You are on page 1of 14
Thermochemistry of the Rocket Motor In Chapter 5 we discussed methods to determine thrust and exhaust velocity om the chamber conditions, i.e. the chamber temperature, T,, the mean ‘molecular weight of the combustion products, Al, and the ratio of specific heats, y. These quantities were supposed to be known. However, in general, we only know the propellant combination used, and we have to determine the composition and temperature of the combustion products in the chamber at a prescribed chamber pressure. _. Moreover, we assumed the ratio of specific heats and the composition of the combustion products constant during the expansion through the nozzle. Jn real rocket motors, there will be deviations from this idealized situation “ due to dissociation and recombination, variation of specific heats with temper- ature, condensation, incomplete reactions, and ionization. Generally, the first two effects dominate. ., Because of dissociation and recombination, the molecular weight of the combustion products, as well as the ratio of specific heats, is not constant. - Moreover, dissociation and recombination are attended with large heat absorptions or releases. Even if no reactions take place in the nozzle, y will vary during expan- sion, as the specific heats depend on the temperature. “1 Because of condensation, which involves heat releases, the density of the combustion products increases. Other effects of two-phase flow have been discussed in Section 5.2. _- In general, the residence time of the species in the combustion chamber is -Tather short, usually less than a few milliseconds. The time required for the formation of a chemical substance may be much larger, and such a substance = May be formed only in small quantities or not at all. “=: Ionization becomes an important factor at temperatures in the order of 5000 K. As in chemical rockets the temperature does not exceed 4000 K, ionization is of minor importance and is not treated in this chapter. 1.1 Concepts of chemical thermodynamics 71.1.1. General definitions Molar quantities. In chemistry and physics, it is convenient to base the Xtensive quantities, such as enthalpy, entropy, etc., on a basic amount of Matter. The basic amount of matter to be used as unit is the mol (mole in 124 Rocket Propulsion & Spaceflight Dynamics older notations). One mol of matter contains the same number of particles. (molecules, atoms, etc.) as there are atoms in exactly 12 g of the isotope °C (In older definitions hydrogen and oxygen have been used as a referen é substance). This number is called Avogadro’s number, N,, Table T.1. The' molecular weight, Jt, of a substance is defined as the mass of one mol of th: substance. It follows from this definition that the molecular weight of 1 isotope *C is 12 g/mol. Extensive quantities, based on the mol (km instead of the gram (kg) are called molar quantities. Instead of expressing the amount of chemicals considered in mols, we may: as well express them in mols per unit volume. In that case we speak of: concentration. Compounds. A molecule of a compound consists of one or more atoms of one or more elements, ie. H as well as NH,CIO, are regarded as co: pounds, The reaction equation. Consider a system initially containing the substances A, We observe that other substances A, are formed and express this by the: reaction equation LA, >YnApy r=1,2,-+°,R and p=R+1,R+2,---,P, ° 4-1); 4 The substances A, are called the reactants, and A, the products. The 4 coefficients n, indicate the concentration or number of mols of Aj, where j=1, 2, +++, P. An example may clarify this. Consider the reaction 2H, + O,~> 2H,0. Here H, and O, are the reactants, H,O is the product, an 2, 1, and 2, respectively, are the concentrations or the number of mols of reactants and products. The equation expressing the formation of products? from the reactants is called the forward reaction. Another reaction is usual going on simultaneously: the back reaction nA, - Yn, (71-9 describing the conversion of products into reactants. The Eqs. (7.1-1) ant (7.1-2) can be combined to give Lind, =D tyAy Chemical equilibrium. If the concentrations of products and reactants re: main constant, the two reactions (forward and back) are in equilibrium wit each other. We define the state of chemical equilibrium as the state in whid there is no spontaneous tendency of the system to change its internal! composition, either by diffusion or by chemical reactions.’ Thermochemistry of the Rocket Motor 125 ‘Stoichiometric coefficients. If it is conceivable that all reactants in Eq. 1-3) disappear to form products, m, and n, are called stoichiometric ‘oefficients, which will be indicated as v, and »,. In the equation 2H, + O,22H,0, e stoichiometric coefficients are 2, 1 and 2. However, we could write as ell H,+30,2H,0. In this case, the stoichiometric coefficients are 1, } and 1. We thus note that ‘the important thing about the stoichiometric coefficients is their ratio, and not their magnitude. A mixture, such that the concentrations of the various species are proportional to the respective stoichiometric coefficients, is called a stoichiometric mixture. Consider a system initially containing n,, mols of A;. These substances, As eact with each other to form products, Aj, according to the stoichiometric (7.1-4) jHRtl---,P. The numbers of mols of the substances A; and A, actually present are indicated by n; and nj, respectively. If no chemical equilibrium is reached yet, the reaction YnAreY nAp ill proceed to the right according to Eq. (7.1-4) in such a way that (7.1-5) “and & is called the extent of reaction. in this case we assumed that the reaction was progressing to the right, thus dn, <0. If the reaction progresses to the left and the substances A; are isappearing, dnj<0. The Eq. (7.1-5) states that the change in concentra- ‘tion or number of mols of chemical substances during a reaction is propor- ional to their respective stoichiometric coefficients. Heat of reaction. The first law of thermodynamics (Appendix 3) for ‘Systems involving only compression and expansion work, and thus applicable to the rocket motor, states: dQ=dE+pdyY, (7.1-6) the heat input dQ into the system balances the increase dE of internal ergy and the work, p dY, performed on the surroundings. The enthalpy, i Rocket Motor 127 126 Rocket Propulsion & Spaceflight Dynamics ‘Thermochemistry of the Roc! 4H, is defined as ombustion chamber has remained the same, all the time. Therefore, the combustion in a rocket motor essentially is a constant pressure reaction. In ‘the following, we therefore will restrict ourselves to constant pressure yeactions. H=E+py, (4 and the first law can also be written as Lt Hess’ law. A direct consequence of the first law of thermodynamics, and H “and E being state variables, is that the heat change is the same whether the teaction takes place in one or in several steps. This is known as Hess’ law, or the law of constant heat summation. Now, consider a chemical reaction taking place at constant pressure and aj a specified temperature. The index 0 indicates the state before the reaction the index e the state after the reaction. The heat put into the system durin; the reaction is denoted by Q. According to Eq. (7.1-8), we find = H.-H, = AH. ur equation. The heat of reaction depends on the temperature at hhoff” Th yf di yn the temper Q=H,-H,=AH. Q Kirchhoff's eq which the reaction takes place. For a constant pressure reaction, we have AH(T) = H.(T)- H,(T), (7.1-14) where the enthalpies are evaluated at the temperature T. Differentiating both sides with respect to temperature leads to wal) _ (2H) _ (ee) Cr p \@T/> \aT]p . and as (aH/dT), = C,, we have Here AH is called the heat change at constant pressure, or commonly the heat of reaction. The reaction is called endothermal if Q>0, and exothermal if Q<0. A combustion process, in general, generates heat and thus is exothermal. If one considers reactions taking place at constant volume and a specifies temperature, the heat Q put into the system during the reaction, accordin, to Eq. (7.1-6) is Qu E,~E,=AE, (7.1-10 where AE is called the heat change at constant volume. There is a direct relationship between the changes in enthalpy and internal energy. Consider a reaction at constant pressure. The volumes o! the system before and after the reaction are indicated by ¥, and Va: respectively. It then immediately follows from Eq. (7.1-7) that AH=AE+p(v.-Y,), (71-11) 3a. G.= (71-15) OOF) £0. -C,=AC, (r 7 Cm Cre = BC This result is known as Kirchhoff’s equation. Integration leads to T, an(r)—aH(T,)= [ AC, aT. (7.1-16a) ry Ih those cases where it is admissible to assume AC, constant, this equation simplifies to AH(T>)— AH(T;) = AC, (Tz T;)- (7.1-16b) Tn general, the heats of reaction at different temperatures are related by AH(T;) = AH(T;) +[H(T2)— H(Ti)e-LA(Ta)- A(T]. (7.1-16¢) - For constant volume reactions, identical results hold if one replaces H by E and C, by C,, as the reader can verify himself. For reactions involving only liquids or solids, the term P(V.-V), in’ general, is negligible small as compared to AE. For reactions involving gases too, this is not the case. The ideal gas law states pY=nRoT, (71-12) where n is the number of gaseous mols in the total volume Y. A reaction, taking place at constant pressure and constant temperature, involving ideal. gases thus yields P(V.