You are on page 1of 60
OVERVIEW chapter 2 Combustion and Thermochemistry In this chapter, we examine several thermodynamic concepts that are impor: tant in the study of combustion, We first briefly review basic property relations for ideal gases and ideal-gas mixtures and the first law of thermodynamics, Although these concepts are likely to be familiar to you from a previous study of thermodynamies, we present them here since they are an integral part of our study of combustion. We next focus on thermodynamic topies related specifi- cally to combustion and reacting systems: concepts and definitions related 10 element conservation; definition of enthalpy that accounts for chemical bonds: and first-law concepts defining heat of reaction, heating values, ete and adiabatic flame temperatures. Chemical equilibrium, a second-law con. cept, is developed and applied to combustion-product mixtures. We emphasize equilibrium because, in many combustion devices, a knowledge of equilibrium states is suflicient to define many performance parame example, the temperature and major species at the outlet of a st combustor are likely to be governed by equilibrium considerations. § ers of the d examples are presented to illustrate these principles, REVIEW OF PROPERTY RELA Extensive and Intensive Properties ‘The numerical value of an exter ive property depends on the amount (mass or number of moles) of the substance considered. Extensive properties are uswally denoted with capital letters; for example, V (mm") for volume, U (J) for internal ° 10 CHAPTER? © Combustion and Thermochemistry energy, H (J) (= U + PY) for enthalpy, ete. An intensive property, on the other hand, is expressed per unit mass (or per mole), and its numerical value is independent of the amount of substance present. Mass-based intensive proper- ties are generally denoted with lower-case letters; for example, v (m’ /kg) for specific volume, u (J/kg) for specific internal energy, / (kg) (= + Po) for specific enthalpy, etc. Important exceptions to this lower-case convention are the intensive properties temperature T and pressure P. Molar-based intensive properties are indicated in this book with an overbar, e.g. i and h (J/kmol). Extensive properties are obtained simply from the corresponding intensive properties by multiplying the property value per unit mass (or mole) by the amount of mass (or number of moles); ie., V = mo(or NB) ey U = mu(or Nid) H = mh(or Nh), ete. In the following developments, we will use either mass- or molar-based inten- sive properties, depending on which is most appropriate to a particular situa tion. ‘Equation of State ‘An equation of state provides the relationship among the pressure, P, tempera~ ture, T, and volume ¥ (or specific volume v) of a substance, For ideul-gas behavior, ie., a gas that can be modeled by neglecting intermolecular forces and the volume of the molecules, the following equivalent forms of the equa- tion of state apply. PV =NR,T, (2.2a) PV =mRT, (2.26) Po=RT, 2.20) or P= pRT, (2.24) where the specific gas constant R is related to the universal gas constant R, (= 8315 J/kmol-K) and the gas molecular weight MW by R=R,/MW. 3) ‘The density p in Eqn. 2.2d is the reciprocal of the specific volume (p = 1/0 = ‘m{V). Throughout this book, we assume ideal-gas behavior for all gascous species and gas mixtures, This assumption is appropriate for nearly all of the systems we wish to consider since the high temperatures associated with com- bustion generally result in sufficiently low densities for ideal-gns behavior to be ‘a reasonable approximation. Review of Property Relations Calorific Equations of State Expressions relating internal energy (or enthalpy) to pressure and temperature are called calorific equations of state, ie., u=u(T,v) 4a) h= WTP) (2.4) ‘The word “calorific” relates to expressing energy in units of calories, which has been superseded by the use of joules in the SI system. General expressions for a differential change in u or h ean be expressed by differentiating Eqns. 2.4a and by v=() ar +(%) ex) 8) ora ( , = (28) or (3), a» on In the above, we recognize the partial derivatives with respect to temperature to be the constant-volume and constant-pressure specific heats, respectively, ie., a= (2), 2) (zz), (2.6b) For an ideal gas, the partial derivatives with respect to specific volume, (@u/e),, and pressure, (3h/3P)z, are zero. Using this knowledge, we integrate Eqn. 2.5, substituting Eqn. 2.6 to provide the following ideal-gas calorific equations of state: r MP)~ te =| ear (74) WT)— hoe = | AT. 2.76) i In a subsequent section, we will define an appropriate reference state th: ‘accounts for the different bond energies of various compounds. For both real and ideal gases, the specific heats ¢, and c, are generally functions of temperature, This is a consequence of the internal energy of a molecule consisting of three components: translational, vibrational, and rota- tional; and the fact that, as a consequence of quantum theory. the vibrational aand rotational energy storage modes become increasingly active as temperature increases. Figure 2.1 schematically illustrates these three energy storage modes by contrasting a monatomic species, whose internal energy consists solely of translational kinetic energy, and a diatomic molecule, which stores energy in a 2 = (@) Monatomie speces IN Pre (© Diatomie species Figure 2.1 (a} Tha internal enargy of monatomic species consists only ofronslotional {Gina energy, stile (4) 0 cictomie species’ ternal energy resus from translation together ‘wth energy fram vibration (potential and Kinetic) ond rotoion [kinetic vibrating chemical bond, represented as a spring between the two nuclei, and by rotation about (wo orthogonal axes, as well as possessing kinetic energy from translation. With these simple models (Fig. 2.1), we would expect the specific heats of diatomic molecules to be greater than monatomic species. In general, the more complex the molecule, the greater its molar specific heat. This can be seen clearly in Fig. 2.2, where molar specific heats for a number of ‘combustion product species are shown as functions of temperature. As a ‘group, the triatomics have the greatest specific heats, followed by the dia- tomies, and, lastly, the monatomics. Note that the triatomic molecules also have a greater temperature dependence than the diatomics, a consequence of the greater number of vibrational and rotational modes that are available to become activated as temperature is increased. In comparison, the monatomic species have nearly constant specific heats over a wide range of temperatures: in fact, the H-atom specific heat is constant (, = 20.786 kl/kmol-K) from 200K to S000K. Constant-pressure molar specific heats are tabulated as a function of tem- perature for various species in Tables A.1 to A.12 in Appendix A. Also pro- Vided in Appendix A are the curvefit coefficients, taken from the Chemkin thermodynamic database [1], which were used to generate the tables. These coefficients can be easily used with spreadsheet software to obtain é, values at any temperature within the given temperature range. Review of Property Relations P os | 4,0 me zo = 0, Lo g 2 co f & i pope} 2 H i 0 % 1000 2000, 3000 “4000 5000 Temperature (K) Figure 2.2 Moler conston-presture speci heals as functions of temperature for monatomic (H, N, and O}, dctomic (CO, Mp, and 3), ond tialomic (Cz, #0, and NOs) Species. Values ore Irom Appendix A. Ideal-Gas Mixtures Two important and useful concepts used to characterize the composition of a mixture are the constituent mole fractions and mass fractions. Consider ‘multicomponent mixture of gases composed of Ny moles of species 1, Ny moles of species 2, etc. The mole fraction of species i, xj, is defined as the fraction of the total number of moles inthe system that are species i i. N N, VSN FN+. N+. Max ea) x Similarly, the mass fraction of species i, ¥j, is the amount of mass of species / compared with the total mixture mass: 3 “4 CHAPTER? © Combustion and Thermochemistry my mu amy + ty +. my Mo 29) Note that, by definition, the sum of all the constitutent mole (or mass) fractions ‘must be unity, ic., (2.104) De Ds (2.106) Mole fractions and mass fractions are readily converted from one to ‘another using the molecular weights of the species of interest and of the mixture: Y= 1MW/MW Qutay 1= YMWyig/ MW uitby “The mixture molecular weight, MH’x:.is easily calculated from a knowledge of either the species mole or mass fractions: MW gue = SMW, 2.12) M ! (2.120) a “von Species mole fractions are also used to determine corresponding species partial pressures. The partial pressure of the ith species, P;, isthe pressure of the ith species if it were isolated from the mixture at the same temperature and volume as the mixture, For ideal gases, the mixture pressure is the sum of the constituent partial pressures: Parr, 2.13) ‘The partial pressure can be related to the mixture composition and total pressure as P= XP. (2.14) For ideal-gas mixtures, many mass- (or molar-) specific mixture properties are calculated simply as mass (or mole) fraction weighted sums of the individ- ual species-specific properties. For example, mixture enthalpies are calculated as Review of Property Relations Froie = 2 Yoh (.15a) fing = Dh 2.150) Other frequently used properties that can be treated in this same manner are internal energies, wand i. Note that, with our ideal-gas assumption, neither the Pure-species properties (i. /,) nor the mixture properties depend on pres sure. ‘The mixture entropy also is calculated as a weighted sum of the constitu cents: Said. P) = YT, Pi) 2.16) Spis( TP Laser, Pa, 2.166) In this case, however, the pure-species entropies (s, and §)) depend on the cs partial pressures as indicated in Eqn. 2.16, The constituent entropies in Eqn. 2.16 can be evaluated from standard-state (Ppp = P’ = Jam) values as P, maa STP) = S(T Pog) — Rn ru (17a) 5A, P) = SAT, Pog) = Ry ln 5 2.176) Poot Standard-state molar specific entropies are tabulated in Appendix A for many species of interest to combustion, Latent Heat of Vaporization In many combustion processes, a liquid-vapor phase change is important. For example, a liquid fuel droplet must first vaporize before it can burn: and, if cooled sufficiently, water vapor can condense from combustion products Formally, we define the latent heat vaporization, hy,, as the heat required in a constant-pressure process to completely vaporize a unit mass of liquid at a -n temperature, ie., gl TP) % Inapod(Ts PY ~ Dish TP), (2.18) where 7 and P are the corresponding saturation temperature and pressure, respectively. The latent heat of vaporization is also known as the enthalpy of vaporization. Latent heats of vaporization for various fuels at their normal (atm) boiling points are tabulated in Table B.1 (Appendix B). The latent heat of vaporization at a given saturation temperature and pressure is frequently used with the Clausius-Clapeyron equation to estimate saturation pressure variation with temperature 16 CHAPTER 2 © Combustion and Thermochemistry Pans _ Iie ATo Pa R Ta This equation assumes that the specific volume of the liquid phase is negligible compared with that of the vapor and that the vapor behaves as an ideal gas. ‘Assuming fig is constant, Eqn. 2.19 can be integrated from (Pyy.1, Ts.1) t© (Psa: Tsus,2) in order to permit, for example, Pyy.2 to be estimated from a knowledge of Pyy,1.Taux,1+ and Tyy.2. We will employ this approach in our discussion of dropiet evaporation (Chapter 3) and combustion (Chapter 10). (2.19) FIRST LAW OF THERMODYNAMICS First Law—Fixed Mass Conservation of energy is the fundamental principle embodied in the first law of thermodynamics. For a fixed mass, ic., a system, (Fig. 2.3a), energy con- servation is expressed for a finite change between two states, | and 2, as 1Q> = Wy = AEi-2 (2.20) Head added v9 Work done by stm Cange in ttn ayer syst in ging ‘on urrundig in Bons ‘ry ging rom fromaste Neowate2 from state fo sate? ‘ane Tose? Both ,@; and ; W are path functions and occur only at the system boundaries; AEj-x(@ Ez — Ej) is the change in the total energy of the system, which is the sum of the internal, kinetic, and potential energies, ie., E=m( wu + ie + ge). 21) Maseapesie nen, Masespecie tem Mass specie sstem Imetal ents wu ene posal eer) we athe aterm Laem mis de Po oahse ° Figure 2.3 _{o) Schematic of fxed-mass sysiom with moving boundary above piston. (b) ‘Conta volume with fixed boundries ond steody flow. First Law of Thermodynamics ‘The system energy is a state variable and, as such, AE does not depend on the path taken to execute a change in state. Equation 2.20 can be converted to unit ‘mass basis or expressed to represent an instant in time. These forms are 12 — 12 = Ae =e; ~e, 22) and o - w = dE/ar 2.23) Phen nse eaten one, "ric of change ‘nto stem power ‘stm ont G- = de/dt, 2.24) where lower-case letters are used to denote mass-specific quantities, e.., e= Em. First Law—Control Volume We next consider a control volume, illustrated in Fig. 2.3b, in which uid may flow across the boundaries. The steady-state, steady-flow (SSSF) form of the first law is particularly useful for our purposes and should be reasonably familiar to you from previous studies of thermodynamics [2-4]. Because of its importance, however, we present a brief discussion here. The SSSF first law is expressed as On = te, fie, $i Pamy — Ped, ieaecey. ra oa 2.28) where the subscripts o and i denote the outlet and inlet, respectively, and sis the mass flowrate. Before rewriting Eqn. 2.25 in a more convenient form, itis appropriate to list the principal assumptions embodied in this relation. 