Combustion modeling of mono-carbon fuels using the rate-controlled constrained-equilibrium method

15
Combustion modeling of mono-carbon fuels using the rate-controlled constrained-equilibrium method Mohammad Janbozorgi a , Sergio Ugarte a , Hameed Metghalchi a, * , James. C. Keck b a Mechanical and Industrial Engineering Department, Northeastern University, Boston, MA 02115, USA b Mechanical Engineering Department, Massachusetts Institute of Technology, Cambridge, MA 02139, USA article info Article history: Received 15 September 2008 Received in revised form 2 February 2009 Accepted 25 May 2009 Available online 5 August 2009 Keywords: Rate-controlled constrained-equilibrium Maximum entropy principle Detailed modeling Reduced kinetics Methane Methanol Formaldehyde abstract The rate-controlled constrained-equilibrium (RCCE) method for simplifying the kinetics of complex react- ing systems is reviewed. This method is based on the maximum entropy principle of thermodynamics and involves the assumption that the evolution of a system can be described using a relatively small set of slowly changing constraints imposed by the external and internal dynamics of the system. As a result, the number of differential and algebraic equations required to determine the constrained-equilib- rium state of a system can be very much smaller than the number of species in the system. It follows that only reactions which change constraints are required to determine the dynamic evolution of the system and all other reactions are in equilibrium. The accuracy of the method depends on both the character and number of constraints employed and issues involved in the selection and transformation of the con- straints are discussed. A method for determining the initial conditions for highly non-equilibrium sys- tems is also presented. The method is illustrated by applying it to the oxidation of methane (CH 4 ), methanol (CH 3 OH), and formaldehyde (CH 2 O) in a constant volume adiabatic chamber over a wide range of initial temperatures, pressures, and equivalence ratios. The RCCE calculations were carried out using 8–12 constraints and 133 reactions. Good agreement with ‘‘Detailed Kinetic Model” (DMK) calculations using 29 species and 133 reactions was obtained. The number of reactions in the RCCE calculations could be reduced to 20 for CH 4 , 16 for CH 3 OH, and 12 for CH 2 O without changing the results significantly affecting the agreement. It may be noted that a DKM with 29 species requires a minimum of 29 reactions. Ó 2009 The Combustion Institute. Published by Elsevier Inc. All rights reserved. 1. Introduction The development of models for describing the dynamic evolution of chemically reacting systems is a fundamental objective of chemical kinetics. The conventional approach to this problem involves first specifying the state and species variables to be included in the model, compiling a ‘‘full set” of rate-equations for these variables, and integrating this set of equations to obtain the time-dependent behavior of the system. Such models are frequently referred to as ‘‘Detailed Kinetic Model”s (DKMs). The problem is that the detailed kinetics of C/H/O/N molecules can easily involve hundreds of chemical species and isomers, and thousands of possible reactions even for system containing only C 1 molecules. Clearly, the computational effort required to treat such systems is extremely large. The difficulties are compounded when considering reacting turbulent flows, where the complexity of turbulence is added to that of the chemistry. As a result a great deal of effort has been devoted to developing methods for reducing the size of DMKs. Among the most promi- nent are: Quasi Steady State Approximation (QSSA) [1], Partial Equilibrium Approximation [2], Intrinsic Low Dimensional Manifolds (ILDM) [3], Computational Singular Perturbation (CSP) [4], Adaptive Chemistry [5], Directed Relation Graph (DRG) [6], and the ICE-PIC method [7]. An alternative approach, originally proposed by Keck and Gillespie [8] and later developed and applied by Keck and co-work- ers [9–11], and others [12–15] is the rate-controlled constrained- equilibrium (RCCE) method. This method is based on the maximum entropy principle of thermodynamics and involves the fundamental assumption that slow reactions in a complex reacting system impose constraints on its composition which control the rate at which it relaxes to chemical-equilibrium, while the fast reactions equilibrate the system subject to the constraints imposed by the slow reactions. Consequently, the system relaxes to chemi- cal-equilibrium through a sequence of constrained-equilibrium states at a rate controlled by the slowly changing constraints. An advantage of the RCCE method is that it does not require a large DKM as a starting point. Instead, one starts with a short list 0010-2180/$ - see front matter Ó 2009 The Combustion Institute. Published by Elsevier Inc. All rights reserved. doi:10.1016/j.combustflame.2009.05.013 * Corresponding author. Fax: +1 617 373 2921. E-mail address: [email protected] (H. Metghalchi). Combustion and Flame 156 (2009) 1871–1885 Contents lists available at ScienceDirect Combustion and Flame journal homepage: www.elsevier.com/locate/combustflame

Transcript of Combustion modeling of mono-carbon fuels using the rate-controlled constrained-equilibrium method

Combustion and Flame 156 (2009) 1871–1885

Contents lists available at ScienceDirect

Combustion and Flame

journal homepage: www.elsevier .com/locate /combustflame

Combustion modeling of mono-carbon fuels using the rate-controlledconstrained-equilibrium method

Mohammad Janbozorgi a, Sergio Ugarte a, Hameed Metghalchi a,*, James. C. Keck b

a Mechanical and Industrial Engineering Department, Northeastern University, Boston, MA 02115, USAb Mechanical Engineering Department, Massachusetts Institute of Technology, Cambridge, MA 02139, USA

a r t i c l e i n f o

Article history:Received 15 September 2008Received in revised form 2 February 2009Accepted 25 May 2009Available online 5 August 2009

Keywords:Rate-controlled constrained-equilibriumMaximum entropy principleDetailed modelingReduced kineticsMethaneMethanolFormaldehyde

0010-2180/$ - see front matter � 2009 The Combustdoi:10.1016/j.combustflame.2009.05.013

* Corresponding author. Fax: +1 617 373 2921.E-mail address: [email protected] (H. Metg

a b s t r a c t

The rate-controlled constrained-equilibrium (RCCE) method for simplifying the kinetics of complex react-ing systems is reviewed. This method is based on the maximum entropy principle of thermodynamicsand involves the assumption that the evolution of a system can be described using a relatively smallset of slowly changing constraints imposed by the external and internal dynamics of the system. As aresult, the number of differential and algebraic equations required to determine the constrained-equilib-rium state of a system can be very much smaller than the number of species in the system. It follows thatonly reactions which change constraints are required to determine the dynamic evolution of the systemand all other reactions are in equilibrium. The accuracy of the method depends on both the character andnumber of constraints employed and issues involved in the selection and transformation of the con-straints are discussed. A method for determining the initial conditions for highly non-equilibrium sys-tems is also presented.

The method is illustrated by applying it to the oxidation of methane (CH4), methanol (CH3OH), andformaldehyde (CH2O) in a constant volume adiabatic chamber over a wide range of initial temperatures,pressures, and equivalence ratios. The RCCE calculations were carried out using 8–12 constraints and 133reactions. Good agreement with ‘‘Detailed Kinetic Model” (DMK) calculations using 29 species and 133reactions was obtained. The number of reactions in the RCCE calculations could be reduced to 20 forCH4, 16 for CH3OH, and 12 for CH2O without changing the results significantly affecting the agreement.It may be noted that a DKM with 29 species requires a minimum of 29 reactions.

� 2009 The Combustion Institute. Published by Elsevier Inc. All rights reserved.

1. Introduction

The development of models for describing the dynamicevolution of chemically reacting systems is a fundamentalobjective of chemical kinetics. The conventional approach to thisproblem involves first specifying the state and species variablesto be included in the model, compiling a ‘‘full set” of rate-equationsfor these variables, and integrating this set of equations to obtainthe time-dependent behavior of the system. Such models arefrequently referred to as ‘‘Detailed Kinetic Model”s (DKMs).