- Vo) =(n,~n,)RyT = AnR,T, or, with Eq. (7.1-11), AH=AE+AnR,T. (71-13) landard heat of formation. The heats of reaction can be calculated theoreti- ly, or be determined by experiments. It would be convenient, of course, to have these heats of reaction in tabular form. It is, however, not possible to list all possible heats of reactions for a certain substance, as it may be formed in infinitely many different ways. Besides, the heats of reaction depend on the state of aggregation of the products and reactants, the ‘mperature and the number of mols involved. As a result of Hess’ law, however, it suffices to list the heat of reaction for the formation of one mol The amount of propellants that enter the combustion chamber of a rocket motor during the time interval At is mAt. After combustion has taken place,’ the volume of this amount of matter has increased considerably, as, 12, general, liquids or solids are converted into gases. The pressure in the. 128 Rocket Propulsion & Spaceflight Dynamics of a substance from its elements, at a certain reference temperature, whil we take the elements and the products in a standard reference state. This heat of reaction is called the standard heat of formation, and is denoted by AH?. The standard reference state for a solid or a liquid is its most stable state at a certain specified temperature and a pressure of 0.101325 MPa(1 atm). F gases, it is convenient and customary to take the standard reference state the (hypothetical) ideal gas condition at a certain specified temperature and: at a pressure of 0.101325 MPa (1 atm). To indicate that a thermodynami quantity corresponds to the standard reference state, the superscript ° used; so we write H°, E°, etc. One finds the heats of formation listed in! thermochemical tables [1, 2, 3]. From its definition, it follows that AHF= for all elements. The reference temperature at which AH; is listed, is mostl 0K or 298.15 K. It should be mentioned that, while the above definitio are the most common ones, some tables use other definitions. Just like potential functions, H and E are fully determined, except for a constant. Often, this constant is taken zero for the elements, either at 0 K o1 at 298.15 K. For a particular substance, the values of the absolute enthalpy, H°, may differ from table to table, depending on the reference temperature and the assigned value of the numerical constant. In general, the tables also ; list the sensible enthalpy, H°(T)—H°(0). These values will be the same for S all tables. From the definition of the standard heat of formation and Eq. (7.1-14) immediately follows that AH(1)= Y [nAH{T)],-Y, [nAHA(T)],, (14-17) where n indicates the number of mols considered and the indices e and 0 indicate the final and initial state respectively. i If the heats of formation are given at a reference temperature T;, while the heat of reaction is evaluated at a temperature T,, and the reactant: initially had the temperature T;, one finds AH(T,) = Y'n,[AH}(T;) + (H(T;)—H°(0))—(H°(T,)— H°0))]. ~ Yn [AH}(T;) + (H°(T5)— H°(0))— (H°(T;)— H°(0)) Joy (7.1-18a) where the equation is written in sensible enthalpies. In the case that one.” uses absolute enthalpies, Eq. (7.1-18a) can be written as § AH(T;) = H.(To)— He(Ta). (7.1-180) If all heat is used to heat up the products, i.e. AH°(T>) = 0, then Tp is called: the adiabatic flame temperature or combustion temperature. A few examples may illustrate how the above derived equations can be used to compute the heat of reaction, or the adiabatic flame temperature. First consider the formation of water from its elements at 500K. We wan! “From Table 7.1 we see (AH;(298.15))s,0 H,0(g25) Hi(gas) 0,(¢23) ‘Temperature H°(T)-H(0) HT) H(T)-H(0) HT) H(T)-H(0) A(T) (K) (kJ/mol) (kJ/mol) (kJ/mol) (kJ/mol) (kI/mol) (kJ/mol) 298.15 9.9065, 57.3158 8.4676 290.5399 8.6805 17.2000 ‘500 16.8427 64.2520 14,3490 296.4236 14.7645 23,2840 118.6197 166.0291 85.1866 368.2589 94,8766 103.3963 2800 123.9543 171.3637 89.8342 371.9066 98.8223 107.3418 2900 129.3178 176.7271 93.5061 375.5784 102.7883 111.3078 (AH#(298.15)) 150 = ~ 2418264 kJ; H°(298.15):4,0,1iquia = 13.3001 kJ/mol io find the heat of reaction and will use the following reaction equation: H,+40,->H,0. —241.8264 kJ. So, using sensi- ble enthalpies, Eq. (7.1-18a) with T;= T,= 500 K yields 4H°(500) =[—241.8264 + 16.8427 — 9.9065]—[14.3490 — 8.4676] —3[14.7645 — 8.6805] = —243.8136 kJ. Using absolute enthalpies, Eq. (7.1-18b) with T;=T,=500K yields AH?