Vous êtes sur la page 1sur 49

Chapter 9

Thermodynamics of reacting
mixtures

9.1 Introduction

For an open system containing several reacting chemical species that can exchange mass
and work with its surroundings the fundamental Gibbs equation relating equilibrium states
is

I
X K
X
T dS = dE + P dV µi dni + Fk dlk . (9.1)
i=1 k=1

The Fk are forces that can act on the system through di↵erential displacements. Ordinarily,
lower case letters will be used to denote intensive (per unit mole) quantities (h, s, e, etc)
and upper case will designate extensive quantities (H, S, E, etc). Heat capacities, pressure
and temperature are symbolized in capital letters. One mole is an Avagadro’s number of
molecules, 6.0221415 ⇥ 1023 .
The chemical potential energy per unit mole µi is the amount by which the extensive energy
of the system is changed when a di↵erential number of moles dni of species i is added or
removed from the system. If the system is closed so no mass can enter or leave and if it is
isolated from external forces, the Gibbs equation becomes

I
X
T dS = dE + P dV µi dni (9.2)
i=1

9-1
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-2

where the di↵erential changes in the number of moles of species i occur through chemical
reactions that may take place within the closed volume. The main di↵erence between
(9.1) and (9.2) is that, in the closed system, changes in mole numbers are subject to the
constraint that the number of atoms of each element in the system is strictly constant.
The precise expression of the chemical potential in terms of conventional thermodynamic
variables of state will be established shortly. For the present it can be regarded as a new,
intensive state variable for the species i. Mathematically, equation (9.2) implies that
✓ ◆
@S
µi (E, V, n1 , . . . , nI ) = T . (9.3)
@ni E,V,nj6=i

If no reactions occur then (9.2) reduces to the familiar form

T dS = dE + P dV. (9.4)

According to the second law of thermodynamics, for any process of a closed, isolated
system

T dS dE + P dV. (9.5)

Spatial gradients in any variable of the system can lead to an increase in the entropy.
Smoothing out of velocity gradients (kinetic energy dissipation) and temperature gradients
(temperature dissipation) constitute the two most important physical mechanisms that
contribute to the increase in entropy experienced by a non-reacting system during a non-
equilibrium process. If the system contains a set of chemical species that can mix, then
changes in entropy can also occur through the smoothing out of concentration gradients
for the various species. If the species can react, then entropy changes will occur through
changes in the chemical binding energy of the various species undergoing reactions.
The inequality (9.5) can be used to establish the direction of a thermodynamic system as
it evolves toward a state of equilibrium.

9.2 Ideal mixtures

Consider a mixture of species with mole numbers (n1 , n2 , . . . , nI ). The extensive internal
energy of the system is E and the volume is V . The extensive entropy of the system is the
function, S (E, V, n1 , n2 , . . . , nI ). An ideal mixture is one where all molecules experience the
same intermolecular forces. In an ideal mixture surface e↵ects, (surface energy and surface
tension) can be neglected and the enthalpy change when the constituents are mixed is
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-3

zero. Ideal mixtures obey Raoult’s law that states that the vapor pressure of a component
of an ideal mixture is equal to the vapor pressure of the pure component times the mole
fraction of that component in the mixture. In the ideal approximation the volume of
the system is the sum of the volumes occupied by the pure species alone. Similarly the
internal energy is the sum of internal energies of the pure species. Most real mixtures
approximate ideal behavior to one degree or another. A mixture of ideal gases is perhaps
the best example of an ideal mixture. Liquid mixtures where the component molecules
are chemically similar, such as a mixture of benzene and toluene, behave nearly ideally.
Mixtures of strongly di↵erent molecules such as water and alcohol deviate considerably
from ideal behavior.
Let the mole numbers of the mixture be scaled by a common factor ↵.

n1 = ↵ñ1 , n2 = ↵ñ2 , n3 = ↵ñ3 , . . . , nI = ↵ñI (9.6)

According to the ideal assumption, the extensive properties of the system will scale by the
same factor.

E = ↵Ẽ, V = ↵Ṽ (9.7)

Similarly the extensive entropy of the system scales as


⇣ ⌘
S (E, V, n1 , . . . , nI ) = ↵S̃ Ẽ, Ṽ , ñ1 , . . . , ñI . (9.8)

Functions that follow this scaling are said to be homogeneous functions of order one.
Di↵erentiate (9.8) with respect to ↵.

@S @S X @S
I ⇣ ⌘
Ẽ + Ṽ + ñi = S̃ Ẽ, Ṽ , ñ1 , . . . , ñI (9.9)
@E @V @ni
i=1

Multiply (9.9) by ↵ and substitute (9.8).

I
@S @S X @S
E +V + ni = S (E, V, n1 , . . . , nI ) (9.10)
@E @V @ni
i=1

The Gibbs equation is


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-4

I
X
dE P µi
dS = + dV dni . (9.11)
T T T
i=1

According to (9.11) the partial derivatives of the entropy are

@S 1
=
@E T
@S P
= (9.12)
@V T
@S µi
= .
@ni T

Inserting (9.12) into (9.10) leads to a remarkable result for an ideal mixture.

I
X
E + PV TS = ni µ i (9.13)
i=1

Equation (9.13) is called the Duhem-Gibbs relation. The combination of state variables
that appears in (9.13) is called the Gibbs free energy.

G = E + PV TS = H TS (9.14)

Equation (9.13) expresses the extensive Gibbs free energy of an ideal mixture in terms of
the mole numbers and chemical potentials.

I
X
G= ni µ i (9.15)
i=1

This important result shows that the chemical potential of species i is not really a new
state variable but is defined in terms of the familiar state variables, enthalpy, temperature
and entropy. The chemical potential of species i is its molar Gibbs free energy.