The problem is that the detailed kinetics of C/H/O/N moleculescan easily involve hundreds of chemical species and isomers, andthousands of possible reactions even for system containing onlyC1 molecules. Clearly, the computational effort required to treatsuch systems is extremely large. The difficulties are compoundedwhen considering reacting turbulent flows, where the complexityof turbulence is added to that of the chemistry.

ion Institute. Published by Elsevier

halchi).

As a result a great deal of effort has been devoted to developingmethods for reducing the size of DMKs. Among the most promi-nent are: Quasi Steady State Approximation (QSSA) [1], PartialEquilibrium Approximation [2], Intrinsic Low DimensionalManifolds (ILDM) [3], Computational Singular Perturbation (CSP)[4], Adaptive Chemistry [5], Directed Relation Graph (DRG) [6],and the ICE-PIC method [7].

An alternative approach, originally proposed by Keck andGillespie [8] and later developed and applied by Keck and co-work-ers [9–11], and others [12–15] is the rate-controlled constrained-equilibrium (RCCE) method. This method is based on themaximum entropy principle of thermodynamics and involves thefundamental assumption that slow reactions in a complex reactingsystem impose constraints on its composition which control therate at which it relaxes to chemical-equilibrium, while the fastreactions equilibrate the system subject to the constraints imposedby the slow reactions. Consequently, the system relaxes to chemi-cal-equilibrium through a sequence of constrained-equilibriumstates at a rate controlled by the slowly changing constraints.

An advantage of the RCCE method is that it does not require alarge DKM as a starting point. Instead, one starts with a short list

Inc. All rights reserved.

1872 M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885

of constraints and rate-limiting reactions, to which more can beadded to improve the accuracy to the desired level. If the only con-straints are those imposed by slowly changing state variables, theRCCE method is equivalent to a local-chemical-equilibrium (LTE)calculation. If the number of constraints in an RCCE model is iden-tical to the number of species in a DKM model, then the total num-ber of equations, differential plus algebraic, to be solved will beidentical and the results will be similar but not identical to thoseof the DKM model. The DKM will include only the specified specieswhereas the RCCE model will include implicitly all species whichcan be made from the elements in the system.

As with all thermodynamic systems, the number of constraintsnecessary to describe the dynamic state of the system within mea-surable accuracy can be very much smaller than the number ofspecies in the system. Therefore, fewer equations are required todescribe the evolution of the system. A further advantage is thatonly rates of the slow constraint-changing reactions are neededand these are the ones most likely to be known. Reactions whichdo not change any constraint are in equilibrium and need not bespecified. It should be emphasized that the successful implementa-tion of the RCCE method depends critically on the constraints em-ployed and a knowledge of the rates of the constraint-changingreactions is required. When rates are uncertain or unknown, thePrinciple of Bayesian Inference suggests that the best choice is toassume that the corresponding reactions are in constrained-equi-librium (CE).

The primary objectives of this paper are to: (1) review theworking equations required to implement the RCCE method forchemically reacting systems, (2) discuss the issues involved inthe selection and transformation of the constraints, and (3) pres-ent a method for determining the initial conditions for highlynon-equilibrium systems, for which concentrations of some spe-cies are zero. To illustrate the method, RCCE calculations of theoxidation of C1 hydrocarbons in a constant volume adiabaticchamber have been made and compared with the results of aDKM.

2. Rate-controlled constrained-equilibrium (RCCE) method

A detailed description of the rate-controlled constrained-equi-librium (RCCE) method is given in Ref. [16]. A concise summaryof the working equations for chemically reacting gas mixtures is gi-ven below.

It is assumed that energy exchange reactions are sufficientlyfast to equilibrate the translational, rotational, vibrational, andelectronic degrees of the system subject to constraints on the vol-ume, V, and the total energy, E, of the system. Under these condi-tions, the energy can be written

E ¼ ETðTÞN ð1Þ

where ETðTÞ is the transpose of the species molar energy vector, Tthe temperature and N is the vector of species mole numbers. It isfurther assumed that the perfect gas model applies and the speciesconstraints, C, can be expressed as a linear combination of the spe-cies mole numbers in the form

C ¼ AN ð2Þ

where A is the nc � nsp constraint matrix, nc the number of con-straints, nsp the number of species and C and N are column vectorsof length nc and nsp. Maximizing the entropy, S(E, V, C), subject tothe constraints (2) using the method of undetermined Lagrangemultipliers, [16], we obtain the constrained-equilibrium composi-tion of the system

Nc ¼ ðM=pÞ expð�l� � lcÞ ð3Þ

where M is the total mole number, p = MRT/V the pressure, andlc ¼ ðh0 � Ts0Þ is the dimensionless standard Gibbs free energyfor nsp species, and

lc ¼ �ATc ð4Þ

is the dimensionless constrained-equilibrium Gibbs free energy ofthe species. In Eq. (4), AT is the transpose of the constraint matrix,and c is the vector of dimensionless constraint potentials (Lagrangemultipliers) conjugate to the constraint vector, C.

Given the values of the constraints, C, and energy, E, of the sys-tem, substitution of Eq. (3) into Eqs. (1) and (2) gives a set of nc + 1transcendental equations which can be solved for the temperature,TðE;V ;C;l�Þ; and the constraint potentials, cðE;V ;C;l�Þ, using gen-eralized equilibrium codes such as GNASA [17] or GSTANJAN [17].Finally substituting Eq. (4) into Eq. (3) gives the constrained-equi-librium composition of the system, NcðE;V ;C; cÞ.

2.1. Rate-equations for the constraints

It is assumed that changes in the chemical composition of thesystem are the results of chemical reactions of the type

m�X $ mþX ð5Þ

where X is the species vector m� and mþ are nr � nsp matrices of stoi-chiometric coefficients of reactants and products, respectively, andnr is the number of reactions. The corresponding rate-equations forthe species can be written

_N ¼ Vmr ð6Þ

where m ¼ mþ � m�, r ¼ rþ � r�, and rþ and r� are the forward andreverse reaction rate column vectors of length nr.

Differentiating Eqs. (1) and (2) with respect to time and usingEq. (6), we obtain equations for the energy

_E ¼ _T CTvN þ ET _N ð7Þ

and constraints

_C ¼ AN ¼ VBr ð8Þ

where

B ¼ Am ð9Þ

is an nc � nr matrix giving the change of constraints due to the nr

elementary chemical reactions among species and Cv � @E=@T isthe molar specific heat vector at constant volume. It follows fromEq. (9) that a reaction k for which all Bik are zero will be in con-strained-equilibrium and that a constraint i for which all Bik are zerowill be conserved. The latter is assumed to be the case for theelements.

Given equations for the state variables, V(t) and E(t), and initialvalues for the species, Nð0Þ, Eqs. (7) and (8) can be integrated instepwise fashion to obtain the temperature, T(t) and species con-straints, C(t). At each time step, a generalized equilibrium code,such as GNASA or GSTANJAN previously cited, must be used todetermine the temperature, T(E, V, C, l�), and constrained-equilib-rium composition, Nc(E, V, C, l�). These, in turn, can be used toevaluate the reaction rates r(T, V, l�, m, Nc) required for the nextstep. Note that only the rates of reactions which change con-straints, i.e. those for which Bik – 0, are required for RCCE calcula-tions. All other reactions are in constrained-equilibrium and neednot be specified.

2.2. Rate equations for the constraint potentials

Although direct integration of the rate-equations for the con-straints is relative straight forward and simple to implement, it

M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885 1873

has proved to be relatively inefficient and time consuming due tothe slowness of the constrained-equilibrium codes currently avail-able [17]. An alternative method, first proposed by Keck [16] andimplemented in later works [18–19] with colleagues, is the directintegration of the rate-equations for the constraint potentials. Thismethod has also recently been investigated by Tang and Pope [13]and Jones and Rigopoulos [14].