(500) = 64.2520 — 296.4236 —3 x 23.2840 = —243.8136 kJ. ‘If a phase transition takes place, i.e. a transition of liquid to gas, etc., this is accompanied by a change in enthalpy: the heat of transition, also denoted by AH}. The heat of transition at a temperature T can be determined with Eq. (7.1-18a) or Eq. (7.1-18b) where one phase of the substance is regarded as reactant and the other one as product. For example, for the transition of ater vapor to water, at 298.15 K, AH = H$,,0 ga H9,0 jiquia = 57-3158— 13.3001 = 44.0157 kJ. » Let us now compute the flame temperature of a mixture of gaseous hydrogen and gaseous oxygen, at an initial temperature of 298.15 K, which -Teacts according to ©) SH, +O,>2H,0 +3Eh. Using sensible enthalpies, we have according to Table 7.1, 2x[-241.8264-+ (H°(T;)— H°(0)),0~ 9.9065] + +3 X[(H°(T;)— H°(0))y,~ 8.4676] = 0- ‘here is no other contribution from H, and O, because T,=T;3 and the 130 Rocket Propulsion & Spaceflight Dynamics. heat of formation for the elements equals zero. So, we are left with 2.x [H°(Ts) —H°(0)lago+ 3 * LH°(To) ~ H°(O)]ex, = 528.8686 ki. By interpolation in Table 7.1 we find for the combustion temperature, T,= 2853 K. Using absolute enthalpies we have 2X (H°(T>))ux,0+ 3 X (H°( To), = 5 X 290.5399 + 17.200 = 1469.6995 ki, and by interpolation we find T, = 2852 K. At these elevated temperatures, however, another phenomenon takes place: dissociation. Compounds will partially break up into electrically: neutral fractions called radicals. Water, for instance, may break up accord- ing to the reaction H,O > n,H,0+ n,H,+n30,+ngH+ns0+n.OH. The last three species in this equation are the radicals mentioned. As disso tion is a highly endothermal process, the flame temperature may drop much below the value as computed by the above method. In rocket motors, the combustion temperatures are so high that the effects of dissociation canno! be ignored and we will deal with this in the next section. The opposite of dissociation is recombination. This phenomenon is important, for determi: ing the performance of a rocket motor, as it may take place in the nozzle. Thermodynamic properties of radicals are, like those for compounds, listed in thermochemical tables. 7.1.2 Chemical equilibrium The state of chemical equilibrium is defined in Section 7.1.1. In this section we will discuss thermodynamic relationships for chemical equilibrium reac- tions. The second law of thermodynamics states for a change between state 1 and state 2: 2 d Ys 5.8 [ @ As for an isolated system dQ=0, the entropy, S, will increase until it has reached its maximum value for that system. In the state of equilibrium, where no irreversible processes take place S= Sirax- (71-19) (7.1-20) Combining the first and second law of thermodynamics for a system al constant temperature yields 2 s-5=5 (aH-¥ ap). (74-21) Thermochemistry of the Rocket Motor 131 ‘If the pressure is also kept constant, we find H,~TS) nH, + njH+ nsOH + ngH,0 +30 + 1602. 138 Rocket Propulsion & Spaceflight Dynamics Thermochemistry of the Rocket Motor 139 The equations for the conservation of elements are th the chamber pressure, p,, determines the exhaust velocity, V., and the xhaust pressure, p,. If one, on the other hand, wants to design a nozzle for iven chamber pressure, pressure ratio and mass flow, the area of the exit face and the exhaust velocity follow from the chamber pressure, pressure ratio, and mass flow. In both cases, the ambient pressure, p,, is an indepen- lent variable that will not interfere with the nozzle flow as long as no shock ‘Waves occur. It is convenient to take the pressure, and not the area of a 2m=2n,t+ n+ n3t+2n,, 2=ngt nyt ngt2ng. We assume the chamber pressure being known. Then it is convenient write the equilibrium constants of formation in terms of the chambe pressure, p., and the number of mols n,,°+-, "6, N, where surface of constant properties, as an independent variable, as the pressure 6 “appears in the equations for the equilibrium constants and the entropy N= zd Ni equation. Therefore, we will determine the local velocity and the local area ie fa surface of constant properties as a function of the local pressure. In general, the composition of the combustion products will change during the expansion through the nozzle. For an exact determination of how the composition changes, the integration of the differential equations for the hemical kinetics and hydrodynamics is necessary. This is beyond the scope of this book. We can, however, distinguish between two extreme cases: may be regarded as an additional unknown. The equilibrium constants of formation for this system are 2 +010; Ky = (Z), Bn \P, ng (N H+O 08; Kina (F). 1. There is no effective change in composition during the expansion ans \P, “through the nozzle. 2H SH; Ky, a4 ( (N’ ). 2 The composition at any location in the nozzle is according to chemical ng \P. -€quilibrium. ng (N AES 205 02; Ky.