µ i = g i = hi T si (9.16)

The enthalpy in (9.16) includes the chemical enthalpy associated with the formation of the
species from its constituent elements.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-5

9.3 Criterion for equilibrium

The Gibbs free energy is sometimes described as the ”escaping tendency” of a substance.
At low temperatures the enthalpy dominates. A chemical species with a positive enthalpy
would like to break apart releasing some of its chemical enthalpy as heat and producing
products with lower enthalpy. A few examples are ozone (O3 ), hydrogen peroxide (H2 O2 ),
and nitrous oxide (N2 O). These are stable chemicals at room temperature but will de-
compose readily if their activation energy is exceeded in the presence of a heat source or a
catalyst. The entropy of any substance is positive and at high temperatures the entropy
term dominates the Gibbs free energy. In a chemical reaction the Gibbs free energy of
any species or mixture will increasingly tend toward a state of higher entropy and lower
Gibbs free energy as the temperature is increased. Take the di↵erential of the Gibbs free
energy.

dG = dE + P dV + V dP T dS SdT (9.17)

For a process that takes place at constant temperature and pressure dT = dP = 0. The
Second Law (9.5) leads to the result that for such a process

dG = dE + P dV T dS  0. (9.18)

A spontaneous change of a system at constant temperature and pressure leads to a decrease


of the Gibbs free energy. Equilibrium of the system is established when the Gibbs free
energy reaches a minimum. This result leads to a complete theory for the equilibrium of a
reacting system.

9.4 The entropy of mixing

Consider the adiabatic system shown in Figure 9.1 consisting of a set of (n1 , n2 , . . . , ni , . . . , nI )
moles of gas species segregated into volumes of various sizes such that the volumes are all
at the same temperature and pressure.

Figure 9.1: A system of gases separated by partitions.


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-6

The total number of moles in the system is

I
X
N= ni . (9.19)
i=1

The entropy per unit mole of an ideal gas is determined using the Gibbs equation

dT dP
ds = Cp Ru (9.20)
T P

where the units of CP are Joules/(mole Kelvin). Tabulations of gas properties are always
defined with respect to a reference temperature and standard pressure. The reference
temperature is universally agreed to be Tref = 298.15 K and the standard pressure is

P = 105 N/m2 = 105 P ascals = 102 kP a = 1 bar. (9.21)

All pressures are referred to P and the superscript ’ ’ denotes a species property evalu-
ated at standard pressure. A cautionary note: In 1999 the International Union of Pure and
Applied Chemistry (IUPAC) recommended that for evaluating the properties of all sub-
stances, the standard pressure should be taken to be precisely 100kP a. Prior to this date,
the standard pressure was taken to be one atmosphere at sea level, which is 101.325kP a.
Tabulations prior to 1999 are standardized to this value. The main e↵ect is a small change
in the standard entropy of a substance at a given temperature tabulated before and after
1999. There are also small di↵erences in heat capacity and enthalpy as well. The IUPAC
continues to provide standards for chemistry calculations and chemical nomenclature.
The pressure has no e↵ect on the heat capacity of ideal gases, and for many condensed
species the e↵ect of pressure on heat capacity is relatively small. For this reason, tabulations
of thermodynamic properties at standard pressure can be used to analyze a wide variety
of chemical phenomena involving condensed and gas phase mixtures. Inaccuracies occur
when evaluations of thermodynamic properties involve phase changes or critical phenomena
where wide deviations from the ideal gas law occur, or condensed phases exhibit significant
compressibility.
Integrating (9.20) from the reference temperature at standard pressure, the entropy per
unit mole of the ith gas species is

Z T ✓ ◆
dT P
si (T, P ) si (Tref ) = Cpi (T ) Ru ln . (9.22)
Tref T P
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-7

where the molar heat capacity Cpi is tabulated as a function of temperature at standard
pressure. The standard entropy of a gas species at the reference temperature is

Z Tref
dT
si (Tref ) = C pi (T ) + si (0) (9.23)
0 T

where the integration is carried out at P = P . To evaluate the standard entropy, heat
capacity data is required down to absolute zero. For virtually all substances, with the
exception of superfluid helium (II), the heat capacity falls o↵ rapidly as T ! 0 so that
the integral in (9.23) converges despite the apparent singularity at T = 0. From the third
law, the entropy constant at absolute zero, si (0), is generally taken to be zero for a pure
substance in its simplest crystalline state. For alloys and pure substances such as CO
where more than one crystalline structure is possible, the entropy at absolute zero may be
nonzero and tabulated entropy data for a substance may be revised from time to time as
new research results become available. Generally the entropy constant is very small.
The entropy per unit mole of the ith gas species is
✓ ◆
P
si (T, P ) = si (T ) Ru ln . (9.24)
P

The entire e↵ect of pressure on the system is in the logarithmic term of the entropy. The
extensive entropy of the whole system before mixing is

I
X I
X I
X ✓ ◆
P
Sbef ore = ni si (T, P ) = ni si (T ) ni Ru ln . (9.25)
P
i=1 i=1 i=1

Figure 9.2: System of gases with the partitions removed at the same pressure, temperature
and total volume as in Figure 9.1.

