6.99 26065 Note 2.

30a1 9.6×1035 −5.57 27600 Note 3.

30a1 5.0×1036 −7.43 28835 Note 4.

30a2 H2NN −→NNH + H 7.2×1028 −7.77 25545 Note 2.

30a2 3.2×1031 −6.22 26330 Note 3.

30a2 5.1×1033 −6.52 27285 Note 4.

30b H2NN + O2−→NH2 + NO2 1.5×1012 300030c1 H2NN + H−→N2H2 + H 1.8×1010 0.97 2250

30c2 H2NN + H−→NNH + H2 4.8×108 1.50 −45030d1 H2NN + O−→NH2 + NO 3.2×109 1.03 1360

30d2 H2NN + O−→OH + NNH 3.3×108 1.50 −45030e1 H2NN + OH−→NH2NO + H 2.0×1012 0

30e2 H2NN + OH−→NNH + H2O 2.4×106 2.00 −600 Note 5.

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) Valid for0.1 atm, 600 to 2500 K. (3) Valid for 1 atm, 600 to 2500 K. (4) Valid for 10 atm, 600 to2500 K. (5) Reaction 30 entries continue in Part V.

Page 211: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

336 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.19(e) Summary of Rate Coefficient Parameters1 Part V.

Reaction A (Note 1) m Ea /R Notes

30f1 H2NN + CH3−→H2C=NNH2 + H 8.3×105 1.93 3270

30f2 H2NN + CH3−→CH3N=NH + H 8.3×105 1.93 327030f3 H2NN + CH3−→CH4 + NNH 1.6×106 1.87 65

30g1 H2NN + NH2−→HNNNH2 + H 7.9×106 1.90 −67030g2 H2NN + NH2−→NH3 + NNH 1.8×106 1.94 −580

30h1 H2NN + HO2−→NH2NO + OH 6.6×105 1.94 3550

30h2 H2NN + HO2−→NNH + H2O2 2.9×104 2.69 −80531a N2H3 −→N2H2 + H 2.3×1043 −9.55 32445 Note 2.

31a 3.6×1047 −10.38 34730 Note 3.

31a 1.8×1045 −9.39 35300 Note 4.

31b N2H3 + H−→N2H2 + H2 2.4×108 1.50 031c1 N2H3 + O−→NH2 + HNO 3.0×1013 0

31c2 N2H3 + O−→NH2NO + H 3.0×1013 031c3 N2H3 + O−→N2H2 + OH 1.7×108 1.50 −325

31d1 N2H3 + OH−→N2H2 + H2O 1.2×106 2.00 −600

31d2 N2H3 + OH−→H2NN + H2O 3.0×1013 0

31e1 N2H3 + CH3 −→N2H2 + CH4 8.2×105 1.87 915

31e2 N2H3 + CH3 −→H2NN + CH4 3.0×1013 0

31f1 N2H3 + NH2 −→N2H2 + NH3 9.2×105 1.94 −58031f2 N2H3 + NH2 −→H2NN + NH3 3.0×1013 0

31g1 N2H3 + HO2 −→NH2NHO + OH 3.0×1013 0

31g2 N2H3 + HO2 −→N2H2 + H2O2 2.9×104 2.69 −805

31g3 N2H3 + HO2 −→N2H4 + O2 9.2×105 1.94 1070

32 N2H4 −→H2NN + H2 4.0×1044 −9.85 35910 Note 2.

32 5.3×1039 −8.35 34880 Note 3.

32 2.5×1039 −8.19 35060 Note 4.

32a N2H4 + H−→N2H3 + H2 9.6×108 1.50 2435

32b N2H4 + O−→N2H3 + OH 6.7×108 1.50 143532c N2H4 + OH−→N2H3 + H2O 4.8×106 2.00 −325

32d N2H4 + CH3 −→N2H3 + CH4 3.3×106 1.87 268032e N2H4 + NH2 −→N2H3 + NH3 3.7×106 1.94 820

33a1 NO + C−→CO + N 1.7×1013 033a2 NO + C−→CN + O 1.1×1013 033c1 NO + HCCO−→HCNO + CO 4.6×1013 35033c2 NO + HCCO−→HCN + CO2 1.4×1013 350