Differentiating Eq. (3) with respect to time and substituting theresult into Eqs. (7) and (8) yields the nc + 1 implicit equations forthe energy,

DC _c� DV

_VV� DT

_TTþ _E ¼ 0 ð10Þ

where

DCi ¼X

j

aijEjNcj ð10aÞ

DV ¼X

j

EjNcj ð10bÞ

DT ¼X

j

ðCvjT þ E2j =RTÞNc

j ð10cÞ

and the constraint potentials,

CC _c� CV

_VV� CT

_TTþ Br ¼ 0 ð11Þ

where

CCik ¼X

j

aijakjNcj ð11aÞ

CVi ¼X

j

aijNcj ð11bÞ

CTi ¼X

j

aijEjNcj =RT ð11cÞ

In this case, given equations for the state variables, V(t) and E(t),and the initial temperature, T(0), and constraint potentials, c(0),Eqs. (10) and (11) can be integrated using implicit ODE integrationroutines such as DASSL [20] to obtain the temperature, T(t), andconstraint potentials, cðtÞ: The constrained-equilibrium composi-tion, Nc(E, V, t), of the system can then be determined using Eq.(3). The number of unknowns is reduced from the nsp + 1 in aDKM calculation to nc + 1 in the RCCE calculation. As previouslynoted, only the rate constants for those reactions which changeconstraints, i.e. Bik – 0, are needed. Note that, once the constraintpotentials have been determined, the constrained-equilibriumconcentration of any species for which the standard Gibbs free en-ergy is known can be calculated whether or not it is explicitly in-cluded in the species list. It also follows that all species areimplicitly included in the RCCE rate-equations.

3. Selection of constraints

The careful selection of constraints is the key to the success ofthe RCCE method. Among the general requirements for the con-straints are that they must (a) be linearly independent combina-tions of the species mole numbers, (b) hold the system in thespecified initial state, (c) prevent global reactions in which reac-tants or intermediates go directly to products, and (d) determinethe energy and entropy of the system within experimental accu-racy. In addition, they should reflect whatever information is avail-able about rate-limiting reactions which control the evolution ofthe system on the time scale of interest.

In the present work, the focus is on applications of the RCCEmethod to chemically reacting gas phase mixtures. In the temper-ature and pressure range of interest, the rates of nuclear and ioni-zation reactions are negligible compared to those for chemical

reactions and the fixed constraints are the neutral elements ofhydrogen, carbon, oxygen, nitrogen, . . . , designated by EH, EC,EO, EN, . . .

Under these conditions, the slowest reactions controlling thechemical composition are three-body dissociation/recombinationreactions and reactions which make and break valence bonds. Suchreactions are slow in the endothermic direction because of the highactivation energies required, and in the exothermic direction be-cause of small three-body collision rates and small radical concen-trations. They impose slowly varying time-dependent constraintson the number of moles, M, of gas and the free valence, FV, of thesystem, respectively. A finite value of FV is a necessary conditionfor chain branching chemical reactions to proceed.

A third important time-dependent constraint, imposed by slowOO bond-breaking reactions, is the free oxygen, FO, defined as anyoxygen atom not directly bound to another oxygen atom. An in-crease in FO is a necessary condition for the formation of the majorreaction products of hydrocarbon oxidation, H2O, CO2 and CO.

Two additional time-dependent constraints which slightly im-prove the agreement between RCCE and DKM calculations undersome conditions are: OHO„OH + O and DCO„HCO + CO. TheOHO constraint is a consequence of the relatively slow con-straint-changing reaction RH + OH M H2O + R coupled with the fastreaction RH + O = OH + R which equilibrates OH and O. The DCOconstraint is a consequence of the slow spin-forbidden reactionCO + HO2 M CO2 + OH coupled with the fast reaction HCO + O2 =CO + HO2 which equilibrates HCO and CO.

For systems involving the elements C, H, and O the eight con-straints EH, EO, EC, M, FV, FO, OHO, and DCO are independent ofthe initial reactants and may, therefore, be considered ‘‘universal”constraints. Along with the equilibrium reactions,

H2 þ O ¼ OHþH ðCE1ÞH2 þHOO ¼ H2O2 þH ðCE2ÞHCOþ O2 ¼ COþHO2 ðCE3Þ

they are sufficient to determine the constrained-equilibrium molefractions of the 11 major hydrocarbon combustion products H, O,OH, HO2, H2, O2, H2O, H2O2, HCO, CO and CO2 under both highand low temperature conditions.

In the present investigation of C1 hydrocarbon oxidation, fouradditional fuel-dependent constraints have been used. The first isa constraint on the fuel, FU, imposed by slow hydrogen – abstrac-tion reactions of the type FU + O2 M FR + HO2 and even slower dis-sociation/recombination of the type AB + M M A + B + M. Thisconstraint is necessary to hold the system in its initial state. Thesecond is a constraint on fuel radicals, FR, which is necessary to pre-vent the equilibration of forbidden exothermic global reactionssuch as CH3 + 2O2 + 2H2O = CO2 + 2H2O2 + H2 + H which wouldotherwise convert fuel radicals directly to major products. Thethird is a constraint on alkylperoxides, APO„CH3OOH + CH3OO +CH2OOH, imposed by slow reactions which convert APO tohydroperoxides coupled with fast reactions which equilibrate thespecies comprising APO, and the fourth is a constraint on alcohol plusformaldehyde, ALCD„CH3OH + CH3O + CH2OH + CH2O imposed byrelatively slow reactions which generate/remove ALCD coupled withfast reactions which equilibrate the species comprising ALCD.

3.1. Transformation of constraints

The integration of Eqs. (10) and (11) for the constraint poten-tials requires inversion of the CC matrix. The performance of theimplicit integrators, such as DASSL, are quite sensitive to the struc-ture of this matrix. In general, the codes work best when the largeelements of the matrix lie on or close to the main diagonal. This isnot usually the case for the initial set of constraints chosen and a

1874 M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885

variety of error messages such as ‘‘singular matrix” or ‘‘failure toconverge” may be encountered. The problem can almost alwaysbe solved by a transformation of the square sub-matrix relatingthe constraints and the major species to a diagonalized form. Thephysical meaning of the transformed matrix may not be clear inall cases but if it improves the speed and reliability of the integra-tor, the desired objective will have been achieved. In this connec-tion, it is important to note that any linear combination of theoriginal linearly independent constraints for which the integratorworks should give the same final result. This can be a useful checkon the numerical results.

A general transformation of the constraint matrix can be madeas follows. Assume G is a non-singular square transformation ma-trix of order nc. Multiplying Eq. (2) through by this matrix yields

C�¼ GC ¼ A

�N ð12Þ

where

A�¼ GA ð13Þ

is the transformed nc � nsp constraint matrix that relates the trans-formed constraint vector, C

�, to the species vector, N. The corre-

sponding transformation for the reaction matrix is

~B ¼ GB ð14Þ

The transformation equations for the constraint potentials can beobtained by noting that the Gibbs free energy, lc; is invariant underthe transformation. Eq. (4) then gives

�lc ¼ ATc ¼ ~AT~c ¼ ðGAÞT c

�¼ AT GT c

�ð15Þ

Multiplying Eq. (15) by A we obtain

�Alc ¼ Sc ¼ SGT ~c ð16Þ

where

S ¼ AAT ð17Þ

is a symmetric matrix of order nc. Assuming that S is non-singular, itfollows from Eq. (16) that2

c ¼ GT ~c ð18aÞ

and, since G is non-singular,

~c ¼ ðGTÞ�1c ð18bÞ

Table 1Definition of the constraints used in the present work.