= (=) In the first extreme case, one assumes the residence time in the nucie to be 5\P. 4 ery short, as compared to the time needed to change the composition “effectively. Moreover, if we assume that the heat capacity of the mixture of combustion products is constant too, we can use the theory as developed in Chapter 5. We call this the constant properties flow. If we do assume a ‘onstant composition throughout the nozzle, but take into account the "|. Variation of the heat capacities with temperature, we speak of a frozen flow. + In the second extreme case, one assumes the reaction rates to be infinite, so,that the composition changes immediately to the one dictated by equilib- lum conditions. In this case we speak of equilibrium flow. where P,=p,/p, the ratio of chamber pressure to reference pressure Finally, we have the energy equation mH Ta, + (To, = mH Ta, + M2H(T a + 3H(Tout +4 H(T 0+ NsH° (Tot neH(T)o,. Here, T, is the temperature at which the respective propellants, H and Oy enter the combustion chamber, and T; is the adiabatic flame temperature, which is to be determined. So, now we have the complete set of equation! for the composition and the adiabatic flame temperature. These equation! . have to be solved numerically. : 73.1 Constant properties flow is the most simple approach. The composition and heat capacities, hich remain constant during the expansion through the nozzle, are known from the chamber conditions. The mean molecular weight, Ml, of the ombustion products, follows from H=¥, (Al) |S (73-2) here n/n, and A, stand for the mol fraction and the molecular weight, k 7.3 Expansion through the nozzle The thrust of a rocket motor is, according to Eq. (6.2-3), given by F=AmV,+(De-Pa)Ac- (73-1 The mass flow, m, equals the rate of propellant flowing into the combustiot: chamber, while A depends on the nozzle shape. For a given rocket motor, the area of the exit surface, A,, in combinati 140 Rocket Propulsion & Spaceflight Dynamics Thermochemistry of the Rocket Motor 141 respectively, of the combustion products A,. The mean molar heat capaci ‘Here, the index g denotes the gaseous species and the index s the condensed at constant pressure, C,, follows from pecies. For a given pressure, the temperature is determined from this last quation. G= y (mC,,) / x Mes (7.3 “In Section 3.7 we derived the energy equation; for stationary flows and the gas constant, R, of the mixture from V-V(h+3V-V)=0. (7.3-9) R=RJ/M. (73 4 ing the same arguments as in Section 5.1.1 we have The ratio of specific heats, y, is determined by A(T) +2V-V = h(T.), (73-10) G the sum of the enthalpy and kinetic energy per unit mass of the 7 (7.3- combustion products is constant and equals the enthalpy per unit mass in the Ro We assumed no heat losses in the combustion chamber, therefore, the’ adiabatic flame temperature equals the temperature of the gases in the’ combustion chamber, i.e. T;= T,. Now that p,, T,, R and y are known, the equations in the Sections 5.1.1 and 5.1.3 allow the direct computation of the velocity, V, the local expan- sion ratio, A/A, and the ratio m/A for a given local pressure, p. If this pressure is taken as the exit pressure, p,, one finds the exhaust velocity, V,' with Eq. (5.1-15), the expansion ratio, A,/A, by Eq. (5.1-42) and the ratio m/A, with Eq. (5.1-37). combustion chamber. In this chapter, we use the molar enthalpy and Eq. (7.3-10) becomes LAAT) = LmBiCL) = Ye) +3) dl, V?. (73-11) 7 c Here, T, is the temperature at which the propellant, A,, enters the combus- nm chamber. T, is the temperature of the combustion products in the combustion chamber. If there are no heat losses T, = T;, the adiabatic flame lemperature. T and V stand for the temperature and velocity, respectively, of the combustion products, A, in the nozzle, while n, stands for the number of mols considered, of the combustion products. The (numerical) ‘olution of Eq. (7.3-11) yields the velocity, V, as for a given pressure, p, Eq. (7.3-8) yields the temperature, T. If p= p,, one finds the exit temperature, T,. The exhaust velocity, V., follows from Eq. (7.3-11): 7.3.2 Frozen flow The main difference of this model with the one above, is that the heat capacities of the various substances are a function of the local temperaturt Therefore, the simple equations, as derived in the Sections 5.1.1 and 5.13 cannot be applied anymore. The expansion process through the nozzle is assumed adiabatic and reversible, i.e. the entropy remains constant. This means that all heat transfer and frictional effects are neglected. For the amount of matter considered, the entropy at any position in the nozzle thus: equals the entropy in the combustion chamber, or at the nozzle entrance. In Eq. (7.1-27) we showed that F,.~Fl=R,T In P,. (73-6) With the definition of free energy, Eq. (7.1-22), and as H(T)=H°(T), we find an a (Sn,81912)~SmsFR(T.) 7 Trl V. (7.3-12) ‘The area of the surface of constant properties per unit mass flow, A/m, is determined from the continuity equation, Eq. (5.1-35). For the nozzle we have - m= pVA, (7.3-13) where m is known. Because the combustion products in the nozzle may S%(T)~ $,(T) = Ry n P,. 