If the partitions are removed as shown in Figure 9.2 then, after complete mixing, each gas
takes up the entire volume and the entropy of the ith species is
✓ ◆
Pi
si (T, Pi ) = si (T ) Ru ln . (9.26)
P
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-8

where Pi is the partial pressure of the ith species. The mixture is ideal so there is no
enthalpy change during the mixing. If the pressure was so high that the potential energy
associated with inter-molecular forces was significant then the enthalpy of mixing would
be non-zero.
The entropy of the system after mixing is

I
X I
X I
X ✓ ◆
Pi
Saf ter = ni si (T, P ) = ni si (T ) ni Ru ln . (9.27)
P
i=1 i=1 i=1

The change of entropy due to mixing is

I
X I
X ✓ ◆! I
X I
X ✓ ◆!
Pi P
Saf ter Sbef ore = ni si (T ) ni Ru ln ni si (T ) ni Ru ln .
P P
i=1 i=1 i=1 i=1
(9.28)
Cancel common terms in (9.28).

I
X ✓ ◆
P
Saf ter Sbef ore = Ru ni ln >0 (9.29)
Pi
i=1

Mixing clearly leads to an increase in entropy. To determine the law that governs the
partial pressure let’s use the method of Lagrange multipliers to seek a maximum in the
entropy after mixing subject to the constraint that

I
X
P = Pi . (9.30)
i=1

That is, we seek a maximum in the function

I I ✓ ◆ I
!
X X Pi X
W (T, n1 , n2 , ..., nI , P1 , P2 , ..., PI , ) = ni s i (T ) ni Ru ln + Pi P
P
i=1 i=1 i=1
(9.31)
where is an, as yet unknown, Lagrange multiplier. The temperature of the system and
number of moles of each species in the mixture are constant. Di↵erentiate (9.31) and set
the di↵erential to zero for an extremum.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-9

@W @W @W @W
dW = dP1 + dP2 + ..... + dPI + d =0 (9.32)
@P1 @P2 @PI @

Now

I ✓ ◆ I
! I
!
X dPi X X
dW = ni Ru + dPi +d Pi P = 0. (9.33)
Pi
i=1 i=1 i=1

The last term in (9.33) is zero by the constraint and the maximum entropy condition
becomes.

I ✓
X ◆
ni Ru
+ dPi = 0 (9.34)
Pi
i=1

Since the dPi are completely independent, the only way (9.34) can be satisfied is if the
Lagrange multiplier satisfies

ni Ru
= (9.35)
Pi

for all i. In the original, unmixed, system each species satisfies the ideal gas law.

P Vi = ni R u T (9.36)

Using (9.35) and (9.36) we can form the sum

I
X I
X P Vi
Pi = . (9.37)
T
i=1 i=1

Finally the Lagrange multiplier is

V
= (9.38)
T
I
X
where, V = Vi . Using (9.35) and (9.38) the partial pressure satisfies
i=1

Pi V = n i R u T (9.39)
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-10

which is Dalton’s law of partial pressures. What we learn from this exercise is, not only
that the entropy increases when the gases mix, but that the equilibrium state is one where
the entropy is a maximum. Using Dalton’s law, the mole fraction of the ith gas species is
related to the partial pressure as follows

ni Pi
xi = = (9.40)
N P

The entropy of a mixture of ideal gases expressed in terms of mole fractions is

I
X ✓ ◆
P
Sgas = ni s igas (T ) Ru ln (xi ) N Ru ln (9.41)
P
i=1

and the entropy change due to mixing, (9.29), is expressed as

I
X
Saf ter Sbef ore = N Ru xi ln (xi ) > 0. (9.42)
i=1

9.5 Entropy of an ideal mixture of condensed species

The extensive entropy of an ideal mixture of condensed species (liquid or solid) with mole
numbers n1 , . . . , nI is S (T, P, n1 , . . . , nI ). If sipure (T, P ) is the entropy of the pure form of
the ith component, then in a system where the mole numbers are fixed, the di↵erential of
the entropy is

I
X
dS (T, P, n1 , . . . , nI ) = ni dsipure (T, P ) (9.43)
i=1

since dn1 = 0, . . . , dnI = 0. If we integrate (9.43) the result is

I
X
S (T, P, n1 , . . . , nI ) = ni sipure (T, P ) + C (n1 , . . . , nI ) (9.44)
i=1

with a constant of integration that is at most a function of the mole numbers.


In order to determine this constant we will use an argument first put forth by Max Planck
in 1932 and also described by Enrico Fermi in his 1956 book Thermodynamics (page 114).
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-11

Since the temperature and pressure in (9.44) are arbitrary let the pressure be reduced and
the temperature be increased until all of the condensed species in the system are fully
vaporized and behave as ideal gases as shown in Figure 9.3.

Figure 9.3: System of condensed species with the temperature increased and pressure de-
creased to fully vaporize the mixture.

Since the number of moles of each constituent has not changed, the constant of integration
must be the same. The entropy of a system of ideal gases was discussed in the previous
section and is given by equation (9.41).

I
X ✓ ✓ ◆◆ I
X ⇣n ⌘
P i
Sgas = ni s igas (T ) Ru ln ni Ru ln (9.45)
P N
i=1 i=1

Comparing (9.44) with (9.45) we can conclude that the constant of integration must
be

I
X ⇣n ⌘
i
C (n1 , . . . , nI ) = Ru ni ln . (9.46)
N
i=1

Therefore, the entropy of a mixture of condensed species is

I
X
S (T, P, n1 , . . . , nI ) = ni sipure (T, P ) Ru ln (xi ) (9.47)
i=1

where the mole fraction, xi = ni /N is used. Finally, the contribution of each component
to the extensive entropy of the mixture of condensed species is

si (T, P ) = sipure (T, P ) Ru ln (xi ) . (9.48)


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-12

The form of this equation is very similar to that for gases and comes as something of a
surprise since it involves the ideal gas constant which would seem to have no particular
relevance to a liquid or a solid. According to Kestin (A Course in Thermodynamics,
Ginn Blaisdell 1966) when (9.48) was first introduced it was ”met with general incredulity,
but its validity has since been confirmed experimentally beyond any reasonable doubt
whatever.”
It can be noted that if the gases in the mixing problem described in the last section were
replaced by ideal liquids, and those liquids were allowed to evolve from the unmixed to the
mixed state, the entropy change would be given by an expression identical to (9.29).