Constraint Definition of the constraint

1 EC Elemental carbon2 EO Elemental oxygen3 EH Elemental hydrogen4 M Total number of moles5 FV Moles of free valance (any unpaired valence electron)6 FO Moles of free oxygen (any oxygen not directly attached to

another oxygen)7 OHO Moles of water radicals (O + OH)8 DCO Moles of HCO + CO9 FU Moles of fuel molecules

10 FR Moles of fuel radicals11 APO Moles of alkylperoxides (CH3OO + CH3OOH + CH2OOH)12 ALCD Moles of alcohols + aldehydes (CH3O + CH3OH + CH2OH + CH2O)

4. Determination of initial conditions

For systems initially in a constrained-equilibrium state, the ini-tial values of the constraint potentials are finite. However, for asystem initially in a non-equilibrium state, where the concentra-tions of one or more species is zero, it can be seen from Eqs. (3)and (4) that one or more constraint potentials must be infinite. Thiscondition is encountered in ignition-delay-time calculations,where the system is initially far from equilibrium and the initialconcentrations of all species except the reactants are assumed tobe zero.

One method of dealing with this problem is to assign small par-tial pressures to as many major product species as required to givefinite values for the constraint potentials. Ideally the choice shouldbe made in such a way that the partial pressures of all other prod-uct species will be smaller than the assigned partial pressures. Areasonable initial choice for a major species corresponding to aconstraint is the species with the minimum standard Gibbs freeenergy in the group of species included in the constraint.

To implement this method, Eq. (2) is decomposed in the form

C ¼ AN ¼ A11 A12

h i N1

N2

� �ð19Þ

where A11 is an nc � nc non-singular, square matrix giving the con-tribution of the major species vector, N1, to the constraint vector, C,and A12 is an nc � ðns � ncÞmatrix, giving the contribution of N2 to C.The corresponding decomposition of Eq. (4) is

l ¼l1

l2

" #¼ �ATc ¼ �

AT11

AT12

" #c ð20Þ

Initial values for the constraint potentials can now be obtained byassuming l1ð0Þ � l2ð0Þ. This gives

cð0Þ ¼ �ðAT11Þ�1l1ð0Þ ¼ �ðAT

11Þ�1ðln p1ð0Þ þ l�Þ ð21Þ

Having determined the initial values of the constraint potentialvector, one can now check the assumption that the initial partialpressures of the minor species are small using the relation

ln p2ð0Þ ¼ �AT12cð0Þ � l�2 ð22Þ

If they are not, an alternative choice for the major species will usu-ally solve the problem.

In initial RCCE calculations using constraints based on kineticconsiderations, problems involving the convergence of the implicitintegrators used were frequently encountered. These were causedprimarily by the fact the CC matrix in Eq. (11) contained large offdiagonal elements. These problems were solved by a transforma-tion which diagonalized the A11 matrix. Using the transformationmatrix G ¼ A�1

11 we obtain from Eq. (12)

C�¼A�1

11 C¼A�111 A11 A12

h i N1

N2

� �¼ I11 A�1

11 A12

h i N1

N2

� �¼A�

N ð23Þ

and it follows from Eq. (15) that

l1ð0Þ ¼ ln p1ð0Þ þ l� ¼ �I11c�ð0Þ ¼ �~cð0Þ ð24Þ

It can be seen from this equation, that in the diagonalized represen-tation, the initial constraint potentials are simply the initial Gibbsfree energies of the major species chosen as surrogates. The trans-formed reaction rate matrix is

B�¼ A�1

11 B ð25Þ

5. RCCE calculations for C1 hydrocarbon oxidation

To illustrate the RCCE method, we consider the homogeneousstoichiometric oxidation of three mono-carbon fuels, namelymethane (CH4), methanol (CH3OH) and formaldehyde (CH2O) with

Table 2bDiagonalized A matrix for the C1 system with 12 constraints and 29 species.

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29CO2 O2 H2 H2O2 HO2 H2O HO CO CH4 CH3 CH3OOH H2CO H CH3O HCO CH CH2 C O CH2OH CH3OH HOCO HOCHO OCHO CH3OO CH2OOH HOOCHO HOOCO OOCHO

CO2 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 0 0 0 1 1 1 0 0 1 1 1O2 0 1 0 0 0 0 0 0 0 0 0 0 0 �1 �1 �2 �2 �2 0 �1 �1 �1 �1 �1 0 0 �1 �1 �1H2 0 0 1 0 0 0 0 0 0 0 0 0 1 0 0 2 2 2 0 0 0 0 0 0 0 0 0 0 0H2O2 0 0 0 1 0 0 0 0 0 0 0 0 �1 0 0 �1 0 �2 �1 0 1 0 1 0 �1 �1 2 1 1HO2 0 0 0 0 1 0 0 0 0 0 0 0 1 1 1 3 2 4 1 1 0 1 0 1 1 1 0 1 1H2O 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 �2 �2 �2 0 0 0 0 0 0 0 0 �1 �1 �1HO 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0CO 0 0 0 0 0 0 0 1 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0CH4 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0CH3 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0CH3OOH 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0 0 0H2CO 0 0 0 0 0 0 0 0 0 0 0 1 0 1 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0

Table 2aConstraint matrix A for the C1 system with 12 constraints and 29 species.

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29CO2 O2 H2 H2O2 HO2 H2O HO CO CH4 CH3 CH3OOH H2CO H CH3O HCO CH CH2 C O CH2OH CH3OH HOCO HOCHO OCHO CH3OO CH2OOH HOOCHO HOOCO OOCHO

EC 1 0 0 0 0 0 0 1 1 1 1 1 0 1 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1EO 2 2 0 2 2 1 1 1 0 0 2 1 0 1 1 0 0 0 1 1 1 2 2 2 2 2 3 3 3EH 0 0 2 2 1 2 1 0 4 3 4 2 1 3 1 1 2 0 0 3 4 1 2 1 3 3 2 1 1M 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1FV 0 0 0 0 1 0 1 0 0 1 0 0 1 1 1 3 2 4 2 1 0 1 0 1 1 1 0 1 1FO 2 0 0 0 0 1 1 1 0 0 0 1 0 1 1 0 0 0 1 1 1 2 2 2 0 0 1 1 1OHO 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0DCO 0 0 0 0 0 0 0 1 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0FU 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0FR 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0APO 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0 0 0ALCD 0 0 0 0 0 0 0 0 0 0 0 1 0 1 0 0 0 0 0 1 1 0 0 0 0 0 0 0 0

M.Janbozorgi

etal./Com

bustionand

Flame

156(2009)

1871–1885

1875

CH4

O2 HO2H2O2

H

O2

O

OH

..

.

.

.

CH3

O2

CH3OO

O

CH3O CH3OH

CH2OH

OH

H2O

CH4

.

. ..

.

. CH3OOH

.

CO

CH2OOH

O2

.

CH2O

CHO.

HO2

.

OH

CO2

HO2

.

1

2

3 4

5

6

7

8

9

10

.

Fig. 1. RCCE reaction flow diagram for CH4 oxidation.

1876 M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885

pure oxygen in a constant volume reactor over a wide range of ini-tial temperatures (900–1500 K) and pressures (1–100 atm), andequivalence ratios (0.8–1.2). The 12 constraints used in the RCCEcalculations are summarized in Table 1.

The DKM calculations with which the RCCE results are com-pared involve only C1 chemistry and include 29 species and 133reactions, without nitrogen chemistry. Twenty species and 100reactions were taken from the widely known and used GRI-Mech3.0 [21] mechanism. This model does not include alkylperox-ides and therefore is not expected to be valid under high pressurelow temperature conditions. To obtain a model valid under theseconditions an additional nine alkylperoxides and organic acidswere included along with 33 reactions with rates taken from [22]or estimated by the authors.

Of the 133 reactions employed in the DKM calculations only102 reactions change one or more of the constraints in Table 1and are therefore of interest. The remaining 31 are in con-strained-equilibrium and are not needed in RCCE calculations. Inprinciple, only the fastest reaction in a group which changes a gi-ven constraint is required for a system to go from the specified ini-tial state to the final chemical-equilibrium state determined by theelements. However, because the fastest reaction may change as thesystem evolves, it is usually necessary to include additional reac-tions to achieve the desired accuracy for the time evolution ofthe system.