73-1). : (T)~ S(T) = Ro In ( tain condensed particles, we write As condensed phases do not contribute to the total pressure, the entropy is found from Ying, +Erglly ri : 5. s(1)=Yn(sx7-R, In (Ps#)) +E n.5x0, (73-8) a a (7.3-14) & s “EAs the condensed particles do not contribute to the pressure, the volume, 7, bln 142 Rocket Propulsion & Spaceflight Dynamics considered, follows from RoTY t=— (7.3-15) From the Eqs. (7.3-13) to (7.3-15) follows RoTiing AL * (73-16) (Selle +Enctt) VP Now that temperature and velocity at a position in the nozzle are known, the area of the surface of constant properties follows from Eq. (7.3-16). If no condensed phases are present, Eq. (7.3-16) reduces to AL RT m MVp’ (7.3-17) where is the mean molecular weight of the combustion products as given by Eq. (7.3-2). The calculations for the frozen flow require the numerical solution of equations. The constant properties flow can be determined analytically. In cases that a frozen flow performance calculation is available, while a neighboring solution is required, the following simple computational scheme can be applied. First, the average heat capacity ratio, 7, for which both the frozen flow and the constant properties flow yield the same results for a given pressure ratio, p/p, is determined with the help of the Poisson relations, Eq. (3.7-20), from which follows ——In(pdp)_ in (p/p) ~In (TITY where T/T is the temperature ratio, as determined by the frozen fiow calculations. As ¥ is only slowly varying with temperature, especially at lower temperatures, a constant properties analysis, using 7 as determined by Eq. (7.3-18) will yield accurate results for nearby pressure ratios, as com- pared to the frozen flow analysis. There is no direct way to determine the expansion ratio A,/A,. One can, for instance, approximate the throat conditions by those that follow from the constant properties flow. These can be used as initial values to determine the throat as the location where M, = 1 by an iterative procedure [9]. In general, one will compute the ratio A/m for various pressures, lying between chamber and exit pressures. These data can be used to determine 4 minimum of A/m. This method also immediately yields p, [13]. (7.3-18) Thermochemistry of the Rocket Motor 143 The throat conditions are determined by A/m being a minimum, Le. a ( ien)= 0 (7.3-19) dp \MVp For the frozen flow, / is constant, and the requirement comes down to aT (1 = (73-20) ap\" Vat) p where use is made of V= V{T(p)}. As the flow is assumed isentropic, we also have, TdS=C,dT-Y¥ dp=0, (73-21) which gives mRoP (2), -- yet a dp/s Yn.C,, LmCyP y From Eq, (7.3-11) it follows that G av, ym (73-23) aT | VI nM, x | Combination of the Eqs. (7.3-20) to (7.3-23) yields Reps rye. (7.3-24) et Tm Val < The simultaneous solution of the Eqs. (7.3-8), (7.3-12) and (7.3-24) yields P, T, and V,. The ratio Am is found from Ar_ Rol (7.3-25) m MVD, The throat conditions, being determined by one method or another, yield, in combination with the values of A/m, the local expansion ratio A_Alm (7.3-26a) A, Adm - The expansion ratio, A./A, is found by setting A= A, Ae_ Adm (7.3-26b) A, Ajm eee 144 Rocket Propulsion & Spaceflight dynamics Thermochemistry of the Rocket Motor 145 f the location in the nozzle is taken as an independent variable. By Eq. 1-42) one then finds a relation betweeen dn,/dT and AH;.,. The solution f the equations in the unknowns n,, dn,/dp, T and V yields the throat nditions [14]. We have discussed three methods that yield results, which are more or less in agreement with experimental results. The equilibrium flow method, iccounting for many variations, requires much computer time. As it is based in the assumption that the composition varies according to chemical equilib- jum, there is no guarantee that its results are a better approximation of eality than the frozen flow. The frozen flow requires less computing time, but assumes a constant composition throughout the nozzle. In fact, the real mposition will lie somewhere between the results as computed by the ‘frozen flow and equilibrium flow method. There is a fourth method, how- “ever, that combines both methods mentioned previously, which we will “discuss below. An advantage of this last method is that the throat conditions can be determined beforehand, while some of the other methods require compu tion of A/m before the throat conditions can be determined. 7.3.3 Equilibrium flow In this case we assume infinitely fast reaction rates, so that the composition of the combustion products is a function of chemical equilibrium at any location in the nozzle. Thus, not only the heat capacities, but also the composition, are dependent on the location in the nozzle. Again, we assume an isentropic flow through the nozzle. To determine the velocity, temperature and composition for a given: pressure ratio, the following equations have to be solved: 1. The equations for the conservation of elements. If e different elements are involved, there are e equations. 2. The equations for the equilibrium constants of formation. If ¢ com- pounds are involved, there are c—e equations. 