I
X
(Saf ter Sbef ore )ideal liquids = N Ru xi ln (xi ) (9.49)
i=1

9.6 Thermodynamics of incompressible liquids and solids

For a single homogeneous substance the Gibbs equation is

dE P
dS = + dV. (9.50)
T T

If the substance is an incompressible solid or liquid the Gibbs equation reduces to

dT
dS = Cv (T ) (9.51)
T

and the entropy is

Z T
dT
S Sref = Cv (T ) . (9.52)
Tref T

The entropy and internal energy of an incompressible substance depend only on its temper-
ature. The constant volume and constant pressure heat capacities are essentially equivalent
Cp (T ) = Cv (T ) = C (T ).
Since liquids and solids tend to be nearly incompressible, the entropy sipure (T, P ) tends to
be independent of pressure and in most circumstances one can use

sipure (T, P ) = si (T ) . (9.53)


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-13

Now the entropy of a condensed species in a mixture really does resemble that of an ideal
gas, but without the dependence on mixture pressure.

si (T, P ) = si (T ) Ru ln (xi ) (9.54)

For a general substance, the di↵erential of the Gibbs function is

dG = dE + P dV + V dP T dS SdT = SdT + V dP (9.55)

where the Gibbs equation (9.4) has been used. The cross derivative test applied to the
right side of (9.55) produces the Maxwell relation
✓ ◆ ✓ ◆
@S @V
= . (9.56)
@P T @T P

For an incompressible material, the entropy is independent of the pressure and (9.56)
implies that the coefficient of thermal expansion of the material is also zero. Therefore, for
an incompressible substance
✓ ◆
@V
=0
✓ @P ◆T
@V
=0
@T P
V = constant (9.57)
CP = C V = C
dE = CdT
C
dS = dT
T
dH = CdT + V dP.

On a per unit mole basis, the enthalpy of an incompressible material referenced to the
standard state is

h(T, P ) = h (T ) + (P P )v (9.58)

where v = constant is the molar volume of the material. The superscript on v is not
really necessary since it assumed to be a constant at all pressures and temperatures but
as a practical matter, when using (9.58) as an approximation for a real condensed solid or
liquid, the value of the molar density will be taken to be at the standard pressure 105 P a
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-14

and the reference temperature 298.15K. Often the pressure term in (9.58) is neglected
when dealing with reacting systems of gases and condensed materials. The reason is that,
while the molar enthalpy of the condensed and gaseous form of a species are of the same
order, the molar volume of a solid or liquid is generally two to three orders of magnitude
smaller than the gaseous molar volume.

9.7 Enthalpy

The enthalpy per unit mole of a gas is determined from

dh = Cp (T )dT. (9.59)

The enthalpy of a gas species is

Z T
hi (T ) hi (Tref ) = hi (T ) hi (Tref ) = Cpi (T )dT. (9.60)
T ref

In principle the standard enthalpy of the ith gas species at the reference temperature could
be taken as

Z Tref
hi (Tref ) = Cpi dT + hi (0). (9.61)
0

In this approach the enthalpy constant is the enthalpy change associated with chemical
bond breaking and making that occurs when the atoms composing the species are brought
together from infinity to form the molecule at absolute zero. Note that even for an atomic
species, the enthalpy constant is not exactly zero. A quantum mechanical system contains
energy or enthalpy arising from ground state motions that cannot be removed completely
even at absolute zero temperature. In practice, enthalpies for most substances are tabulated
as di↵erences from the enthalpy at the reference temperature of 298.15K which is much
more easily accessible than absolute zero so the question of the zero point enthalpy rarely
comes up. Thus the standard enthalpy of a species is

Z T
hi (T ) = Cpi (T )dT + hf i (Tref ) (9.62)
T ref

where hf i (T ) is the enthalpy change that occurs when the atoms of the species are
brought together from at infinity at the finite temperature T . The enthalpy including the
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-15

heat of formation (9.62) is sometimes called the complete enthalpy. In practice certain con-
ventions are used to facilitate the tabulation of the heat of formation of a substance.

9.7.1 Enthalpy of formation and the reference reaction

The enthalpy of formation of a substance, denoted hf (Tref ), is defined as the enthalpy


change that occurs when one mole of the substance is formed from its elements in their
reference state at the given temperature T and standard pressure P . The reference state
for an element is generally taken to be its most stable state at the given temperature and
standard pressure. The reference reaction for a substance is one where the substance is the
single product of a chemical reaction between its elements in their most stable state.
This convention for defining the heat of formation of a substance is useful even if the
reference reaction is physically unlikely to ever actually occur. A consequence of this
definition is that the heat of formation of a pure element in its reference state at any
temperature is always zero. For example, the enthalpy of formation of any of the diatomic
gases is zero at all temperatures. This is clear when we write the trivial reaction to form,
for example hydrogen, from its elements in their reference state.

H2 ! H2 (9.63)

The enthalpy change is clearly zero. In fact the change in any thermodynamic variable
for any element in its reference state is zero at all temperatures. A similar reference
reaction applies to any of the other diatomic species O2 , N2 , F2 , Cl2 , Br2 , I2 , and the heat
of formation of these substances is zero at all temperatures. The most stable form of carbon
is solid carbon or graphite and the reference reaction is

C(s) ! C(s) (9.64)

with zero heat of formation at all temperatures.