1 The control volume is fixed relative to the coordinate system, This eliminates any work interactions associated with a moving boundary, as well as climinating the need to consider changes in the kinetic and potential energies of the control volume itself 2. The properties of the fluid at each point within the control volume. or on the control surface, do not vary with time. This assumption allows us to treat all processes as steady. v CHAPTER? —« Combustion and Thermochemistry 3, Fluid properties are uniform over the inlet and outlet flow areas. This allows us to use single values, rather than integrating over the area, for the inlet and exit stream properties. 4. There is only one inlet and one exit stream. This assumption is invoked to keep the final result in a simple form and can be easily relaxed to allow multiple inlet/exit streams. ‘The specific energy e of the inlet and outlet streams consists of the specific internal, kinetic, and potential energies, ic., e = « + f (2.26) Tlulemny Intra enrry Kine egy peunemee —ferinkims rants where v and z are the velocity and elevation, respectively, of the stream where it crosses the control surface. “The pressure-specifie volume product terms associated with the flow work in Eg, 2.25 can be combined with the specific internal energy of Eqn. 2.26, which we recongize as the useful property, enthalpy: ut Posut Pip. 27) Combining Eqns. 2.25-2.27, and rearranging, yields our final form of energy conservation for a control volume. Geo — Wea = rll — A) +40 — ¥) + 8G ~ 2d] (2.28) ‘The first law can also be expressed on a mass-specific basis by dividing Eqn. 2.28 by the mass flowrate sit, ic, hy ~ by 402 — 9) + eC — 20. (229) oo Wen In Chapter 7, we present more complete expressions of energy conservation that are subsequently simplified for our objectives in this book. For the time being, however, Eqn. 2.28 suits our needs. REACTANT AND PRODUCT MIXTURES Stoichiometry The stoichiometric quantity of oxidizer is just that amount needed to com- pletely burn a quantity of fuel. If more than a stoichiometric quantity of oxidizer is supplied, the mixture is said to be fuel lean, or just lean; while supplying less than the stoichiometric oxidizer results in a fuel-rich, or rich mixture. The stoichiometric oxidizer- (or air-) fuel ratio (mass) is determined by writing simple atom balances, assuming that the fuel reacts to form an ideal Reactant and Product Mixtures set of products. For a hydrocarbon fuel given by C,H, the stoichiometric relation can be expressed as CH, + a(O3 + 3.76N2) + xCOy + (9/2)H20 + 3.760N: (230) where x+y/4, 231) For simplicity, we assume throughout this book that the simplified composi- tion for air is 21 percent O, and 79 percent N3 (by volume), ie, that for each ‘mole of O) in air, there are 3.76 moles of Nz, The stoichiometric air-fuel ratio can be found as /Page = (Mit) = where MW, and MWg are the molecular weights of the air and fuel, respec- tively. Table 2.1 shows stoichiometric air-fuel ratios for methane and solid carbon. Also shown is the oxygen-fuel ratio for combusion of Hy in pure >. For all of these systems, we see that there is many times more oxidizer than fuel, (2.32) Table 2.1 __ Some combustion propertos of methane, hydrogen, and solid carbon for reactors a! 298 K ‘he ‘he OP? (SPhgud (shan) (esik) CH + air 55.538 —3,066 Wat Hy + Oy 142919 ~15380 80 cy) 32794 2,648 na 0 /F i ho exif aio, where for combustion wth ot, he ois he onder not jt the oxygen in thea The equivalence ratio, @, is commonly used to indicate quantitatively whether a fuel-oxidizer mixture is rich, lean, or stoichometric. The equivalence ratio is defined as = APP asic (FV) OSE) Fac (2.33a) From this definition, we see that for fuel-rich mixtures, © > 1, and for fuel- lean mixtures, @ < 1, For a stoichiometric mixture, & equals unity. In many combustion applications, the equivalence ratio is the single most important factor in determining a system’s performance. Other parameters frequently used to define relative stoichiometry are percent stoichiometric air, which relates to the equivalence ratio as ” 20 CHAPTER 2 © Combustion and Thermochemistry 1002 % stoichometric air = 207% (2.336) and percent excess air, or % excess air =F = 100%, 0339 Example 2.1 A small,low-cmisson, stationary gusturbine engine (ae Fig. 24) operates at full ad {G950W) at an equivalence ratio oF 0.286 with an air flowrate of 15.9 kp. The equiv lent composition of the fel atual gus) i C Hy- Determine the fuel mas flowrate and the operating air-fuel ratio for the engine Solution Given: @ = 0286, MIs, = 2885, fing = 159Kgl6,— MVng = 1.161201) + 4.321.008) = 18.286 Find: Fas and (4/8). We will prooced by first finding (4/F) and then ig. The solution requires only the application of definitions expresed in Eqns. 2.32 and 2.33, ie, where a= x-+y/4-= 1.16-4432/4 = 224. Thus, 2a A/Puge = 4760.28 788 _ 63, ad, from Ean. 2.33, (A/F aic _ 1682 CaF = = oe = OE Since (A/F) is the ratio of the air lowrate tothe fuel flowrate, 159ks/s Ti] = tts = BOKS . [OaTO RR Comment Note that even at fll power, a large quantity of exes airs supplied tothe engine. Example 2.2 A natural gassed industrial boller (oe Fig. 2.) operates with an oxygen concentration of 3 mole percent in the flue gases. Determine the operating air-fuel ratio and the ‘equivalence ratio. Treat the natural gas as methane. Solution Given: yo, = 0.03, MW = 16.04, MW yg = 28.85, Find: (4/F) and ©. ‘We can use the given O2 mole fraction to find the air-fuel ratio by writing an overall combustion equation assuming “complete combustion,” ie. no dissociation (all fuel C is found in CO, and al fuel H is found in HO): CHAPTER? wider Injection fel spoke (18) Primary airflow ° Channel fuel 3 injection Pile fuel Reactant and Product Mixtures 2 Figure 2.5 Two 10:MW (34 milion BTU/hr) natural-gos burners fre into © boil combustion chamber 3m deep. Air enters the burners through the forge verhcel pipes, wile ‘he natural gos enters through the herizonta! pipe onthe let T'SOURCE: Cousesy of North Amercan Manufataing Co (CH, 4 a(0; + 3.76N;) + COs +2H,0 +502 +3.76aN3, here wand b are related from conservation of © atoms, +242 baa-2 From the definition of a mole fraction (Ean. 2.8) No, 6 a 20: = Nig TED TBE Ta Ta Substituting the known value of zo,(= 0.03) and then solving for a