34a NO2 + H−→NO + OH 1.3×1014 180

34b NO2 + O−→NO + O2 3.9×1012 −120

34c NO2 + M−→NO + O + M 5.7×1015 3019034d1 NO2 + NH2−→N2O + H2O 1.5×1016 −1.44 135

34d2 NO2 + NH2−→NH2O + NO 6.6×1016 −1.44 13534e NO2 + CH3−→NO + CH3O 1.4×1013 0

35 N2O + OH−→N2 + HO2 1.3×10−2 4.72 18400

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) Valid for0.1 atm, 600 to 2500 K. (3) Valid for 1 atm, 600 to 2500 K. (4) Valid for 10 atm, 600 to2500 K.

Page 212: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 337

TABLE 2.19(f) Summary of Rate Coefficient Parameters1 Part VI.

Reaction A (Note 1) m Ea /R Notes

36a HNO + M−→H + NO + M 2.6×1016 2450036b 2HNO−→N2O + H2O 8.5×108 1550 1.0 atm

36c HNO + OH−→NO + H2O 1.3×107 1.88 −48036d1 HNO + H−→H2 + NO 4.5×1011 0.72 330

36d2 HNO + H−→O + NH2 3.5×1015 −0.30 14730 Note 2.

36e HNO + O−→OH + NO 4.5×1011 0.72 33036f HNO + NH2 −→NH3 + NO 9.2×105 1.94 −580

36g HNO + NO−→N2O + OH 8.5×1012 1489036h HNO + O2 −→NO + HO2 2.0×1013 8000

36i HNO + CH3 −→NO + CH4 8.2×105 1.87 480

37a NH2O + M−→HNO + H + M 2.8×1024 −2.83 32670

37b NH2O−→HNOH 8.2×1025 −4.94 22040 0.1 atm

37b 1.3×1027 −4.99 22135 1.0 atm

37b 2.6×1028 −5.06 22530 10 atm

37b NH2O + M−→HNOH + M 1.1×1029 −3.99 22135 Note 3.

37c1 NH2O + H−→NH2 + OH 4.0×1013 0

37c2 NH2O + H−→HNO + H2 4.8×108 1.50 78537d NH2O + O−→HNO + OH 3.3×108 1.50 245

37e NH2O + OH−→HNO + H2O 2.4×106 2.00 −60037f1 NH2O + CH3 −→CH3O + NH2 2.0×1013 0

37f2 NH2O + CH3 −→CH4+ HNO 1.6×106 1.87 1490

37g NH2O + NH2 −→HNO + NH3 1.8×106 1.94 −58037h1 NH2O + HO2 −→HNO + H2O2 2.9×104 2.69 −805

37h2 NH2O + HO2 −→O2 + NH2OH 2.9×104 2.69 −80538a HNOH + M−→H + HNO + M 2.0×1024 −2.84 2966038b1 HNOH + H−→NH2 + OH 4.0×1013 0

38b2 HNOH + H−→HNO + H2 4.8×108 1.50 19038c1 HNOH + O−→HNO + OH 7.0×1013 038c2 HNOH + O−→HNO + OH 3.3×108 1.50 −18038d HNOH + OH−→HNO + H2O 2.4×106 2.00 −600

38e1 HNOH + CH3 −→CH3N.H + OH 2.0×1013 038e2 HNOH + CH3 −→CH4 + HNO 1.6×106 1.87 1055

38f1 HNOH + NH2 −→N2H3 + OH 6.7×106 1.82 36038f2 HNOH + NH2 −→H2NN + H2O 4.6×1019 −1.94 970

38f3 HNOH + NH2 −→HNO + NH3 1.8×106 1.94 −58038g1 HNOH + HO2 −→HONHO+ OH 4.0×1013 0

38g2 HNOH + HO2 −→HNO + H2O2 2.9×104 2.69 −805

38g3 HNOH + HO2 −→NH2OH + O2 2.9×104 2.69 −805

391HNOO + M−→OH + NO + M 1.5×1036 −6.18 15670

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) Reverse ofreaction 15c. (3) T > 1000 K.