The reference reaction for carbon dioxide at 298.15K is

C(s) + O2 ! CO2 hfCO (298.15) = 393.522 kJ/mole. (9.65)


2

Here the carbon is taken to be in the solid (graphite) form and the oxygen is taken to be
the diatomic form. Both are the most stable forms over a wide range of temperatures.
Even if the temperature is well above the point where carbon sublimates to a gas (3915 K)
and significant oxygen is dissociated, the heats of formation of C(s) and O2 remain zero
even though the most stable form of carbon at this temperature is carbon gas.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-16

The heats of formation of metal elements are treated a little di↵erently. The heat of
formation of crystalline aluminum is zero at temperatures below the melting point and the
heat of formation of liquid aluminum is zero at temperatures above the melting point. The
same applies to boron, magnesium, sulfur, titanium and other metals.
The enthalpy (9.62) is usually expressed in terms of tabulated data as

hi (T ) = hi (T ) = hf i (Tref ) + {hi (T ) hi (Tref )} . (9.66)

For a general reaction the enthalpy balance is

h (Tf inal ) =

Iproduct
X Ireactant
X ⇣ ⌘ (9.67)
niproduct hiproduct (Tf inal ) nireactant hireactant Tinitialireactant .
iproduct ireactant

So for example, to determine the heat of formation of CO2 at 1000 K where the initial
reactants are also at 1000 K, the calculation would be

hfCO (1000) =
2

⇣ n o⌘
hfCO (298.15) + hfCO (1000) hfCO (298.15)
2 2 2

⇣ n o⌘ (9.68)
hf C (298.15) + hfC (1000) hf C (298.15)
(s) (s) (s)

⇣ n o⌘
hfO (298.15) + hfO (1000) hfO (298.15) .
2 2 2

Putting in the numbers from tabulated data (See Appendix 2) gives

hfCO (1000) =
2 (9.69)
[ 393.522 + 33.397] [0 + 11.795] [0 + 22.703] = 394.623

which is the tabulated value of the heat of formation of carbon dioxide at 1000 K. Note that
the enthalpy of the reference reactants at the reaction temperature must be included in the
calculation of the heat of formation calculation. Further discussion of heats of formation
can be found in Appendix 1 and tables of thermo-chemical data for selected species can be
found in Appendix 2.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-17

9.8 Condensed phase equilibrium

The pressure and temperature of the system may be such that one or more or all of the
species may be evolving with their condensed phase. In this case it may be necessary to
vary the volume to keep the temperature and pressure constant until the system reaches
equilibrium.
Figure 9.4 depicts the various species of the system in several phases all of which are in con-
tact with each other. Let the total number of moles in each phase be N1 , N2 , . . . , Np , . . . , NP .
Generally only a subset of the species will be evolving with their condensed phase and so
there is a di↵erent maximal index Ip for each phase. E↵ects of surface tensions between
phases are ignored.

Figure 9.4: System of molecular species with several phases.

Note the total numbers of moles in each phase are related to the species mole numbers
by

I1
X I2
X IP
X
N1 = n1i N2 = n2i · · · NP = nP i . (9.70)
i=1 i=1 i=1

The mole fractions in each phase are

npi
xpi = . (9.71)
Np

In the gas phase, the partial pressures of the gas species add up to the mixture pres-
sure

I1
X
P = Pi . (9.72)
i=1
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-18

The mole fractions of the species in the gas phase are related to the partial pressures
by

Pi
x1i = . (9.73)
P

Finally the mole fractions in each phase add to one.

Ip
X
xpi = 1 (9.74)
i=1

The extensive Gibbs free energies in each phase are

( I1 ✓ ◆)
X P
G1 (T, P, n11 , n12 , n13 , ..., n1I1 ) = N1 x1i (g1i (T ) + Ru T ln (x1i )) + Ru T ln
P
i=1
I2
X
G2 (T, P, n21 , n22 , n23 , ..., n2I2 ) = N2 x2i (g2i (T, P ) + Ru T ln (x2i ))
i2 =1
..
.
IP
X
GP (T, P, nP 1 , nP 2 , nP 3 , ..., nP IP ) = NP xpi (gpi (T, P ) + Ru T ln (xpi ))
iP =1
(9.75)
where phase 1 is assumed to be the gas phase. In (9.75) we have assumed that all condensed
phases are ideal mixtures and the entropy per mole of the condensed phase of species i
does not depend on the pressure. This is a reasonable assumption for a condensed phase
that is approximately incompressible as was argued earlier in conjunction with equation
(9.58). There are a few examples of liquids such as liquid helium and liquid nitrous oxide
that are quite compressible and the assumption would break down. This treatment of the
condensed phases is similar to that used by Bill Reynolds in the development of STANJAN.
The NASA Glenn code CEA is a little di↵erent. In CEA, each condensed phase is treated
as a pure substance. For example, if there are two species that condense out as liquids, each
is treated as a separate, distinct phase. In the approach used by CEA the mole fractions
of each condensed species are one by definition. In CEA the pressure term in (9.58) is
neglected. See equation 2.11 in NASA Reference Publication 1311 (1994) by Gordon and
McBride.
In the case of a gas, the Gibbs free energy does depend on pressure through the dependence
of entropy on pressure. This is connected to the fact that, as quantum mechanics tells us,
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-19

the number of energy states that a gas can occupy (and therefore the entropy of the gas)
increases with the volume containing the gas (see Appendix 1 of the AA210a notes).
The Gibbs free energy of the whole system is

G (T, P, n11 , n12 , . . . , n1I1 , n21 , n22 , . . . , n2I2 , ..., nP 1 , nP 2 , ..., nP IP ) =

IP
P X
X (9.76)
npi gpi (T, P, npi ) .
p=1 i=1

With the gas phase written separately Equation (9.76) is

G (T, P, n11 , n12 , . . . , n1I1 , n21 , n22 , . . . , n2I2 , ..., nP 1 , nP 2 , ..., nP IP ) =


0 0 11
IP
P X
X XI1
n1i @g 1i (T ) + Ru T ln (n1i ) Ru T ln @ n1î AA+
p=1 i=1 î=1 (9.77)
0 0 11
IP
P X
X XIP
npi @gpi (T, P ) + Ru T ln (npi ) Ru T ln @ npî AA .
p=2 i=1 î=1

Di↵erentiate (9.77).