yields est: Tra a= 2.368, ery CHAPTER? «Combustion and Thermochemistry ‘The mass air-fuel rato, in general, is expressed as (arm = 2280289 fg To find ©, we need to determine (A/F)... From Eq. 231, a= 2; hence, Poa BREE <7 Applying the definition of (Eqn. 2.33), APPace _ VTL ap, on Comment In the solution, we assumed that the ©; mole fraction was on a “wet basis,” ie. moles of O2 per mole of moisture-containing flue gases. Frequently, in the measurement of exhaust species, moisture is removed to provent condensation in the analyzers; thus, xo, cean also be reported on a “dry basis” (see Chapter 15). Absolute (or Standardized) Enthalpy and Enthalpy of Formation In dealing with chemically reacting systems, the concept of absolute enthalpies is extremely valuable. For any species, we can define an absolute (or standar- dized) enthalpy that is the sum of an enthalpy that takes into account the energy associated with chemical bonds (or lack thereof), the enthalpy of formation, i, and an enthalpy that is associated only with the temperature, the sensible ‘enthalpy change, f,. Thus, we can write the molar absolute enthalpy for species i as iT) = (Tet) + Abyi(Trt), (2.34) Abst eilny otha of formation Sense enalpy ‘Wenger ‘a sandard tense hangs nn rom state Tar oF where Alig, = f(T) ~ lif (Tr) To make practical use of Eqn. 2.34, it is necessary to define a standard reference state. We employ a standard-state temperature, Tyg = 25°C (298.15K), and. standard-state pressure, Pr = P? = | atm (101,325 Pa), con- sistent with the Chemkin [1] and NASA [5] thermodynamic databases. Furthermore, we adopt the convention that enthalpies of formation are zero for the elements in their naturally occurring state at the reference state Reactant and Product Mixtures temperature and pressure. For example, at 25°C and Latm, oxygen exists as diatomic molecules; hence, (io, where the superscript 0 is used to denote that the value is for the standard-state pressure. To form oxygen atoms at the standard state requires the breaking of a rather strong chemical bond. The bond dissociation energy for O) at 298 K is 498,390 kJ/kmolo,. Breaking this bond creates two O atoms; thus, the enthalpy of formation for atomic oxygen is half the value of the O; bond dissociation energy, ie, BS io) =249.19543/Amolg ‘Thus, enthalpies of formation have a clear physical interpretation as the net change in enthalpy associated with breaking the chemical bonds of the stan- dard state elements and forming new bonds to create the compound of interest. Representing the absolute enthalpy graphically provides a useful way to understand and use this concept, In Fig. 2.6, the absolute enthalpies of atomic oxygen (0) and diatomic oxygen (02) are plotted versus temperature starting from absolute zcro. At 298.15K, we see that fi, is zero (by definition of the standard-state reference condition) and the absolute enthalpy of atomic oxygen equals its enthalpy of formation, since the sensible enthalpy at 298.15 K is zero, AC the temperature indicated (4000 K), we see the additional sensible enthalpy contribution to the absolute enthalpy. In Appendix A, enthalpies of formation 400,000 300 00] O stom Ah, (4000) 200,000 (4000 K) 100.000} | 0; molecules, i Absolute enthalpy (Km 100,000 '0 10003000 3000 #000060 000 Temperature (K) Figure 2.6 Grophicol interpretation of ebsolute enthelpy, heat of formation, ond sensible halpy. 5 % CHAPTER? —» Combustion and Thermochemistry at the reference state are given, and sensible enthalpies are tabulated as a function of temperature for a number of species of importance in combustion Enthalpies of formation for reference temperatures other than the standard state 298.15K are also tabulated. Example 2.3, ‘A gas stream at | atm contains a mixture of CO, COs, and Nz in which the CO mole fraction is 0.10 and the CO; mole fraction is 0.20. The gas-stream temperature is 1200K. Determine the absolute enthalpy of the mixture on both @ mole basis (ki/ 4kmol) and a mass basis (kJ/kg). Also determine the mass fractions of the three compo- nent gases. Solution Given: co = 0.10, T= 1200K, co, = 0.20, P=latm Find: finies bass Yoos Yeo, and Yw, Finding ligix Tequires the straightforward application of the ideal-gas mixture law, Eqn. 2115, and, to find zy, the knowledge that Ey, = 1 (Eqo, 210. Thus, an, = 1 = xe0, = 0.70 and fix Buh = xe0ffiico + (i ~Ft8)c0] + xe0,[liico, + (HT) — F.28)c0,] + w[fin, + (HD — Han), ‘Substituting values from Appendix A (Table A.1 for CO, Table A.2 for CO, and Table A for N:} 1o[-110,541 + 28,440] + 0.20[-393,546 + 44,488] + 0:70{0 + 28.118) 58,339.11 /kinolys Lins To find we need to determine the molecular weight of the mixture: Main = BKM, 10(28.01) + 0.20(44.01) + 0.7028.013) 1.212 Then, = fn 58391 ‘nin = Wag 31.212 Since we have previously found MWyyx, calculation of the individual mass fractions follows simply from their definitions (Eqn. 2.11) = 1869.12 K0/Keaa) Reactant and Product Mixtures 3801 10 5573 — 0.0897 39 401 20 s5 = 02820) ‘As a check, we see that 0.0897 + 0.2820 + 0.628 Comment 1.000, as required. Both molar and mass units are frequently used in combustion. Because of this, you should be quite comfortable with their interconversions, 27 Enthalpy of Combustion and Heating Values Knowing how to express the enthalpy for mixtures of reactants and mixtures of products allows us to define the enthalpy of reaction, or, when dealing specif- ically with combustion reactions, the enthalpy of combustion. Consider the steady-flow reactor, shown in Fig. 2.7, in which a stoichiometric mixture of reactants enters and products exits, both at standard-state conditions (25°C, | atm), The combustion process is assumed to be complete, ic. all of the fuel ‘carbon is converted to CO and all of the fuel hydrogen is converted to H;O. For the products to exit at the same temperature as the entering reactants, heat ‘must be removed from the reactor, The amount of heat removed can be related to the reactant and product absolute enthalpies by applying the steady-flow form of the first law (Eqn. 2.29) hy- 4 gros tree 2.33) The definition of the enthalpy of reaction, or the enthalpy of combustion, Aha (per mass of mixture, is Alin = deo = hot ~ Pracs (2.36) Ressants —>! 