Page 213: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

338 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.19(g) Summary of Rate Coefficient Parameters1 Part VII.

Reaction A (Note 1) m Ea /R Notes

40a HONO + M−→OH + NO + M 2.0×1031 −4.56 2575540b1 HONO + H−→H2 + NO2 2.0×108 1.55 3330

40b2 HONO + H−→H2O + NO 8.1×106 1.89 193540b3 HONO + H−→OH + HNO 5.6×1010 0.86 250040c HONO + O−→OH + NO2 1.7×108 1.50 1525

40d HONO + OH−→H2O + NO2 1.2×106 2.00 −300

40e HONO + CH3 −→NO2 + CH4 8.1×105 1.87 2770

40f HONO + NH2 −→NO2 + NH3 9.2×105 1.94 96541 HNO2 −→HONO 7.1×1027 −5.40 26440 0.1 atm

41 1.3×1029 −5.47 26580 1.0 atm

41 2.0×1030 −5.50 27020 10 atm

41a HNO2 + H−→H2 + NO2 2.4×108 1.50 2095

41b HNO2 + O−→OH + NO2 1.7×108 1.50 119041c HNO2 + OH−→H2O + NO2 1.2×106 2.00 −400

41d HNO2 + CH3 −→NO2 + CH4 8.1×105 1.87 2435

41e HNO2 + NH2 −→NO2 + NH3 9.2×105 1.94 440

42 HCN−→HNC 1.5×1023 −4.20 24890 0.1 atm

42 1.9×1024 −4.23 24950 1.0 atm

42 5.3×1025 −4.34 25260 10 atm

42 HCN+ M−→HNC+ M 1.6×1026 −3.23 24950 Note 2.

42a HCN + OH−→CN + H2O 3.9×106 1.83 518042b1 OH + HCN−→HNCO + H 4.4×103 2.26 322042b2 OH + HCN−→HOCN + H 1.1×106 2.03 673042b3 OH + HCN−→NH2 + CO 1.6×102 2.56 453042b4 OH + HCN−→ .N=CHOH 1.7×1029 −6.31 2580 0.1 atm

42b4 2.8×1030 −6.37 2690 1.0 atm

42b4 1.1×1032 −6.53 3140 10 atm

42c1 HCN + O−→NH + CO 5.4×108 1.21 377042c2 HCN + O−→NCO + H 2.0×108 1.47 382042c3 HCN + O−→CN + OH 4.2×1010 0.40 1040543a O + HNC−→NH + CO 4.6×1012 110043b OH + HNC−→HNCO + H 2.8×1013 186043c1 HNC + O2 −→HNCO + O 1.5×1012 0.01 2070

43c2 HNC + O2 −→NH + CO2 1.6×1019 −2.25 89544a CN + H2 −→HCN + H 3.6×108 1.55 1510

44b CN + H2O−→HCN + OH 7.8×1012 3750 Note 3.

44c CN + O−→CO + N 7.7×1013 044d CN + O2 −→NCO + O 1.0×1013 0

44e CN + OH−→NCO + H 4.0×1013 044f CN + HCN−→NCCN + H 1.5×107 1.71 77044g CN+N2O−→NCN + NO 4.2×1011 3610

44h CN + NO2 −→NCO + NO 6.2×1015 −0.75 175

44i CN + CH4 −→HCN + CH3 1.2×105 2.64 −80

44j CN + NH3 −→HCN + NH2 9.2×1012 −180

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) T > 1000K. (3) Reverse of reaction 42a.

Page 214: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 339

TABLE 2.19(h) Summary of Rate Coefficient Parameters1 Part VIII.