@G @G
dG = dT + dP +
@T @P
@G @G @G
dn11 + dn12 + · · · + dn1I1 +
@n11 @n12 @n1I1
(9.78)
@G @G @G
dn21 + dn22 + · · · + dn2I2 + . . . +
@n21 @n22 @n2I2

@G @G @G
dnP 1 + dnP 2 + · · · + dnP IP .
@nP 1 @nP 2 @nP IP

Written out fully (9.78) is


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-20

0 0 11
I1
X I1
X
dG = dn1i @g1i (T ) + Ru T ln (n1i ) Ru T ln @ n _ AA+
1i
i=1 _
0 0 11 i =1
P
I1

I1
X B ✓ ◆ B _ dn1_i CC ✓ ◆X I1
B dn1i B i =1 CC P
B
n1i BRu T B
Ru T B I C C
CC + Ru T ln P dn1i +
i=1 @ n1i @ P 1
n_ A A i=1
_ 1i
0 i =1 0 11
XI1 XI1
@g (T )
n1i @ 1i + Ru ln (n1i ) Ru ln @ n _ AA dT +
@T 1i
i=1 _
! ! i =1
X I1 ✓ ◆ XI1 ✓ ◆
dP P
n1i Ru T + n1i Ru ln dT +
P P
i=1 0 i=1 0 11
P X Ip Ip
X X
dnpi @gpi (T, P ) + Ru T ln (npi ) Ru T ln @ n _ AA +
pi
p=2 i=1 _
0 0 11 i =1
PIp
B ✓ ◆ B dn _ CC
XP X Ip B dnpi B_ p i CC
B B i =1 CC
npi BRu T Ru T B I CC+
B npi B P p CC
p=2 i=1 @ @ n _ AA
i=1 pi
0 0 11
Ip
P X
X Ip
X Ip
P X
X ✓ ◆
@gpi (T, P ) @gpi (T, P )
npi @ + Ru ln (n1i ) Ru ln @ n _ AA dT + npi dP .
@T 1i @T
p=2 i=1 _ p=2 i=1
i =1
(9.79)
Note that
0 1
Ip
P
B dn _ C
Ip
X B dn _ pi C
B pi i =1 C
npi B C = dNp dNp = 0. (9.80)
B npi PIp C
i=1 @ n _
A
_ pi
i =1

At constant temperature and pressure dT = dP = 0 and(9.79) becomes


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-21

0 0 1 1
I1
X I1
X ✓ ◆
P
dG = dn1i @g1i (T ) + Ru T ln (n1i ) Ru T ln @ n _ A + Ru T ln A+
1i P
i=1 _
0 0 i =1 11
P X Ip Ip
X X
dnpi @gpi (T, P ) + Ru T ln (npi ) Ru T ln @ n _ AA .
pi
p=2 i=1 _
i =1
(9.81)
In terms of mole fractions (9.81) reads

I1
X ✓ ✓ ◆◆
P
dG = dn1i g1i (T ) + Ru T ln (x1i ) + Ru T ln +
P
i=1
Ip
P X . (9.82)
X
dnpi (gpi (T, P ) + Ru T ln (xpi )).
p=2 i=1

At equilibrium dG = 0. For species that are only present in one phase, dnpi = 0 at
equilibrium. For those species that are in equilibrium with another phase, dnphase1i =
dnphase pi and equation dG = 0 can only be satisfied if

gphase1i = gphase pi . (9.83)

If phase 1 is a gas species (9.83) is

✓ ◆
P
gphase1i (T )+Ru T ln (xphase1i )+Ru T ln = gphase pi (T, P )+Ru T ln (xphase pi ) . (9.84)
P

At equilibrium, the Gibbs free energy (or chemical potential) of species i is the same
regardless of its phase. For example if gas species 1 is in equilibrium with its liquid phase,
then ggas1 (T, P ) = gliquid1 (T, P ). There can also be more than one solid phase and so the
total number of phases can exceed three. For example, in helium at very low temperature
there can be multiple liquid phases.
Suppose phase p in is a pure liquid so the mole fraction is one. Also assume the liquid is
incompressible so that the Gibbs function of the liquid is given by
✓ ◆
P
ggasi (T ) + Ru T ln (xgasi ) + Ru T ln = gliquidi (T ) + (P P )v liquid . (9.85)
P
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-22

Solve for the partial pressure of the species in the gas phase.
!
(P P )vliquid gliquid (T ) ggas (T )
Pi i i i
=e Ru T e Ru T (9.86)
P

The term in brackets is the classical form of the Clausius-Clapeyron equation that relates
the vapor pressure of a gas in equilibrium with its condensed phase to the temperature of
the system.

9.9 Chemical equilibrium, the method of element poten-


tials

If the species are allowed to react at constant temperature and pressure, the mole fractions
will evolve toward values that minimize the extensive Gibbs free energy of the system
subject to the constraint that the number of moles of each element in the mixture remains
fixed. The number of moles of each atom in the system is given by

Ip
P X
X
aj = npi Apij (9.87)
p=1 i=1

where Apij is the number of atoms of the jth element in the ith molecular species of the
pth phase. The appropriate picture of our system is shown in Figure 9.5.