3. The entropy equation 7.3.4 The modified Bray approximation 2. Tn a classical article, Bray [15] studied the effects of recombination in the nozzles of hypersonic windtunnels. He showed that, by good approximation, the flow in the nozzle can be divided into three regions: equilibrium flow, transition region, and frozen flow. The transition region, in general, is very small and may be approximated by an infinitely small region, the Bray freezing point. The various reactions taking place in the nozzle will yield freezing points at different locations. In general, however, there is one dominating reaction with respect to energy release, and this is the one that ‘determines the Bray freezing point for the rocket motor [13]. Reaction rates increase with temperature, while the recombination rates increase with pressure. It is these recombinations that yield the high S(T.) = Ying) Ry ( Fa) +Snsn, 3.20 which equates the total entropy in the chamber to the total entropy at some. location in the nozzle for the amount of matter considered. 4. The energy equation from which the velocity, V, follows: we Ld, (73-28) nergy release. For moderate chamber pressures (~5.5 MPa) it is possible to oid the exact determination of the Bray freezing point, which requires knowledge of the reaction rates, but to assume the freezing point location The exhaust velocity, exit temperature, T,, and the composition at the at the throat. The justification is that temperatures and pressures before the ‘throat remain relatively high, while they rapidly drop after the throat. The Teverse is the case for the fluid velocity. Experience shows this approxima- tion to be very good [13]. -For this modified Bray approximation an equilibrium flow calculation is Performed from the chamber to the throat. It is followed by a frozen flow alculation from the throat to the exit. The throat conditions are determined according to equilibrium flow calculation. The Figs. 7.1 to 7.3 show the results of performance computations for a H,-O, rocket motor (based on LH, and LOX). ‘As recombination is accompanied by energy release, the calculations for he equilibrium flow yield higher specific impulses than for the frozen flow. e dependence of the maximum specific impulse on the mixture ratio and properties is found by Eq. (7.3-17). The average heat capacity ratio, 7 found by Eq. (7.3-18), with T,/T now as determined by the equilibrium flow calculations. . To determine the throat conditions, we can use the same iterative proce- dures as were-indicated to compute the throat conditions for the frozen flow: We can also use the simple condition of A/m being at a minimum. As the, | composition now varies according to chemical equilibrium, one also has to. determine the differential quotients dn,/dp from algebraic equations. Thes¢ equations are found by differentiating the e equations for the conservation of elements and the c—e equations for the K,,’s with the respect to pressur One should be aware that K,y(T)= Key{T(p)} if the local pressure, p, instead 146 Rocket Propulsion & Spaceflight Dynamics 450, 4S Vaeal Expansion, 9, 0101325 MPa Line of Maximum "J Equilibrium Flow a Mature Ratio jpg, 20.25 gino - a | +43 8100 3350 z g 3 WR \__| z 300 4 E 2 210} a & | io Equilibrium Flow 350 250 |-35| Ideot Expansion ‘Ay=0101325 MF 10 0 005 00 O15 020 025 030 O35 Mixture Ratio ™y2/™92 200 poopesire i 0 005 O10 Of O20 O25 030 O35 Mixture Ratio 49/779 = Bo BMG 5 10 1s 20g B Chamber Pressure A,(MPal 2 = 7h [a= 3 3 = 86 t poi) We & bons! & 110 2 0 a 2 35. a g 10 3 a € al 248 Frozen Flow. z g Ideal Expansion & 5 5 2 2 & 2 8 0 20 Chamber Pressure, (MPa) Fig. 7.1 The specific impulse of an ideal H,-O, rocket motor versus chamber =}. pressure and mixture ratio. The increments in specific impulse are plotted with respect to the equilibrium flow results chamber pressure is shown for the equilibrium flow. With increasing chamber pressures, the optimum mixture ratio shifts towards the stoichiometric mixture ratio, my,/mo, = 0.125. As the degree of dissociation increases with temperature, one finds the largest differences between the frozen and equilibrium flow calculations around the stoichiometric mixture ratio. For a mixture ratio 0.25, the differences between the Bray cycle and the frozen flow are small, while the J,, computed according to the method of constant gas properties does not differ much from those found by equilib- ” tium or frozen flow calculations. The expansion ratio determined by the method of constant gas properties yields larger differences. Finally, one can, by Fig. 7.3, determine the effect of a non-ideal expansion ratio on the exit pressure. For example, with (p,/p.)iaeai = 110, my,/mo,= 0.175 and an offset from the ideal expansion ratio, A(A,/A,) = 10™*, we see from Fig. 7.3 that A(p,/p.)