Reaction A (Note 1) m Ea /R Notes

45a H2CN−→HCN + H 1.3×1029 −6.03 15045 0.1 atm

45a 6.0×1031 −6.46 16160 1.0 atm

45a 3.5×1029 −5.46 16380 10 atm

45b H2CN + HO2 −→CH2NO + OH 3.0×1013 0

45b1 H2CN + HO2 −→HCN + H2O2 1.4×104 2.69 −810

45b2 H2CN + HO2 −→H2CNH + O2 1.4×104 2.69 −81045c H2CN + O2 −→CH2O + NO 3.0×1012 3000

45d H2CN + CH3 −→HCN + CH4 8.1×105 1.87 −560

45e H2CN + OH−→HCN + H2O 2.1×1017 −1.68 160 0.1 atm

45e 1.5×1019 −2.18 1090 1.0 atm

45e 9.5×1021 −2.91 2835 10 atm

45e2 H2CN + OH−→HCN + H2O 1.2×106 2.00 −60045f H2CN + N−→N2 + CH2 6.0×1013 200

45g H2CN + H−→HCN + H2 2.4×108 1.50 −450

45h H2CN + NH2 −→HCN + NH3 9.2×105 1.94 −58045i1 H2CN + O −→HCN + OH 1.7×108 1.50 −450

45i2 H2CN + O −→HNCO + H 6.0×1013 045i3 H2CN + O −→HCNO + H 2.0×1013 0

46 HCNH−→HCN + H 7.7×1025 −5.20 11065 0.1 atm

46 6.1×1028 −5.69 12215 1.0 atm

46 6.2×1026 −4.77 12490 10 atm

46a1 HCNH + H−→H2CN + H 2.0×1013 0

46a2 HCNH + H−→HCN + H2 2.4×108 1.50 −450

46b1 HCNH + O−→HNCO + H 7.0×1013 046b2 HCNH + O−→HCN + OH 1.7×108 1.50 −45046c HCNH + OH−→HCN + H2O 1.2×106 2.00 −600

46d HCNH + CH3−→HCN + CH4 8.2×105 1.87 −560

47a HCNN + O2 −→H + CO2 + N2 4.0×1012 047b HCNN + O2 −→HCO + N2O 4.0×1012 0

48a1 H2C=NH + H−→H2CN + H2 2.4×108 1.50 3685

48a2 H2C=NH + O−→H2CN + OH 1.7×108 1.50 233048a3 H2C=NH + OH−→H2CN + H2O 1.2×106 2.00 −45

48a4 H2C=NH + CH3 −→H2CN + CH4 8.2×105 1.87 3585

48a5 H2C=NH + NH2 −→H2CN + NH3 9.2×105 1.94 223548b1 H2C=NH + H−→HCNH + H2 3.0×108 1.50 3085

48b2 H2C=NH + O−→HCNH + OH 2.2×108 1.50 2720

48b3 H2C=NH + OH−→HCNH + H2O 2.4×106 2.00 230

48b4 H2C=NH + CH3 −→HCNH + CH4 5.3×105 1.87 4875

48b5 H2C=NH + NH2 −→HCNH + NH3 1.8×106 1.94 3065

48c H2C=NH + O−→CH2O+NH 1.7×106 2.08 0

All pressures refer to N2 buffer gas. Note: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s.

Page 215: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

340 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.19(i) Summary of Rate Coefficient Parameters1 Part IX.

Reaction A (Note 1) m Ea /R Notes

49a CH3NH−→H2C=NH + H 1.6×1036 −7.92 18290 Note 2.

49a 1.3×1042 −9.24 20805 Note 3.

49a 2.3×1044 −9.51 22770 Note 4.

49b CH3NH + H−→H2C=NH + H2 7.2×108 1.50 −450

49c CH3NH + O−→H2C=NH + OH 5.0×108 1.50 −450

49d CH3NH + OH−→H2C=NH + H2O 3.6×106 2.00 −60049e CH3NH + CH3−→H2C=NH + CH4 2.4×106 1.87 −560

50 CH2NH2 −→H2C=NH + H 1.1×1045 −10.24 24065 Note 2.

50 2.4×1048 −10.82 26190 Note 3.

50 3.2×1046 −9.95 26940 Note 4.