Figure 9.5: System of reacting molecular species in several phases at constant temperature
and pressure with fixed number of moles of each element.

Generally the number of species in each phase will be di↵erent. This may be the case even
in a situation where, at a given temperature and pressure, there is phase equilibrium for
a given species. Consider graphite in equilibrium with its vapor at low temperature, the
mole fraction in the vapor phase may be so small as to be essentially zero and the species
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-23

may be rightly excluded from the vapor mixture. The volume required to maintain the
system at constant pressure and temperature must be allowed to vary.
The Gibbs free energy of the system is

P Ip
X X
G (T, P, n11 , n12 , ..., np1 , ..., nP IP ) = Np xpi gpi (T, P, xpi ) . (9.88)
p=1 i=1

With the gas phase written separately

I1
X ✓ ◆
P
G (T, P, n11 , n12 , ..., n21 , ..., nP IP ) = N1 x1i (g1i (T ) + Ru T ln (x1i )) + N1 Ru T ln +
P
i=1
P Ip .
X X
Np xpi (gpi (T, P ) + Ru T ln (xpi ))
p=2 i=1
(9.89)
We will use the method of Lagrange multipliers to minimize the Gibbs free energy subject
to the atom constraints. Minimize the function

W T, P, n11 , n12 , ..., n1I1 , n21 , n22 , ..., n2I2 , ...., np1 , np2 , ..., npIp , 1 , ..., J =

G T, P, n11 , n12 , ..., n1I1 , n21 , n22 , ..., n2I2 , ...., np1 , np2 , ..., npIp
0 1 (9.90)
J Ip
P X
X X
Ru T j @ npi Apij aj A .
j=1 p=1 i=1

where the J unknown Lagrange multipliers, j, are dimensionless. Our modified equilib-
rium condition is

@W @W @W @W @W @W
dW = dT + dP + dn11 +....+ dnpIP + d 1 +....+ d J = 0. (9.91)
@T @P @n11 @npI @ 1 @ J

Substitute (9.76) into (9.91) and impose dP = dT = 0.

Ip
P X j Ip
P X
X X X
dW = npi dgpi (T, P, xpi ) + gpi (T, P, xpi ) dnpi Ru T j dnpi Apij (9.92)
p=1 i=1 j=1 p=1 i=1
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-24

The order of the sums can be rearranged so (9.92) can be written as

0 1
Ip
P X Ip
P X J
X X X
dW = npi dgpi (T, P, xpi ) + @gpi (T, P, xpi ) Ru T j Apij A dnpi = 0.
p=1 i=1 p=1 i=1 j=1
(9.93)
The di↵erential of the molar Gibbs free energy is

@gpi @gpi dxpi


dgpi = dT + dP + Ru T . (9.94)
@T @P xpi

For a process that takes place at constant temperature and pressure

0 1
Ip
P X Ip
P X J
X dxpi X X
dW = Ru T npi + @gpi (T, P, xpi ) Ru T j Apij A dnpi = 0. (9.95)
xpi
p=1 i=1 p=1 i=1 j=1

The first sum in (9.95) can be re-written as follows

0 1
P Ip Ip
P X J
X X dxpi X X
dW = Ru T Np xpi + @gpi (T, P, xpi ) Ru T j Apij A dnpi =0
xpi
p=1 i=1 p=1 i=1 j=1
(9.96)
or

0 1
P Ip Ip
P X J
X X X X
dW = Ru T Np dxpi + @gpi (T, P, xpi ) Ru T j Apij A dnpi = 0. (9.97)
p=1 i=1 p=1 i=1 j=1

But the normalization conditions for the mole fractions of each phase imply that
0 1
Ip Ip
X X
dxpi = d @ xpi A = d (1) = 0. (9.98)
i=1 i=1

Finally our modified equilibrium condition is


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-25

0 1
Ip
P X J
X X
dW = @gpi (T, P, xpi ) Ru T j Apij A dnpi = 0. (9.99)
p=1 i=1 j=1

Since the dnpi are completely free, the condition (9.99) can only be satisfied if

J
X
gpi (T, P, xpi ) = Ru T j Apij . (9.100)
j=1

The Gibbs free energy of the system is

Ip J
P X
X X
G (T, P, n11 , n12 , ..., np1 , ..., nP IP ) = Ru T npi j Apij . (9.101)
p=1 i=1 j=1

Each atom in the mixture contributes equally to the extensive Gibbs free energy regardless
of which molecule it is in or which phase it is in. The molar Gibbs free energy of the ith
gas phase species is
✓ ◆
P
g1i (T, P ) = g1i (T ) + Ru T ln (x1i ) + Ru T ln . (9.102)
P

Insert into (9.100). For the gas phase species

✓ ◆ J
X
g1i (T ) P
+ ln (x1i ) + ln = j A1ij . (9.103)
Ru T P
j=1

For the condensed phase species

X J
gpi (T, P )
+ ln (xpi ) = j Apij p = 2, . . . , P . (9.104)
Ru T
j=1

Solve for the mole fraction of the ith species in the pth phase.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-26

8 9
< g (T ) ✓ ◆ X J =
1i P
x1i = Exp ln + j A1ij
: Ru T P ;
8 j=1
9 . (9.105)
< g (T, P ) X J =
pi
xpi = Exp + j Apij p = 2, . . . , P
: Ru T ;
j=1

The constraints on the atoms are

P Ip
X X
aj = Np xpi Apij . (9.106)
p=1 i=1

Substitute (9.105) into (9.106).