1+ rotate (echionerie J” (complete conbusion folate steundad ate 2 standard state conditions) tendons) Figure 2.7 Steady-low reactor used to determine enthalpy of combustion CHAPTER? © Combustion and Thermochemistry or, in terms of extensive properties, Ag = pros ~ Hac (2.366) The enthalpy of combustion can be illustrated graphically, as shown in Fig. 2.8. Consistent with the heat transfer being negative, the absolute enthalpy of the products lies below that of the reactants, For example, at 25°C and I atm, the reactants enthalpy of a stoichiometric mixture of CHy and air, where | mol of fuel reacts, is 74,831 kJ. At the same conditions (25°C, latm), the combustion products have an absolute enthalpy of —877,236KJ. Thus, AHg = ~877,236 ~ (74,831) = 802,405 kJ This value can be adjusted to a per-mass-of-fuel basis: Aig H-) = Hg M Mt 237) Fi on(@e,) = (-802,405/16.043) = —s0,016. (0) Enihelpy of reaction using representative values foro stoichiometric mathons— ‘The water inthe products i ossumed fo bein the vopor sate Reactant and Product Mixtures ee ee 1) Ahyi| —) —, (2.38) (ia) y where (239) (ATFV+1 From Table 2.1, we see that the stoichiometric air-fuel ratio for CHy is 17.11; kJ —50,016 Note that the value of the enthalpy of combustion depends on the temperature chosen for its evaluation since the enthalpies of both the reactants and products vary with temperature; ic., the distance between the Hyog and Hae lines in Fig. 2.8 is not constant. The heat of combustion, Ai, (known also as the heating value). is numeri- cally equal to the enthalpy of reaction, but with opposite sign, The upper oF higher heating value, HHL, is the heat of combustion calculated assuming that all of the water in the products has condensed to liquid. This scenario liberates the most amount of energy, hence the designation “upper.” The lower heating value, LHV, corresponds to the ease where none of the water is assumed to condense. For CH, the upper heating value is approximately 11 percent larger than the lower. Standard-state heating values for a variety of hydrocarbon fuels are given in Appendix B. » A. Determine the upper and lower heating values at 298K of gaseous decane, Ciobz:. per kilomole of fuel and per kilogram of fuel. The molecular weight of 2- decane is 142.284, B. If the enthalpy of vaporization of n-decane i 359 kJ/kgiua at 298K, what are the upper and lower heating values of liquid n-decane? le of C\oH,2, the combustion equation can be written as CyoHan(e) + 15.5(0s +3.76N2) > 10CO; + HHsOU or g) + 15.503.76)N> For cither the upper or lower heating value, AH, = —8 Ha = Hex ~ Hoos where the numerical value of Hpyog depends on whether the HO in the products is Tiquid (determining higher heating value) or gaseous (determining lower heating value). The sensible enthalpies for all species involved are zero since we desire ‘Ad, at the reference state (298 K). Furthermore, the enthalpies of formation of the (0; and Ny are also zero at 298K. Recognizing that Example 2.4 CHAPTER? « Combustion and Thermochemistry Moe = Ni 208 Hyg = NI we obtain AH 4,09 = HAY = (i cy ~ [ ico, + Wino} Table A.6 (Appendix A) gives the enthalpy of formation for gaseous water and the enthalpy of vaporization. With these values, we can calculate the enthalpy of Formation for the liquid water (Eqn. 2.18): Feo) = ano ~ hy = ~241.847 ~ 44.010 = 285,857 3k. ‘Using this value, together with enthalpies of formation given in Appendiees A and B, we obtain the higher heating value: A n(-2960 2) wu io = [to(-9 5 1) (ane #2) $30,065) AH.a,00) and 820,096 moe, ah__ 6830096 5 Wega 4.286 (ee, 800307 For the lower heating value, we use Mf. x.o%) Hy sou) = ~285,857kI/kmol. Thus, 41,847 kI/kmol in place of Bi, = 6S ROT mol ah. 001 ke, B. For CigHz, in the liquid state, Feats ~ ha) an (iat) — a (PH) — Reactant and Product Mixtures o { 3 | é | 3 = » be no (Gare ct / nes Not by Ss Temperature Figure 2.9 Entholpy-temperature plot ilustating calculation of heating volves in Exomple 2a Thus, |, (higher) = 48,003 ~ 339, 7.644 KI Pk Ah, (lower) = 44,601 — 359 I 242KI Ker Comment Graphical representations of the various definitions and/or thermodynamic processes are valuable aids in setting up problems or in checking their solutions. Figure 29 illustrates, on fT coordinates, the important quantities used in this example. Note that the enthalpy of vaporization given for n-decane is for the standard-state tempera ture (298.15), while the value given in Appendix B is at the boiling point (447.4 k). a 32 CHAPTER? © Combustion and Thermochemistry ADIABATIC FLAME TEMPERATURES We define two adiabatic flame temperatures: one for constant-pressure com- bustion and one for constant-volume. Ifa fuel-air mixture burns adiabatically at constant pressure, the absolute enthalpy of the reactants at the initial state (say, T = 298K, P = 1atm) equals the absolute enthalpy of the products at the final state (7 = Typ, P= Latm), ie., application of Eqn. 2.28 results in Heel Ty P) = Hy TatsP) (2.400) ‘or, equivalently, on a per-mass-of-mixture basis, Iga Tis P) = Igo Tats Pe (2.406) ‘This first-law statement, Eqn. 2.40, defines what is called the constant-pressure adiabatic flame temperature. This definition is illustrated graphically in Fig. 2.10. Conceptually, the adiabatic flame temperature is simple; however, eval- tating this quantity requires knowledge of the composition of the combustion products. At typical flame temperatures, the products dissociate and the (beg) Figure 2.10 _Illstration of constont-pressute adiabatic Fame temperature on h-T coordinotes. Adiabatic Flame Temperatures mixture comprises many species. As shown in Table 2.1 and Table BI in Appendix B, flame temperatures are typically several thousand kelvins Calculating the complex composition by invoking chemical equilibrium is the subject of the next section. The following example illustrates the fundamental concept of constant-pressure adiabatic flame temperatures, while making crude assumptions regarding the product mixture composition and evaluation of the product mixture enthalpy. 33 Estimate the constant-pressure adiabatic lame temperature for the combustion of a stoichiometric CHy-air mixture, The pressure is | atm and the initial reactant tempera- ture is 298K Use the following assumptions “Complete combustion” (no dissociation), ic. the product mixture consists of only CO;, H,0, and Ns, 2. The product minture enthalpy is estimated using constant specific heats evaluated at 1200 K (&0.5(T, + Ty), where Tyy is guessed to be about 210K). Solution Mixture composition: (CH +210 + 3.76N3) —> 1 CO} + 24:04 7.52N, Properties (Appendices A and B) Enthalpy of Formation @ 298K Specific Heat a) 1200K Species 4, alae “yu (ka/Amotk) cH — 74881 ©o: 393 6 soa ho 241 sas ae ™ ° m1 ° ° First law (Eqn. 2.40) Fro yes = Mihi Hyege, = (IX —74,831) + 20) + 7.52(0) = =74831 43 peod = DNAs + Gy (Tas ~ 298] {393,546 + 56.21( Tay ~298)) + (Q-241,845 + 43.877, — 298)] + G520-4 38.71(Tay ~ 298). Equating Mec 10 Hyg and solving for Tuy yields Toy = 2318K] Example 2.5 CHAPTER? + Combustion and Thermochemistry Comments Comparing the