50a1 CH2NH2 + O2 −→H2C=NH + HO2 1.0×1022 −3.09 3400

50a2 CH2NH2 + O2 −→NH2 + CH2O + O 6.0×1018 −1.59 1519550b CH2NH2 + H−→H2C=NH + H2 4.0×108 1.50 −450

50c1 CH2NH2 + O−→CH2O + NH2 7.0×1013 050c2 CH2NH2 + O−→H2C=NH + OH 3.3×108 1.50 −450

50d1 CH2NH2 + OH−→CH2OH + NH2 4.0×1013 050d2 CH2NH2 + OH−→H2C=NH + H2O 2.4×106 2.00 −600

50e1 CH2NH2 + CH3 −→C2H5 + NH2 2.0×1013 1360

50e2 CH2NH2 + CH3 −→H2C=NH + CH4 1.6×106 1.87 −31551a1 CH3NH2 + H−→CH2NH2 + H2 5.6×108 1.50 2750

51a2 CH3NH2 + O−→CH2NH2 + OH 4.0×108 1.50 261551a3 CH3NH2 + OH−→CH2NH2 + H2O 3.6×106 2.00 120

51a4 CH3NH2 + CH3 −→CH2NH2 + CH4 1.5×106 1.87 461551a5 CH3NH2 + NH2 −→CH2NH2 + NH3 2.8×106 1.94 2765

51b1 CH3NH2 + H−→CH3NH + H2 4.8×108 1.50 4885

51b2 CH3NH2 + O−→CH3NH + OH 3.3×108 1.50 319551b3 CH3NH2 + OH−→CH3NH + H2O 2.4×106 2.00 225

51b4 CH3NH2 + CH3 −→CH3NH + CH4 1.6×106 1.87 445051b5 CH3NH2 + NH2 −→CH3NH + NH3 1.8×106 1.94 3595

52a NCCN + M−→CN + CN + M 1.6×1034 −4.32 6546552b NCCN + H−→HCN + CN 1.4×1014 4000 Note 5.

52c NCCN + O−→NCO + CN 4.6×1012 447052d NCCN + OH−→HOCN + CN 2.0×1012 956053a6 NCO + NO −→Products 1.4×1018 −1.73 385 Note 6.

53a1 NCO + NO−→CO2 + N2 7.8×1017 −1.73 385

53a2 NCO + NO−→N2O + CO 6.2×1017 −1.73 38553b NCO + M−→N + CO + M 3.3×1014 2720053c NCO + H2 −→HNCO + H 7.6×102 3.00 2000

53d1 NCO + O−→NO + CO 4.2×1013 053d2 NCO + O−→N + CO2 8.0×1012 126053e NCO + H−→NH + CO 5.2×1013 053f NCO + N−→N2 + CO 3.3×1013 0 Note 7.

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) Valid for0.1 atm, 600 to 2500 K. (3) Valid for 1 atm, 600 to 2500 K. (4) Valid for 10 atm, 600 to2500 K. (5) Reverse of reaction 44f. (6) Total for all channels of reaction 53a. (7) Reaction53 entries continue in Part X.

Page 216: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 341

TABLE 2.19(j) Summary of Rate Coefficient Parameters1 Part X.

Reaction A (Note 1) m Ea /R Notes

53g1 NCO + OH−→HNCO + O 7.8×104 2.27 −50053g2 NCO + OH−→HON + CO 5.3×1012 −0.07 258053g3 NCO + OH−→H + CO + NO 8.3×1012 −0.05 908053h1 NCO + NO2 −→CO2 + N2O 2.3×1012 −440

53h2 NCO + NO2 −→CO + 2NO 2.1×1011 −44053i NCO + CH4 −→HNCO + CH3 9.8×1012 4090

53j NCO + NH3 −→HNCO + NH2 2.8×104 2.48 495

54 HCNO−→HCN + O 2.0×1030 −6.03 30565 0.1 atm

54 4.2×1031 −6.12 30805 1.0 atm

54 5.9×1031 −5.85 31170 10 atm

54a HCNO + H−→HNCO + H 2.1×1015 −0.69 143554b HCNO + H−→HCN + OH 2.7×1011 0.18 106554c HCNO + H−→NH2 + CO 1.7×1014 −0.75 145554d HCNO + H−→HOCN + H 1.4×1011 −0.19 125054e HCNO + O−→HCO + NO 7.0×1013 054f HCNO + OH−→HCOH + NO 4.0×1013 055a HOCN + H−→HCN + OH 2.0×1013 −0.04 1075 Note 2.