In this work, excellent agreement between DKM calculationsand RCCE calculations was obtained for all C1 species using 12 con-straints and 102 reactions and acceptable agreement was obtainedusing 12 constraints and 20 reactions for methane, 10 constraintsand 16 reactions for methanol, and nine constraints and 12 reac-tions for formaldehyde.

5.1. Methane (CH4) oxidation

The constraint matrix for CH4 used in this work is shown inTable 2a and the corresponding diagonalized matrix used toset the initial conditions and carry out the numerical calcula-tions is shown in Table 2b. 12 constraints and 29 species areincluded in the Tables. The corresponding RCCE reaction flowdiagram is shown in Fig. 1. The constrained species are enclosedby dashed lines. Except for the initiation steps, this diagram alsoincludes the sub-mechanisms involved in the oxidation ofCH3OH and CH2O.

As can be seen, the oxidation process is initiated by the highlyendothermic CH4 constraint-changing H-abstraction reaction

CH4 þ O2 $ CH3 þHO2 ðRC1Þ

This is followed by the two competing CH3 constraint-changingreactions

CH3 þ O2 þM$ CH3OOþM ðRC2ÞCH3 þ O2 $ CH3Oþ O ðRC3Þ

The first is most important at low temperatures and is followed bythe constrained-equilibrium reactions

CH3OOþH2O2 ¼ CH3OOHþHO2 ðCE1ÞCH3OOHþHO2 ¼ CH2OOHþH2O2 ðCE2Þ

which equilibrate the alkylperoxides in APO. The second is mostimportant at high temperatures and is followed by the con-strained-equilibrium reactions

H2 þ O ¼ OHþH ðCE3Þ

which equilibrates the water radicals in OHO, and constrained-equilibrium reactions

CH3OþH2O2 ¼ CH3OHþHO2 ðCE4ÞCH3OHþHO2 ¼ CH2OHþH2O2 ðCE5ÞCH3Oþ O2 ¼ CH2OþHO2 ðCE6Þ

which equilibrate the free oxygen species in ALCD. The CH4 con-straint-changing reaction

OHþ CH4 $ CH3 þH2O ðRC4Þ

converts OH in OHO to the product H2O and regenerates CH3.The stable intermediate CH2O in ALCD is oxidized by the con-

straint-changing H-abstraction reaction

CH2Oþ O2 $ CHOþHO2 ðRC5Þ

and the constrained-equilibrium reaction and the fast equilibriumreaction

CHOþ O2 ¼ COþHO2 ðCE7Þ

to form DCO. The CO in DCO is converted to the product CO2 by theFO constraint-changing reaction

COþHO2 $ CO2 þ OH ðRC6Þ

Finally the eight major H/O species are determined by the four con-straint-changing reactions

Logt(sec)

Tem

pera

ture

(K)

-4 -3 -2 -1 0 1

1000

1500

2000

2500

3000

3500

4000

4500RCCEDKM

P=1atm

P=100atm

P=50atmP=10atm

Logt(sec)

Tem

pera

ture

(K)

-7 -6 -5 -4 -3

1500

2000

2500

3000

3500

4000

4500RCCEDKM

P=1atm

P=100atm

P=50atmP=10atm

(a) (b)

Fig. 2. Comparison of RCCE and DKM temperature vs. time profiles for a stoichiometric mixture of CH4/O2 at different initial pressures and initial temperatures of 900 K and1500 K (time in s).

M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885 1877

HO2 $ Hþ O2 ðRC7ÞHO2 þHO2 $ H2O2 þ O2 ðRC8ÞHO2 þH$ OHþ OH ðRC9ÞHþ O2 $ OHþ O ðRC10Þ

which change M, FV, FO, and OHO, respectively, plus the two con-strained-equilibrium reactions

H2 þ O ¼ OHþH ðCE3Þ

H2 þHO2 ¼ H2O2 þH ðCE8Þ

and the elemental constraints EH and EO.As the above reactions proceed, the radical population increases

rapidly and all constraint-changing reactions included in the ki-netic model become involved in determining the evolution of thesystem and its approach to final chemical-equilibrium. Of particu-lar importance are reactions of the type

CH4 þ Q $ CH3 þHQ ðRC1:1Þ

and

Time(ms)

Tem

pera

ture

(K)

14 15 16 17 18

1000

2000

3000

4000

5000RCCEDKM

φ =0.8φ =1.2

φ =1.0

(a) (b

Fig. 3. Comparison of RCCE and DKM temperature vs. time profiles for CH4/O2 mixtures100 atm (a) and 1500 K and 1 atm (b) (time in s).

CH2Oþ Q $ CHOþHQ ðRC5:1Þ

where Q can be any radical, e.g. HO2, OH, H, O. Numbers in circles inFig. 1 correspond to the rate-controlling reactions mentioned above,e.g. 1 corresponds to (RC1), 2 corresponds to (RC2) and so on.

5.1.1. CH4 results for 133 reaction setTime-dependent temperature profiles of stoichiometric mix-

tures of methane and oxygen at initial temperatures of 900 K and1500 K and different initial pressures are shown in Fig. 2. All 12constraints listed in Table 2a have been included. As previouslynoted, only 102 of the reactions in the full set of 133 reactionschange constraints and are therefore required. The remaining 31are in constrained-equilibrium and are not needed in RCCEcalculations.

It can be seen that the agreement with DKM calculations isexcellent over the entire range of pressure and temperature cov-ered. Note that the temperature overshoot at low pressures dueto slow three-body recombination and dissociation reactions iswell reproduced by the constraint M on the total moles. Also, thesame comparisons have been made for rich (/ ¼ 1:2) and lean

Time(ms)

Tem

pera

ture

(K)

0.2 0.25 0.3

1500

2000

2500

3000

3500

4000RCCE,Phi=0.8DKM,Phi=0.8RCCE,Phi=1.0DKM,Phi=1.0RCCE,Phi=1.2DKM,Phi=1.2

)

with different equivalence ratios at initial temperatures and pressures of 900 K and

Log (Time)

Log

|T-T

i|

-10 -8 -6 -4 -2 0

-10

-5

0

5

T>TiiT<T

Log (Time)

Log

(Mol

e Fr

actio

n)

-10 -8 -6 -4 -2 0-14

-12

-10

-8

-6

-4

-2

0

CO2

O2

CO

H2O

CH4

Log (Time)

Tran

sfor

med

Con

stra

int P

oten

tials

-10 -8 -6 -4 -2 00

20

40

60

80

100

120

O2H2O

CH3CO2

H2

H2O2

HO2

CH4

H2

O2

H2O

HO2

CH3

H2O2

CH4

CO2

Log (Time)

Log

(Mol

e Fr

actio

n)

-10 -8 -6 -4 -2 0-14

-12

-10

-8

-6

-4

-2

0

CH3OOH

CH2O

CH3

CH2OCH3OO

CH3

CH3OOH

CH3OO

Log (Time)

Con

stra

int P

oten

tials

-10 -8 -6 -4 -2 0-150

-100

-50

0

50

100

EC

FOEO

FR

MFV

FU

EC

EHEH

Log (Time)

Log

(Mol

e Fr

actio

n)

-10 -8 -6 -4 -2 0

-14

-12

-10

-8

-6

-4

-2

0

H

H2

OH

H2O2

HO2

Log (Time)

Log

(Con

stra

int M

ole

Num

ber)

-10 -8 -6 -4 -2 0

-12

-9

-6

-3

0

3EH EO

EC

FR

M

FV

FO

FU

FU

FR

Log (Time)

Log

(Mol

e Fr

actio

n)

-10 -8 -6 -4 -2 0

-14

-12

-10

-8

-6

-4

-2

0

HOCO

OCHO

HOCHO

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)

Fig. 4. Comparison of various RCCE and DKM time profiles for stoichiometric CH4/O2 mixtures at initial temperature and pressure of 900 K and 100 atm (time in s).