above result vith the equiirium-comporition based computation shown in Table 2.1 (Tguq = 2226) shows that the simplified approach overestimates Tra by slightly ess thas 10DK. Considering the crudeness of the assumptions, this appears to be rather surprisingly good agreement. Removing assumption 2 and re- ululating Tyy using variable specific heat, i, fi | an. yields Tyy = 2328 K. (Note that Appendix A provides tabulations of these integrated ‘quantities. Similar tabulations are found in the JANAF tables (6}) Since this result is quite close to our constant-c, solution, we conclude that the ~ 100K difference is the result of neglecting dissociation. Note that dissociation causes a lowering of Tay since ‘more energy is tied up in chemical bonds (enthalpies of formation) at the expense of the sensible enthalpy. In the above, we dealt with a constant-pressure system, which would be appropriate in dealing with a gas-turbine combustor, or a furnace. Let us look ‘now at constant-volume adiabatic flame temperatures, which we might require in aan ideal Otto-cycle analysis, for example. The first law of thermodynamics (Eqn. 2.20) requires reac (Ti, Pini) = Uproa(Taas Ps at) where U is the absolute (or standardized) internal energy of the mixture. Graphically, Eqn. 2.41 resembles the sketch (Fig. 2.10) used to illustrate the constant-pressure adiabatic flame temperature, except the internal energy replaces the enthalpy. Since most compilations or calculations of thermo- dynamic properties provide values for H (or h) rather than U (or w) (1, 6), ‘we can rearrange Eqn, 2.41 to the following form: Hae ~ Hest ~ VPs ~ P) = 0 (2.42) We can apply the ideal-gas law to eliminate the PY terms: Paut¥ = YO NRuTinit = NreweRe Tine P/V = NR Tat = Novos Re Tate fel Thus, Hae ~ Hyros ~ RalNreac Trt ~ Nproa Tat) = 0- 243) ‘An alternative form of Eqn. 2.43, on a per-mass-of mixture basis, can be obtained by dividing Eqn. 2.43 by the mass of mixture, mai, and recognizing that Mic! Nreac = MW rene Adiabatic Flame Temperatures or main] Npeos = MW pro We thus obtain fr, r, Seat Ru spi — aif) =°- aay re NM MW pt Since the equilibrium composition of the product misture depends upon both temperature and pressure, as we will se inthe next section, ullizing Ean. 243 or 2.44 with the ideal-gas law and appropriate calorific equations of state, eg MT, P) =h (T only, ideal gas), to find T,y is straightforward, but non. Estimate the constant-volume adiabatic flame temperature for a stoichiometric CH,_air ‘mixture using the same assumtions as in Example 2. Initial conditions are 7; = 298K, ‘atm (= 101,325 Pa). Solution ‘The same composition and properties used in Example 2.5 apply here, We note, how= ever, that the ¢,,; values should be evaluated at a temperature somewhat greater than 1200K. since the constant-volume Tj will be higher than the constant-pressure Tyy Nonetheless, we will use the same values as before. First law (Ego. 2.43) Hrcac~ Hyaos ~ Bu(Neoe Tine ~ Nprod Tad) = DENA ~ RiNrTi ~ Novos To Substituting numerical values, we have He (74.831) + 2(0) + 7.520) 74831 KI Hyg = (IN ~393,546 + $6.21(T yy ~ 298)] + Q)[-241 845 + 43.87(Tag — 298)] + .52)0-4 33.71(Tas ~ 298)] = 877,236 + 397.5(Ty ~ 298)kI and Ry(NrweTint ~ Nod Tad) = 8:315(00.52)298 = where Neus = Npyoe = 10.52 kml. Reassembling Eqn. 2.43 and solving for Tay Example 2.6 36 CHAPTER? 6 Combustion and Thermochemistry Comments (For the same initial conditions, constant-volume combustion results in much higher temperatures (571K higher in this example) than for constant-pressure combustion. This isa consequence of the pressute forces doing no work wien the volume is fixed Gi) Note, aso, that the number of moles was conserved in going from the inital to final state. This fortuitous esult for CH, and does not occu for othe fuel. ii) The inal pressure is well above the inital pressure: Py = Pig( T/T) = 9.69atm CHEMICAL EQUILIBRIUM In high-temperature combustion processes, the products of combustion are not a simple mixture of ideal products, as may be suggested by the simple atom- balance used to determine stoichiometry (ef. Eqn. 2.30). Rather, the major species dissociate, producing a host of minor species. Under some conditions, what ordinarily might be considered a minor species is actually present in rather large quantities. For example, the ideal combustion products for burn- ing a hydrocarbon with air are CO;, HzO, O2, and N;, Dissociation of these species and reactions among the dissociation products yields the following species: Hp, OH, CO, H, O, N, NO, and possibly others. The problem we address in this section is the calculation of the mole fractions of all of the product species at a given temperature and pressure, subject to the constraint of conserving the number of moles of each of the elements present in the initial mixture. This element constraint merely says that the number of C, H, O, and N atoms is constant, regardless of how they are combined in the various species. ‘There are several ways to approach the calculation of equilibrium compo- sition. To be consistent with the treatment of equilibrium in most undergrad- uate thermodynamics courses, we focus on the equilibrium-constant approach and limit our discussion to the application of ideal gases. For descriptions of other methods, the interested reader is referred to the literature (5, 7. Second-Law Considerations The concept of chemical equilibrium has its roots in the second law of thermo- dynamics. Consider a fixed-volume, adiabatic reaction vessel in which a fixed mass of reactants form products. As the reactions proceed, both the tempera- ture and pressure rise until a final equilibrium condition is reached. This final state (temperature, pressure, and composition) is not governed solely by first- law considerations, but necessitates invoking the second law. Consider the combustion reaction C0440; + COs 0.43) Chemical Equilibrium If the final temperature is high enough, the CO; will dissociate. Assuming the products to consist only of COz, CO, and Op, we can write [co+10)],, + [aco +eco+$ 04], (2.46) where a is the fraction of the CO) dissociated. We ean calculate the adiabatic flame temperature as a function of the dissociation fraction, a, using Eqn. 2.42. For example, with a= 1, no heat is released and the mixture temperature, pressure, and composition remain unchanged; while with a = 0, the maximum amount of heat release occurs and the temperature and pressure would be the highest possible allowed by the first law. This variation in temperature with a is plotted in Fig. 2.1) ‘What constraints are imposed by the second law on this thought experi- ment where we vary