55b HOCN + H−→HNCO + H 3.1×108 0.84 96555c HOCN + H−→NH2 + CO 1.2×108 0.61 104555d HOCN + H−→H2 + NCO 2.4×108 1.50 3330

55e HOCN + O−→OH + NCO 1.7×108 1.50 208055f HOCN + OH−→H2O + NCO 1.2×106 2.00 −125

55g HOCN + CH3 −→CH4 + NCO 8.2×105 1.87 3330

55h HOCN + NH2 −→NCO + NH3 9.2×105 1.94 183556a HNCO + M−→NH + CO + M 1.3×1016 4246056b HNCO + H−→NH2 + CO 3.6×104 2.49 1180

56c1 HNCO + O−→HNO + CO 1.7×106 2.08 056c2 HNCO + O−→NH + CO2 1.7×106 2.08 056d1 HNCO + OH−→NH2 + CO2 6.3×1010 −0.06 5860

56d2 HNCO + OH−→NCO + H2O 5.2×1010 −0.03 8840

56e HNCO + H−→NCO + H2 1.8×105 2.40 4990 Note 3.

56f HNCO + O−→NCO + OH 3.1×106 1.94 3250 Note 4.

56g HNCO + OH−→NCO + H2O 3.6×107 1.50 1810

56h HNCO + CH3 −→NCO + CH4 1.0×1012 5000 Note 5.

56i HNCO + NH2 −→NCO + NH3 1.0×1012 4500 Note 6.

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s. (2) Reverse ofreaction 42b2. (3) Reverse of reaction 53c. (4) Reverse of reaction 53g1. (5) Reverse ofreaction 53i. (6) Reverse of reaction 53j.

Page 217: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

342 Anthony M. Dean, Joseph W. Bozzelli

TABLE 2.19(k) Summary of Rate Coefficient Parameters1 Part XI.

Reaction A (Note 1) m Ea /R Notes

57a CH2NO−→HNCO + H 6.9×1041 −9.30 26020 Note 2.

57a 2.3×1042 −9.11 27095 Note 3.

57a 1.7×1038 −7.64 26965 Note 4.

57b CH2NO + O2 −→CH2O+ NO2 1.2×1015 −1.01 10130 Note 5.

57b CH2NO + O2 −→CH2O+ NO2 2.9×1012 −0.31 8910 Note 6.

57c1 CH2NO + H−→CH3 + NO 4.0×1013 057c2 CH2NO + H−→HCNO + H2 4.8×108 1.50 −450

57d1 CH2NO + O−→CH2O + NO 7.0×1013 0

57d2 CH2NO + O−→HCNO + OH 3.3×108 1.50 −45057e1 CH2NO + OH−→CH2OH + NO 4.0×1013 0

57e2 CH2NO + OH−→HCNO + H2O 2.4×106 2.00 −60057f1 CH2NO + CH3 −→C2H5+ NO 3.0×1013 0

57f2 CH2NO + CH3 −→HCNO + CH4 1.6×106 1.87 −56057g1 CH2NO + NH2 −→CH2NH2 + NO 3.0×1013 0

57g2 CH2NO + NH2 −→HCNO + NH3 1.8×106 1.94 −580

58a CH3NO + H−→CH2NO + H2 4.4×108 1.50 19058b CH3NO + O−→CH2NO + OH 3.3×108 1.50 1820

58c CH3NO + OH−→CH2NO + H2O 3.6×106 2.00 −600

58d CH3NO + CH3 −→CH2NO + CH4 7.9×105 1.87 272558e CH3NO + NH2 −→CH2NO + NH3 2.8×106 1.94 540

58f CH3NO + H−→CH3 + HNO 1.8×1013 140058g CH3NO + O−→CH3 + NO2 1.7×106 2.08 0

58h CH3NO + OH−→CH3 + HONO 2.5×1012 50059a HON + M−→NO + H + M 5.1×1019 −1.73 807559b1 HON + H−→HNO + H 2.0×1013 059b2 HON + H−→OH + NH 2.0×1013 059c HON + O−→OH + NO 7.0×1013 059d HON + OH−→HONO + H 4.0×1013 059e HON + O2 −→HONO + O 1.0×1012 250060 HCOH−→CH2O 3.5×1017 −2.86 4470 0.1 atm

60 2.1×1019 −3.07 4800 1.0 atm

60 1.8×1021 −3.32 5465 10 atm

Notes: All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressedin the form k = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s.(2) Valid for 0.1 atm, 600 to 2500 K. (3) Valid for 1 atm, 600 to 2500 K. (4) Valid for 10atm, 600 to 2500 K. (5) Valid for 1000 to 2500 K. (6) Valid for 300 to 1000 K.