/A(A¢/A,) = 10~ and hence, for p, = 11 MPa, we then find p, =—-1.1x10-> MPa. . We have determined the rocket performance parameters either by assum- ing constant gas properties, frozen flow or equilibrium flow. It is possible, Ideal Expansion, ,20101325 MPa ae T Eeiiiorbtn Flow . Equilibrium Flow < r ideal Expansion | Mixture Ratio my, m9,=0.25 Ss [2/2 Ng -0s01825 MPa ¢ [oz Pn Ra Ss > tree | x 2 6 Maximum (spa, 2 é -— @ 8 2 |un Mm | § 10 g 2 Ge & ee 5 Thermochemistry of the Rocket Motor 147 20 O05 010 O'S 020 O25 O30 O35 Mixture Ratio my, /M™Mo, < 64 20 5 70 1s 20 3 $ Ghamber Pressure Po) S 218 2 S\ayro= é eo | 3-0/4 = si é 70 G2 Se § aN a 3 wo | B4 Lavo 08 ot Frozen Flow! = £ | 4H ieeat Exponsion & 9 & Ns fea bo Lt geo 5 0 15 2 05 O10 GS 029 O25 030 035 Chamber Pressure p, (MPa) Mixture Ratio ™419/7M0, Fig. 7.2 The expansion ratio for an ideal H,-O2 rocket motor versus chamber Pressure and mixture ratio. The increments in expansion ratio are plotted with ~-fespect to the equilibrium flow results & A (PefPeif (Aer) 10->|_t10} zo | 4 0 005 01 O15 020 025 030 035 Mixture. Ratio tue Ratio Myo, Fig. 7.3 The increment in the pressure ratio due to an increment in the xpansion ratio for an ideal H,-O, rocket motor, versus mixture ratio 148 Rocket Propulsion & Spaceflight Dynamics however, to combine the equations of chemical kinetics and hydrodynamic: [16] to obtain a better insight in the processes occurring in the nozzle. A different approach to obtain more accurate results for the, nozzle flow i: followed by Prud’homme [17], who uses a quasi-equilibrium method to obtain a more realistic picture of the composition. These techniques, how. ever, are beyond the scope of this book. There is also the non-equilibrium effect due to the presence of condensed particles, and non-equilibrium chamber effects [13], which is not treated here either. It may be of impor- tance, however. An excellent discussion on the calculation of the chemical equilibrium, especially with respect to rocket motors, is given by van , Zeggeren and Storey [18]. These authors also pay much attention to various — computational procedures and discuss the respective pros and cons. References 1 Weast, R. C. (ed.) (1978), Handbook of Chemistry and Physics, 59th ed., C.R.C.-Press, Cleveland, 2 McBride, B. J., Heimel, S., Ehlers, J. G. and Gordon, S. (1963), Thermodynamic Properties to 6000K for 210 Substances Involving the First 18 Elements, NASA SP-3001, Washington. 3 Stull, D. R. and Prophet, H. (proj. dir.) (1971), JANAF Thermo- chemical Tables, 2nd edn, National Bureau of Standards NSRDS- NBS37, Washington. 4 Andrews, F. C. (1971), Thermodynamics: Principles and Applications, « Wiley-Interscience, New York, p. 243. 5 Penner, S. S. (1957), Chemistry Problems in Jet Propulsion, Pergamon Press, London, p. 193. . 6 Lewis, G. N. and Randall, M. (1923), Thermodynamics and the Free Energy of Chemical Substances, McGraw-Hill, New York, p. 190-202. 7 Rossini, F. D. (ed.) (1955), Thermodynamics and physics of matter, Vol. 1, in: High Speed Aerodynamics and Jet Propulsion, Princeton University Press, Princeton, p. 85-86. 8 Kestin, J. (1968), A Course in Thermodynamics, Vol. Il, Blaisdell, Waltham, p. 257. 9 Gordon, S. and McBride, B. J. (1971), Computer Program for Calculation of Complex Chemical Equilibrium Compositions, Rocket Performance, Incident and Reflected Shocks and Chapman-Jouguet Detonations, NASA SP-273, Washington. 10 Svehla, R.A. and McBride, B. J. (1973), Fortran IV Computer Program For Calculation of Thermodynamic and Transport Properties of Complex Chemical Systems, NASA TN D-7056, Washington. 11 Huff, V. N., Gordon, S. and Morrel, V. E. (1951), General Method and - Thermodynamic Tables for Computation of Equilibrium Composition and Temperature of Chemical Reactions, NACA Report 1037, Washington. © AS v4 15 “16 “7 18 Thermochemistry of the Rocket Motor 149 Barrére, M., Jaumotte, A., Fraeijs de Veubeke, B. and Vandenkerck- hove, J. (1960), Rocket Propulsion, Elsevier, Amsterdam, p. 140-153. Glassman, I. and Sawyer, R. F. (1970), The Performance of Chemical Propellants, AGARDograph 121, Technivision Services, Slough. Schéyer, H. F. R. (1976), The Determination of the Throat Conditions for Equilibrium and Frozen Flows in Rocket Motor Nozzles, Rpt. VTH- 209, Delft University of Technology, Dpt. of Aerospace Engineering, Delft. Bray, K. N. C. (1959), Atomic recombination in a hypersonic wind- tunnel nozzle, J. Fluid Mech., 6, 1-32. Bitther, D. A. and Scullin, V. J. (1972), General Chemical Kinetics Computer Program for Static and Flow Reactions, with Application to Combustion and Shock-Tube Kinetics, NASA TN D-6586, Washington. Prud’homme, R. (1974), Méthode de Quasi-Equilibre pour I’ Etude des Ecoulements Relaxés, paper presented at the 43rd PE.P.-AGARD specialist meeting, Liége. Van Zeggeren, F. and Storey, S. H. (1970), The Computation of Chemi- cal Equilibria, Cambridge University Press, Cambridge.

You might also like