a? The entropy of the product mixture can be calculated by summing the product species entropies, ie. Doser where 1V; is the number of moles of species / in the mixture. The individual species entropies are obtained from Swis(Ty P) (1 -adico, taico +5, (2.47) aT P, 5 git T Rell pe (2.48) Hw +f TER) or Sy (RK) a OF or] oF 70 Fraction of CO, undiesociated, I~ Figure 2.11 tusrotion of chemical equilibrium for @fixed-mos isolated sytem, a 38 CHAPTER? «© Combustion and Thermochemistry where ideal-gas behavior is assumed, and P, is the partial pressure of the ith species. Plotting the mixture entropy (Eqn. 2.47) as a function ofthe dissocia- tion fraction, we see that a maximum value is reached at some intermediate value of a, For the reaction chosen, CO-+ $0, > COs, the maximum entropy occurs near 1 ~a@ = 0.5, For our choice of conditions (constant U, V, and m, which implies no heat or work interactions), the second law requires that the entropy change internal to the system ds>0. (2.49) Thus, we see that the composition of the system will spontaneously. shift toward the point of maximum entropy when approaching from either side, since dS is positive. Once the maximum entropy is reached, no further change in composition is allowed, since this would require the system entropy to decrease in violation of the second law (Eqn. 2.49). Formally, the condition for equilibrium can be written 4S)y,72m = 0. (2.50) In summary, if we fix the internal energy, volume, and mass ofan isolated system, the application of Eqn. 2.49 (second law), Eqn. 2.41 (first law), and Eqn. 2.2 (equation of state) define the equilibrium temperature, pressure, and chemical composition. Gibbs Function Although the foregoing was useful in illustrating how the second law comes into play in establishing chemical equilibirum, the use of an isolated (fixed- energy) system of fixed mass and volume is not particularly useful for many of the typical problems involving chemical equilibrium. For example, there is frequently a need to calculate the composition of a mixture at a give tempera- ture, pressure, and stoichiometry. For this problem, the Gibbs free energy, G, replaces the entropy as the important thermodynamic property. ‘As you may recall from your previous study of thermodynamics, the Gibbs free energy is defined in terms of other thermodynamic properties as HTS. @.s1) The second law can then be expressed as (4G)r,pym <9, 2.52) which states that the Gibbs function always decreases for a spontaneous, is0- thermal, isobaric change of a fixed-mass system in the absence of all work effects except boundary (P-dV) work. This principle allows us to calculate the equilibrium composition of a mixture at a given temperature and pressure. ‘The Gibbs function attains a minimum in equilibrium, in contrast to the max- ‘Chemical Equilibrium imum in entropy we saw for the fixed-energy and fixed-volume ease (Fig. 2.11) Thus, at equilibrium, (Dr.pm =0- (2.53) For a mixture of ideal gases, the Gibbs function for the jh species is given by Hirt RTP) 2.54) where gi is the Gibbs function of the pure species at the standard-state pressure (.e., P; = P") and P; is the partial pressure. The standard-state pres- sure, /”, by convention taken to be I atm, appears in the denominator of the logarithm term, In dealing with reacting systems, a Gibbs function of formation, is frequently employed: an- Y GAD) a (2.55) where the vate the stoichiometric coefficients ofthe clements requited to form ‘one mole of the compound of interest. For example, the coefficients are 1), = | and for forming a mole of CO from Oy and C, respectively. As with enthalpies, the Gibbs functions of formation of the naturally occurring ele- ‘ments are assigned values of zero at the reference state. Appendix A provides tabulations of Gibbs function of formation over a range of temperatures for selected species. Having tabulations of g7(7) as a function of temperature quite useful. In later calculations, we will need to evaluate differences in Zr between different species at the same temperature. These differences can be obtained easily by using the Gibbs function of formation at the temperature of interest, values of which are provided in Appendix A. Tabulations for over 1,000 species can be found in the JANAF tables (6. The Gibbs function for « mixture of ideal gases can be expressed as Gon = ONiGir = DON [ir + RT IPP) (2.56) where N; is the number of moles of the ith species. For fixed temperature and pressure, the equilibrium condition becomes AGnig = 0 2.57) Daniele + RTI / PY] + DN. dlaty + RTI /P] =0. 2.58) ‘The second term in Eqn. 2.58 can be shown to be zero by recognizing that (in P,) = dP,/P, and that 3>4P; = 0, since all changes in the partial pressures ‘must sum to zero because the total pressure is constant. Thus, GGnix =0 = SPAN [Sir + RLTIN(P,/P?)} (2.59) For the general system, where AA+bB4... @ eB 4/F+..., 2.60) 39 CHAPTER? © Combustion and Thermochemistry the change in the number of moles of each species is directly proportional to its stoichiometric coefficient, ie., dN, = —Ka 2.61) Ng = —«b Ng = +xe Np = 44h Substituting Eqn. 2.61 into Eqn. 2.59 and canceling the proportionality constant x, we obtain —alair + RT In(Pa/P*)] — bfér + RuT ln(Po/P)] — (2.62) ele, + RyT In(Pe/P)] + S[eh.r + RuT In(Pe/ PY) +... = 0. Equation 2.62 can be rearranged and the log terms grouped together to yield ~ (ear + fbr +... — Bh — ba —--.) (2.63) (PefPY «(Pr PY! ete. (Pq/PPY" (Py / Po? ete. ‘The term in parentheses on the leftchand-side of Eqn. 2.63 is called the standard-state Gibbs function change AG, i.e., AG, = (eBbr +487 +. — Bir — bBo —--) (2.64a) or, alternately, AGF = (CBf.n + /8e + 08} 4 ~ bg ---) (2.64b) ‘The argument of the natural logarithm is defined as the equilibrium constant K, for the reaction expressed in Eqn. 2.60, ie., (Po/P’Y «(Pri PY ete. (Px/PY (Pp / Poy ete. With these definitions, Eqn. 2.63, our statement of chemical equilibrium at constant temperature and pressure, is given by AG} =-RTInky, (2.668) =R,Tin K, 2.65) or K, = exp(-AG7/R.T). (2.666) From the definition of K, (Eqn. 2.65) and its relation to AG (Eqn. 2.66), we can obtain a qualitative indication of whether a particular reaction favors products (goes strongly to completion) or reactants (very little reaction occurs) ‘Chemical Equilibrium at equilibrium. If AG; is positive, reactants will be favored since In K, is negative, which requires that A, itself is less than unity. Similarly, if AG? is, negative, the reaction tends to favor products. Physical insight to this behavior can be obtained by appealing to the definition of AG in terms of the enthalpy and entropy changes associated with