Page 218: 2 Combustion Chemistry of Nitrogen · Combustion Chemistry of Nitrogen Anthony M. Dean1 Joseph W. Bozzelli1,2 2.1 INTRODUCTION The purpose of this chapter is to examine reactions

2. Combustion Chemistry of Nitrogen 343

TABLE 2.19(l) Summary of Rate Coefficient Parameters1 Part XII.

Reaction A (Note 1) m Ea /R Notes

61b1 NH2OH + H−→HNOH + H2 4.8×108 1.50 3145

61b2 NH2OH + H−→NH2O + H2 2.4×108 1.50 2550

61c1 NH2OH + O−→HNOH + OH 3.3×108 1.50 194561c2 NH2OH + O−→NH2O + OH 1.7×108 1.50 1515

61d1 NH2OH + OH−→HNOH + H2O 2.4×106 2.00 −16561d2 NH2OH + OH−→NH2O + H2O 1.2×106 2.00 −300

61e1 NH2OH + CH3−→HNOH + CH4 1.6×106 1.87 3195

61e2 NH2OH + CH3−→NH2O + CH4 8.2×105 1.87 276561f1 NH2OH + NH2−→HNOH + NH3 1.8×106 1.94 1625

61f2 NH2OH + NH2−→NH2O + NH3 9.2×105 1.94 950

61g1 NH2OH + HO2−→HNOH + H2O2 2.9×104 2.69 481061g2 NH2OH + HO2−→NH2O + H2O2 1.4×104 2.69 3230

62a NH2NO−→N2 + H2O 4.1×1033 −7.18 17700 0.1 atm

62a 3.1×1034 −7.11 18260 1.0 atm

62a 2.9×1031 −5.91 18205 10 atm

62b NH2NO + H−→HNNO + H2 4.8×108 1.50 373062c NH2NO + O−→HNNO + OH 3.3×108 1.50 2365

62d NH2NO + OH−→HNNO + H2O 2.4×106 2.00 −35

62e NH2NO + CH3−→HNNO + CH4 1.6×106 1.87 361562f NH2NO + NH2−→HNNO + NH3 1.8×106 1.94 2285

62g NH2NO + HO2−→HNNO + H2O2 2.9×104 2.69 635563a H2NNHO−→NH2 + HNO 2.7×1039 −8.74 20945 0.1 atm

63a 2.4×1040 −8.73 20940 1.0 atm

63a 1.2×1041 −8.64 20925 10 atm

63b H2NNHO + H−→HNNHO + H2 4.8×108 1.50 −450

63c H2NNHO + O−→HNNHO + OH 3.3×108 1.50 −45063d H2NNHO + OH−→HNNHO + H2O 2.4×106 2.00 −600

63e H2NNHO + CH3−→HNNHO + CH4 1.6×106 1.87 190

63f H2NNHO + NH2−→HNNHO + NH3 1.8×106 1.94 −58063g H2NNHO + HO2−→HNNHO + H2O2 2.9×104 2.69 −805

64a ClNO + M−→NO + Cl + M 1.9×1015 1610064b O + ClNO−→ClO + NO 5.0×1012 152064c1 OH + ClNO−→HOCl + NO 5.4×1012 113064c2 OH + ClNO−→HONO + Cl 5.5×1010 −24064d Cl + ClNO−→Cl2 + NO 4.0×1013 −13064e H + ClNO−→HCl + NO 4.6×1013 455

All pressures refer to N2 buffer gas. Notes: (1) Rate coefficients are expressed in the formk = A T m exp(−Ea/RT ) for concentration units mol/cm3 and time in s.


Recommended