1878 M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885

(/ ¼ 0:8) mixtures and are shown in Fig. 3. RCCE predictions ofignition-delay-times at low temperature, high pressure are within

1% of those by DKM. At high temperature, low pressure the resultsagree within 1–5% of accuracy.

Log|

T-T

i|

-10 -8 -6 -4 -2-10

-5

0

5 RCCE (12,102)DKM (29,133)

T > TiiT < T Lo

g(M

ole

Fra

ctio

n)

-10 -8 -6 -4 -2-14

-12

-10

-8

-6

-4

-2

0 O2

CO2CH4

H2O CO

CH4

Tra

nsfo

rmed

Co

nstr

ain

tPo

ten

tials

-10 -8 -6 -4 -220

40

60

80

100

O2

H2O

CH3CO2

H2

H2O2

HO2

CH4

H2

O2

H2O

HO2

CH3

H2O2

CH4

CO2

Log

(Mo

leF

ract

ion)

-10 -8 -6 -4 -2-18

-16

-14

-12

-10

-8

-6

-4

-2

0

CH3OOH

CH2O

CH3

CH2O

CH3OO

CH3

CH3OOH

CH3OO

Co

nst

rain

tPo

ten

tials

-10 -8 -6 -4 -2

-100

-50

0

50

EC

FO

EOFR

MFV

FU

EO

EH

EHEC

Log(

Mo

leF

ract

ion)

-10 -8 -6 -4 -2

-14

-12

-10

-8

-6

-4

-2

0

H2

HO2

H2O2

OH

H

H2O2

HO2

OH

H

H2

Log (Time) Log (Time)

Log (Time)Log (Time)

Log (Time) Log (Time)

Log (Time)Log (Time)

Log

(Con

stra

ins

Mol

eN

um

ber)

-10 -8 -6 -4 -2 0

-12

-9

-6

-3

0

3

EH EO

EC

FR

M

FV

FO

FU

FU

FR Log(

Mo

leF

ract

ion)

-10 -8 -6 -4 -2-18

-16

-14

-12

-10

-8

-6

-4

-2

0

HOCO

OCHO

HOCHO

(a)

(b)

(c)

(d)

(e)

(f)

(g)

(h)

Fig. 5. Comparison of various RCCE and DKM time profiles for stoichiometric CH4/O2 mixtures at initial temperature and pressure of 1500 K and 1 atm (time in s).

M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885 1879

0

0.005

0.01

0.015

0.02

0.025

0.03

9th = FR 10th = ALCD 11th = APO 12th = DCO

Igni

tion

Del

ay T

ime

(sec

)(T

i=90

0K, P

i=10

0 at

m)

RCCE DKM

0

0.0001

0.0002

0.0003

9th = FR 10th = ALCD 11th = APO 12th = DCO

Igni

tion

Del

ay T

ime

(sec

)(T

i = 1

500

K, P

i = 1

atm

)

RCCE DKM

Fig. 6. Comparison of ignition-delay-times for RCCE models using 9–12 constraints with the DKM for stoichiometric CH4/O2 mixtures at initial temperatures and pressures of900 K and 100 atm (top) and 1500 K and 1 atm (bottom).

1880 M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885

Additional results are shown on log–log plots in Fig. 4 for theinitial conditions 900 K and 100 atm, where the dominant radicalsare HO2 � CH3 at early times and CH3OO at late times, and in Fig. 5

Table 3Reduced reaction matrix B for C1 hydrocarbons (Units: cc, mole, kcal, s).

X

X

X X

CH2O CH3OH CH4 REK CH4 CH3 CH3OH C

X 1 CH4 + O2 = CH3 + HO2 �1 1 0 0X 2 CH4 + HO2 = CH3 + H2O2 �1 1 0 0X 3 CH4 + OH = CH3 + H2O �1 1 0 0X 4 CH4 + CH3OO = CH3 + CH3OOH �1 1 0 0X 5 CH3 + O2 + M = CH3OO + M 0 �1 0 0X 6 CH3 + O2 = CH3O + O 0 �1 0 0X 7 CH3 + HO2 = CH3O + OH 0 �1 0 0X 8 CH3 + CH3OO = CH3O + CH3O 0 �1 0 0

X 9 CH3OH + O2 = CH2OH + HO2 0 0 �1 0X 10 CH3OH + O2 = CH3O + HO2 0 0 �1 0X 11 CH3OH + HO2 = CH2OH + H2O2 0 0 �1 0X 12 CH3OH + OH = CH3O + H2O 0 0 �1 0

X X X 13 CH2O + O2 = CHO + HO2 0 0 0 �X X X 14 CH2O + HO2 = CHO + H2O2 0 0 0 �X X X 15 CH2O + OH = CHO + H2O 0 0 0 �X X X 16 H + O2 + M = HO2 + M 0 0 0 0X X X 17 OH + H + M = H2O + M 0 0 0 0X X X 18 OH + OH + M = H2O2 + M 0 0 0 0X X X 19 H + O2 = OH + O 0 0 0 0X X X 20 CH3OO + HO2 = CH3OOH + O2 0 0 0 0X X X 21 HO2 + HO2 = H2O2 + O2 0 0 0 0X X X 22 CO + HO2 = CO2 + OH 0 0 0 0X X X 23 H + HO2 = OH + OH 0 0 0 0X X X 24 CO + OH = CO2 + H 0 0 0 012 16 20

for the initial conditions 1500 K and 1 atm, where the dominantradicals are HO2 � CH3OO at all times. As can be seen in Figs. 3aand 4a, the temperature first decreases due to the fact that the

X X X X X CH2O Er Uncert.factor

X X X X X X CH3OH

X X X X X X X CH4

H2O M FV FO OHO DCO ALCD APO Log(Ar)

0 2 0 0 0 0 0 13.6 57.3 50 0 0 0 0 0 0 11.3 18.7 50 0 0 �1 0 0 0 13.5 2.1 1.40 0 0 0 0 0 0 11.3 18.6 10�1 0 0 0 0 0 1 12.0 0.0 30 2 2 1 0 1 0 12.9 29.4 30 0 2 1 0 1 0 13.3 0.0 30 0 2 0 0 2 �1 13.4 0.0 30 2 0 0 0 0 0 13.4 45.2 100 2 0 0 0 0 0 13.4 45.2 100 0 0 0 0 0 0 11.0 12.6 100 0 0 �1 0 0 0 13.7 1.5 2

1 0 2 0 0 1 �1 0 14.0 40.0 21 0 0 0 0 1 �1 0 13.6 12.0 31 0 0 0 �1 1 �1 0 13.7 �0.4 2

�1 0 0 0 0 0 0 15.4 0.0�1 �2 0 �1 0 0 0 15.5 0.0 2�1 �2 �2 �2 0 0 0 12.5 0.0 20 2 2 2 0 0 0 14.0 17.0 1.50 �2 0 0 0 0 0 13.5 4.9 5

0 �2 0 0 0 0 0 11.1 �1.6 30 0 2 1 �1 0 0 14.2 23.6 30 0 2 2 0 0 0 14.1 0.9 20 0 0 �1 �1 0 0 11.8 0.1 1.5

Log (Time)

Tem

pera

ture

(K)

-4 -3 -2 -1 0 1

1000

1500

2000

2500

3000

3500

4000

4500

5000 RCCE(12,20)DKM(29,133)

Pi =1atm

Pi =100atm

Pi =10atm

Log (Time)

Tem

pera

ture

(K)

-8 -6 -4 -

1500

2000

2500

3000

3500

4000

4500

5000 RCCE(12,20)DKM(29,133)

Pi=1atm

Pi=100atm

Pi=10atm

Fig. 7. Comparison of reduced RCCE(12, 20) and DKM temperature vs. time profiles for stoichiometric mixtures of CH4/O2 for different initial pressures at initial temperaturesof 900 K (left) and 1500 K (right). The reduced RCCE(12, 20) model employed 12 constraints and 20 reactions listed in Table 3 (time in s).