the reaction, From Eqn. 2.51, we can write AG; = AH" — Tas" which can be substituted into Eqn. 2.66b: Kae INT ASIN, For K, to be greater than unity, which favors products, the enthalpy change for the reaction, AH”, should be negative, ie., the reaction is exothermic and the system energy is lowered. Also, positive changes in entropy, which indicate greater molecular chaos, lead to values of K, > | Consider the dissociation of CO; as a function of temperature and pressure CO, & CO+403, Find the composition of the mixture, ie. the mole fractions of COs, CO, and O, that results from subjecting originally pure CO; to various temperatures ("= 1500, 2000, 2500, and 3000K) and pressures (0.1, 1, 10, and 100atm). Solution To find the three unknown mole fractions, xco,. xc0: Nd xo, We will need three ‘equations. The first equation will be an equilibrium expression, Eqn. 2.66. The other {to equations will come from element conservation expressions that state thatthe total amounts of C and O are constant, regardless of how they arc distributed among the three species, since the original mixture was pure COs, To implement Eqn. 2.66, we recognize that a= 1,b= 1, and c (CO, @ (1)CO + Os, ‘Thus, we can evaluate the standard-state Gibbs function change. For example, at T = 250K, AG; (Who, + DEf.co ~ (NEF con}ra2sn = (0 + (1-327,248) — (396,152) 8.907 kI/kmol. ‘The values above are taken from Appendix Tables A.1, A.2, and A.11 From the definition of K,, we have (Poof PN Pot PY? (Peo, 4a Exam a CHAPTER? «Combustion and Thermochemistry We can rewrite K, in terms of the mole fractions by reco, ee pe 7eor Substituting the above into Eqn. 2.66b, we have Heoxd(P/PY _ [AGF Heo, LRT spl 258.907 P[(8.315)2500), soot P/F 03635. o 0, We create a second equation to express consertation of elements: No. of carbon atoms _1___x00-+ x00, No. of oxygen atoms ~ 2 ~ Xco + 2xc0, + 2¥0, We can make the problem more general by defining the C/O ratio to be a parameter Z that can take on different values depending on the initial composition of the mixture: Xoo + x00, Xeo + 2xco, + Bro, (Z— Vxco+ QZ ~ Ixco, + Zx0, =O. an To obtain a third and final equation, we require that all of the mole fractions sum to unity Heo + Xco, + Xo, = amy Simultaneous solution of Eqns. I, I and IIT for selected values of P, T, and Z yield ‘values for the mole fractions xco, coy» and o,. Using Eqns. 1 and III to eliminate eo, And xo,, Eqn. I becomes Kcoll = 22 + Zyco)(P/PI" — BZ ~ I + Z)xcolexpl—AGF/R,T) = 0. ‘The above expression is easily solved for yoo by applying Newton-Raphson iteration, which can be implemented simply using spreadsheet software. The other unknowns, xco, land %o,. are then recovered using Equations 11 and I Results are shown in Table 2.2 for four levels each of temperature and pressure Figure 2.12 shows the CO mole fractions over the range of parameters investigated, Comments ‘Two general observations concerning these results can be made. First, at any fixed temperature, increasing the pressure suppresses the dissociation of CO; into CO and. (On; second, increasing the temperature at a fixed pressure promotes the dissociation, Both ofthese trends are consistent with the principle of Le Chitelir that states that any Chemical Equilibrium Table 2.2 Equilibrium compositions ot various temperotures ond pressures for CO, CO+ Jr Patan P=Wam P= 100atm T = 1500K, AGF = 1.5268 108 kmot xco 1785-10 3.601 -10-* 1672-104 776.10 Xen 0.9088, 9998 0.9997 0.9999 x0, 3877-10 Exot 10* 8357-105 358-10 T= 2000K, AG} = 1.10462. 108 J/hamol xo 0315 none 696.107 3283-10 Xe. 09877 osm 0950s 99st Xo wise oor 348-107) 1622+ 10-" T= 2500K, AG; = 6.8907 107 J)hmot xco 0260, 0.1210 0.0802 0289 Xeo, 0.6610 ogiss 0.9036 0.9566 x0, 0.1130 0.0605 0.0301 oss 100K, AG} = 2.7878 107 kao oo oot aia ci = i a con can a es “es rey wore, 10° eo _ 300K ww} So = i ~ Zi 20k : — 8 — ane 10) oe T = ce Figure 2.12 The CO mole fractions resulting from dissociation of pure CO, at various prestures ond femperoivr. CHAPTER? «© Combustion and Thermochemistry system initially in state of equilibrium when subjected to a change (e.., increasing pressure or temperature) will shift in composition in such @ way as to minimize the change. For an increase in pressure, this translates to the equilibrium shifting in the direction to produce fewer moles. For the CO; < CO + 40, reaction, this means a shift to the left, to the CO» side. For equimolar reactions, pressure has no effect. When the temperature is inereased, the composition shifts inthe endothermic direction. Since heat is absorbed when COs breaks down into CO and Os, inereasing the temperature produces shift to the right, to the CO +$0» sie ‘Complex Systems ‘The preceding sections focused on simple situations involving a single equilib- rium reaction; however, in most combustion systems, many species and several simultaneous equilibrium reactions are important. In principle, the previous example could be extended to include additional reactions. For exam- ple, the reaction Oy ¢ 20 is likely to be important at the temperatures con- sidered. Including this reaction introduces only one additional unknown, zo. We easily add an additional equation to account for the Op dissociation: (x0/x0,)P/P? = exp(—AGF/R,T), where AG# is the appropriate standard-state Gibbs function change for the O2 20 reaction. The element-conservation expression (Eqn. Il) is modified to account for the additional O-containing species, No. of C atoms co + Xco3 No, of 0 atoms ~ xeo + 2xc0, + 2x0, + Xo" and Eqn. III becomes Xco + Xco, + Xo, + Xo We now have a new set of four equations with four unknowns to solve. Since two of the four equations are nonlinear, it is likely that some method of simultaneously solving nonlinear equations would be applied. Appendix E presents the generalized Newton's method, which is easily applied to such systems. ‘An example of the above approach being applied to the C, H, N, O system is the computer code developed by Olikara and Borman [8]. This code solves for 12 species, invoking seven equilibrium reactions and four atom-c« tion relations, one each for C, O, H, and N. This code was developed specifi- cally for internal combustion engine simulations and is readily imbedded as a subroutine in simulation codes. This code is used in the software provided with this book, as explained in Appendix F. ‘One of the most frequently used general equilibrium codes is the powerful NASA Chemical Equilibrium Code [5}, designated CEC86. The code is fre- quently updated so the “86” represents the update year. This code is capable of handling over 400 different species, and many special problem features are

You might also like