M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885 1881

initiation reactions are all endothermic then later increases as exo-thermic reactions become important. The agreement betweenRCCE and DKM calculations, especially with regard to the time atwhich the temperature difference becomes positive, is excellent.

Figs. 3b and 4b show the time dependence of the constraintpotentials for the diagonalized constraint matrix and Figs. 3c and4c show the corresponding constraint potentials for the primaryconstraint matrix. Note that for the primary constraint matrix, allthe time-dependent constraint potentials go to zero at equilibriumas required, while those for the elements go to values identicalwith those obtained from the STANJAN equilibrium code [23]. Oncethe constraint potentials have been determined, the constrained-equilibrium mole fractions of any species for which the standardGibbs free energy is known can be calculated from Eq. (3).

The fixed elemental constraints and the most important time-dependent constraints M, FV, FO, FU, and FR are shown in Figs. 3dand 4d and the mole fractions of the major species are shown in Figs.3e–h and 4e–h. It can be seen that overall agreement is very good.

To investigate the sensitivity of the ignition-delay-time to thenumber of constraints used, a series of RCCE calculations was car-

Log (Time)

Log(

X H2)

-8 -6 -4 -2 0

-20

-15

-10

-5

0 RCCE (12,20)DKM(29,133)

Ti = 900 KPi = 100 atm

φ = 1

Fig. 8. Comparison of reduced RCCE(12, 20) and DKM H2 mole fractions vs. time profitemperatures of 900 K (left) and 1500 K (right). Note that H2 is not explicitly includedconstraint potentials.

ried out starting with the eight constraints EH, EC, EO, M, FV, FO,OHO, and FU and adding additional constraints one at a time.The results for both high and low temperature conditions are com-pared with those of the DKM in Fig. 6. It can be seen that for hightemperature conditions, nine constraints are sufficient to give theagreement within 5%, while for low temperature conditions, 11constraints are required to give the same agreement.

5.1.2. Reduced CH4 reaction mechanismIn the initial studies, a set of 133 reactions was used for both the

RCCE and DKM calculations. As previously discussed, not all of theseare of equal importance, especially in RCCE calculations where, inprinciple, only one independent reaction for each time-dependentconstraint is required to allow the system to relax from the specifiedinitial state to the correct final chemical-equilibrium state.

By systematically eliminating reactions, a reduced mechanismfor C1 hydrocarbons involving the 24 constraint-changing reac-tions in Table 3 has been found. The set of reactions used forthe individual fuels: CH4, CH3OH and CH2O, are indicated bycross marks in the left columns of the Table. The constraints

Log (Time)

Log

(XH

2)

-8 -6 -4 -2 0-16

-14

-12

-10

-8

-6

-4

-2

0 RCCE (12,20)DKM(29,133)

Ti = 1500 KPi = 1 atm

φ = 1

les for stoichiometric mixtures of CH4/O2 for different initial pressures at initialin the RCCE (12, 20) model and the H2 mole fractions were calculated from the

1882 M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885

used for each of the fuels are indicated by cross marks in the toprows of the Table. Arrhenius rate-parameters and uncertaintyfactors taken from Tsang and Hampson [22] are also shown inthe right hand columns of the Table. It can be seen that manyof the rates have estimate uncertainties greater than a factorof three even though these are among the simplest and bestknown reactions. For each constraint where more than one reac-tion is listed, the redundant reactions are important at differentstages in the evolution of the system.

-35

-30

-25

-20

-15

-10 -5 0

H2OCOOHCO2O2H2OHHO2H2O2HCOHOCOHOCHOCH2OCH2OHCCH3CH2CHHCCOCH4CH3OHCH3OCH2COC2H2CH3OOHCCOHCH3OOHC2HHOOCOOCHOCH2OOHC2H3CH3CHOC2H4C2H5OOCHOC2H6HOOCHOC3H7C3H8

STANJAN

RC

CE

Detailed

Ti = 900 K, Pi = 100 atm

Fig. 9. Comparison of final C1 and C2 species mole fractions calculated using RCCE(12, 133for initial temperatures and pressures 900 K and 100 atm (left) and 1500 K and 1 atm (r

Reduced RCCE calculations for CH4 were carried out using 12constraints and 20 reactions. The temperature vs. time plots areshown in Fig. 7 and it can be seen that the RCCE calculations giveignition-delay-times within a few percent of those obtained usingthe DKM in both low and high temperature regimes.

In the RCCE approach, the moles of any species can be foundfrom the constraint potentials using Eq. (3). It follows that any spe-cies which can be made from the elements will evolve dynamicallyeven if there is no kinetic path provided in the mechanism. This

-30

-25

-20

-15

-10 -5 0

H2OCOOHH2O2HCO2OHO2H2O2HCOHOCOCH2OHOCHOCCHCH2CH3CH2OHCH4HCCOCH3OCH3OHCH2COC2H2OCHOHOOCOCH3OOC2HHCCOHCH3OOHCH2OOHC2H3CH3CHOC2H4OOCHOHOOCHOC2H5C2H6C3H7C3H8

STANJAN

RC

CE

Detailed

Ti = 1500 K, Pi = 1 atm

) and DKM and C1 kinetics with equilibrium mole fractions obtained using STANJANight). Note that the DKM C2 mole fractions are frozen at the initial assigned values.

CH3OH

O2

CO

O2

CHO

OH

.

HO2

.

.

CH3O CH2OH . .

CH2O

CO2

HO2H2O2

H

O2

O

OH

..

.

.

.

. HO2

Fig. 10. RCCE reaction flow diagram of CH3OH/O2 oxidation.

CH2O

O2

CO

CHO

HO2

.

HO .

CO2

HO2H2O2

H

O2

O

OH

..

.

.

.

HO2

.

.

Fig. 12. RCCE reaction flow diagram for CH2O/O2 oxidation.

M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885 1883

point is illustrated in Fig. 8 which compares reduced RCCE(12, 20)and DKM(29, 133) results for the mole fractions of H2 as a functionof time. Although no reactions involving H2 are explicitly includedin the RCCE(12, 20) calculations and H2 is not on the species list,both the ignition-delay-time and the equilibrium mole fractionsare in good agreement with the DKM(29, 133) results. This pointis further illustrated in Fig. 9 where the final equilibrium mole frac-tions of the most abundant C1 and C2 species calculated using onlyC1 kinetics in both the RCCE model and DKM are compared withvalues obtained using the STANJAN equilibrium code. For C1 spe-cies, all values agree perfectly. However for C2 species, the RCCEand STANJAN values agree but the DKM values are zero for the

Log (Time)

Tem

pera

ture

(K)

-4 -3 -2 -1 0 1

1000

1500

2000

2500

3000

3500

4000

4500 RCCE (10,16)DKM(29,133)

P = 50 atm

P = 1 atm

P = 10 atm

P = 100 atm

Fig. 11. Comparison of reduced RCCE(10, 16) and DKM temperature vs. time profiles fotemperatures of 900 K (left) and 1500 K (right). The reduced RCCE(10, 16) model for CH

obvious reason that the list of species in the DKM does not includeC2 species. The ability to estimate the mole fractions of species noton a reaction list can be useful in determining whether they arelikely to be kinetically important and therefore to be added tothe model.

The fact that RCCE calculations can be carried out with fewerrate-equations than unknowns and still give constrained-equilib-rium mole fractions for all species that can be made from the ele-ments in the system is an important feature of the RCCE methodthat distinguishes it from most other reduction techniques thatdo not use a constrained-equilibrium manifold for reconstructingthe missing species concentrations.

5.2. Methanol (CH3OH) oxidation

The same set of basic constraints used for CH4 can be used forthe oxidation of methanol. However in this case, the fuel moleculeFU is CH3OH and fuel radical constraint, FR, is CH3O + CH2OH + -CH2O. In addition, an examination of the kinetics shows that alkyl-peroxides reactions are not important and the APO constraint canbe eliminated. This results in a reduction of the total number of

Log (Time)

Tem

pera

ture

(K)

-8 -6 -4 -2

1500

2000

2500

3000

3500

4000

4500RCCE (10,16)DKM (29,133) P = 100 atm

P = 1 atm

P = 10 atm

P = 50 atm

r stoichiometric mixtures of CH3OHOHO/O2 for different initial pressures at initial3OHOHO/O2 is shown in Table 3 (time in s).

Log (Time)

Tem

pera

ture

(K)

-5 -4 -3 -2500

1000

1500

2000

2500

3000

3500

4000

4500

5000RCCE(9,12)DKM(29,133)

Pi =1atm

Pi =100atm

Pi =10atm

Pi =50atm

Log (Time)

Tem

pera

ture

(K)

-8 -7 -6 -5 -4

1500

2000

2500

3000

3500

4000

4500

5000RCCE(9,12)DKM(29,133)

Pi =100atm

Pi =1atmPi =10atm

Pi =50atm

Fig. 13. Comparison of reduced RCCE(9, 12) and DKM temperature vs. time profiles for stoichiometric mixtures of CH2OOHO/O2 for different initial pressures at initialtemperatures of 900 K (left) and 1500 K (right). The reduced RCCE(9, 12) model for CH2OOHO/O2 is shown in Table 3 (time in s).

1884 M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885

constraints from 12 to 10. The corresponding reaction diagram isshown in Fig. 10.

5.2.1. Reduced CH3OH reaction mechanismA reduced set of 16 reactions for CH3OH oxidation is included in

Table 3. RCCE calculations using these reactions and the seventime-dependent constraints included in Table 3 are compared withDKM calculations using 133 reactions and 29 species in Fig. 11. Itcan be seen that the temperature vs. time plots are in excellentagreement over the entire range of temperature and pressure cov-ered. The agreement for all other variables is similar to that for CH4.

5.3. Formaldehyde (CH2O) oxidation

The reaction diagram for CH2O oxidation is shown in Fig. 12 and areduced kinetic model is included in Table 3. The reduced kineticmodel in this case is even simpler than that for CH3OH and only nineconstraints and 12 reactions are required in the RCCE calculations. Acomparison of temperature vs. time plots is shown in Fig. 13 andagain it can be seen that the agreement between RCCE and DKM cal-culations is excellent. As in the case of CH3OH, plots of all other vari-ables are also in excellent agreement and are similar to those for CH4.

6. Summary and conclusions

RCCE calculations of methane, methanol and formaldehyde oxi-dation over a wide range of initial temperatures and pressureshave been made using up to 12 constraints and 133 reactionsand excellent agreement with ‘‘Detailed Kinetic Model” (DKM) cal-culations using 29 species and the same 133 reactions has been ob-tained. In addition, reduced sets of 20 reactions for methane, 16reactions for methanol and 12 reactions for formaldehyde havebeen found, which when used in RCCE calculations give resultsidentical to those obtained using the full 133 reactions.

Among the important features of the RCCE method for simplify-ing the kinetics of hydrocarbon oxidation are:

1. It is based on the well established Maximum Entropy Principleof Thermodynamics rather than mathematical approximations.

2. The entropy always increases as the system evolves and anapproach to the final chemical-equilibrium state determinedby the specified elements is guaranteed. This is not true forDKMs where only the listed species are included and all othersare missing.

3. The total number of constraints required to determine theconstrained-equilibrium state of a complex chemical systemcan be very much smaller than the number of species inthe system, and an estimate of the concentrations of any spe-cies for which the standard Gibbs free energy is known canbe obtained even though the species is not explicitly includedin the model used.

4. Since the number of constraints required to determine the con-strained-equilibrium state of a system can be very much smal-ler than the number of species in the system, fewer rate-equation are required to determine its evolution.

5. If the only constraints on the system are the elements and thestate variables, RCCE calculations automatically reduce tolocal-thermodynamic-equilibrium (LTE) calculations.

6. If the number of constraints in an RCCE model is equal to thenumber of species in a DKM, the number of rate-equationswill be the same but the species mole fractions will differbecause the RCCE model will include all species which canbe made from the elements but the DKM will include onlythe listed species.

7. The accuracy of the results can be improved by adding con-straints one at a time until no further improvement is obtain.This can be done using either information available about prob-able rate-controlling reactions or trail and error.

Acknowledgments

This work was partially supported by ARO under technicalmonitoring of Dr. Ralph Anthenion. The authors are indebted toDr. Yue Gao and Dr. Donald Goldthwaite for helpful commentsand stimulating discussions.

References

[1] S.W. Benson, J. Chem. Phys. 20 (1952) 1605.[2] M. Rein, Phys. Fluids A 4 (1992) 873.[3] U. Mass, S.B. Pope, Combust. Flame 88 (1992) 239.[4] S.H. Lam, D.A. Goussis, Proc. Combust. Inst. 22 (1988) 931.[5] O. Oluwole, B. Bhattacharjee, J.E. Tolsma, P.I. Barton, W.H. Green, Combust.

Flame 146 (1–2) (2006) 348–365.[6] T. Lu, C.K. Law, Combust. Flame 144 (1–2) (2006) 24–36.[7] Z. Ren, S.B. Pope, A. Vladimirsky, J.M.J. Guckenheimer, The ICE-PIC method for the

dimension reduction of chemical kinetics, Chem. Phys. 124 (114111) (2006).[8] J.C. Keck, D. Gillespie, Combust. Flame 17 (1971) 237.[9] R. Law, M. Metghalchi, J.C. Keck, Proc. Combust. Inst. 22 (1988) 1705.

M. Janbozorgi et al. / Combustion and Flame 156 (2009) 1871–1885 1885

[10] P. Bishnu, D. Hamiroune, M. Metghalchi, J.C. Keck, Combust. Theory Model. 1(1997) 295–312.

[11] D. Hamiroune, P. Bishnu, M. Metghalchi, J.C. Keck, Combust. Theory Model. 2(1998) 81.

[12] V.A. Yousefian, Rate-controlled constrained-equilibrium thermochemistryalgorithm for complex reacting systems, Combust. Flame 115 (1998)66–80.

[13] Q. Tang, S.B. Pope, A more accurate projection in the rate-controlledconstrained-equilibrium method for dimension reduction of combustionchemistry, Combust. Theory Model. (2004) 255–279.

[14] W.P. Jones, S. Rigopoulos, Rate-controlled constrained equilibrium:formulation and application of non-premixed laminar flames, Combust.Flame 142 (2005) 223–234.

[15] W.P. Jones, S. Rigopoulos, Reduced chemistry for hydrogen and methanolpremixed flames via RCCE, Combust. Theory Model. 11 (5) (2007) 755–780.

[16] J.C. Keck, Prog. Energy Combust. Sci. 16 (1990) 125.[17] P.S. Bishnu, D. Hamiroune, M. Metghalchi, Development of constrained

equilibrium codes and their applications in nonequilibrium thermodynamics,J. Energy Resour. Technol. 123 (3) (2001) 2001.

[18] Y. Gao, Rate Controlled Constrained-Equilibrium Calculations of FormaldehydeOxidation, Ph.D. Thesis, Northeastern University, Boston, 2003.

[19] S. Ugarte, Y. Gao, H. Metghalchi, Int. J. Therm. 3 (2005) 21.[20] L. Petzold, SIAM J. Sci. Stat. Comput. 3 (1982) 367.[21] <http://www.me.berkeley.edu/gri-mech/version30/text30.html>.[22] W. Tsang, R.F. Hampson, J. Phys. Chem. Ref. Data 15 (3) (1986) 1087–1193.[23] W.C. Reynolds, STANJAN Program, Stanford University, ME270, HO#7.