8 9
I1
X < g (T ) ✓ ◆ X J =
1i P
a j = N1 A1ij Exp ln + j1 A1ij1 +
: Ru T P ;
i=1 8 j1 =1
9 (9.107)
XP Ip
X < gpi (T, P ) X
J =
Np Apij Exp + A
j1 pij1 j = 1, ..., J.
: Ru T ;
p=2 i=1 j1 =1

Note that we have to introduce the dummy index j1 in the formula for xpi when we make
the substitution. The normalization conditions on the mole fractions give
8 9
< g (T ) ✓ ◆ J
X =
1i P
ln + j A1ij =1 (9.108)
: Ru T P ;
j=1

and
8 9
Ip
X < g (T, P ) XJ =
pi
Exp + j Apij =1 p = 2, ..., P . (9.109)
: Ru T ;
i=1 j=1

The total number of moles in the mixture is

P
X
Np = N. (9.110)
p=1
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-27

Equations, (9.107), (9.108), (9.109) and (9.110) are J + P + 1 equations in the unknowns
1 , ...., J , N1 , ..., NP and N .

As a practical matter, it is easier to compute the solution to equations (9.107), (9.108),


(9.109) and (9.110) by reformulating the equations to get rid of the exponentials. De-
fine

g1i (T )
B1i (T ) ⌘ Exp (9.111)
Ru T

and for the condensed species, p > 1



gpi (T, P )
Bpi (T, P ) ⌘ Exp . (9.112)
Ru T

In addition, define

yj = Exp ( j ) . (9.113)

The mole fractions become

✓ ◆ J
Y
P
x1i = B1i (yj )A1ij
P
j=1
J (9.114)
Y Apij
xpi = Bpi (yj ) p = 2, . . . , P.
j=1

The system of equations that needs to be solved now becomes


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-28

✓ ◆ I1
X J
Y P
X Ip
X J
Y
P
N1 A1ij B1i (yj1 )A1ij1 + Np Apij Bpi (yj1 )Apij1 = aj , j = 1, . . . , J
P
i=1 j1 =1 p=2 i=1 j1 =1
✓ ◆X
I1 J
Y
P
B1i (yj )A1ij = 1
P
i=1 j=1
Ip J
X Y
Bpi (yj )Apij = 1, p = 2, . . . , P
i=1 j=1
P
X
Np = N.
p=1
(9.115)
Note that in this formulation only the always positive yj are needed to determine the mole
fractions. The element potentials j , which are the logarithm of the yj , never actually need
to be calculated.
A key advantage of this formulation of the problem is that the equations that need to
be solved for the unknown yi and Np are multivariate polynomials, and algorithms are
available that enable the roots to be determined without requiring an initial guess of the
solution. Typically a number of real and complex roots are returned. The correct root is
the one with all positive real values of the yi and Np . In general, there is only one such
root.

9.9.1 Rescaled equations

The equations (9.115) admit an interesting scaling invariance where any constants, say ↵j ,
j = 1, . . . , J can be added to the normalized Gibbs free energies as long as the coefficients
and unknowns are transformed as

J
Y
Bpi = B̃pi (↵j )Apij
j=1 (9.116)
ỹj
yj = .
↵j

If the transformations (9.116) are substituted into (9.115) the system of equations remains
the same but expressed in tildaed variables. This leads to the following reformulation of
the problem.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-29

Rewrite (9.111) and (9.112) by adding and subtracting the standard Gibbs free energies of
the elements that make up the given species.

8 9
< g (T ) X J
g j (T ) X J
gj (T ) =
1i
B1i = Exp + A1ij A1ij
: Ru T Ru T Ru T ;
8 j=1 j=1 9 (9.117)
< g (T, P ) X J
gj (T, P ) XJ
gj (T, P ) =
pi
Bpi = Exp + Apij Apij p>1
: Ru T Ru T Ru T ;
j=1 j=1

The first two terms in the bracket in equation (9.117) constitute the Gibbs free energy of
formation of the given species from its individual elements
8 9
< g1i (T )
J
X gj (T ) =
B1i = Exp A1ij
: Ru T Ru T ;
8 j=1 9 (9.118)
< gpi (T, P )
J
X gj (T, P ) =
Bpi = Exp Apij p>1
: Ru T Ru T ;
j=1

where

J
X
g1i (T ) = g1i (T ) A1ij gj (T )
j=1
J (9.119)
X
gpi (T, P ) = gpi (T, P ) Apij gj (T, P ) p > 1.
j=1

We can write (9.118) as

J
Y ✓ ◆
gj (T )
B1i = B̃1i Exp A1ij
Ru T
j=1
J ✓ ◆ (9.120)
Y gj (T, P )
Bpi = B̃pi Exp Apij p>1
Ru T
j=1

where
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-30


g1i (T )
B̃1i = Exp
⇢ Ru T
(9.121)
gpi (T, P )
B̃pi = Exp p > 1.
Ru T

The system of equations (9.115) can now be written as

✓ ◆ I1
X J
Y P
X Ip
X J
Y
P
N1 A1ij B̃1i (ỹj1 )A1ij1 + Np Apij B̃pi (ỹj1 )Apij1 = aj , j = 1, . . . , J
P
i=1 j1 =1 p=2 i=1 j1 =1
✓ ◆X
I1 J
Y
P
B̃1i (ỹj )A1ij = 1
P
i=1 j=1
Ip J
X Y
B̃pi (ỹj )Apij = 1 p = 2, . . . , P
i=1 j=1
P
X
Np = N
p=1
(9.122)
where

✓ ◆
gj (T )
ỹj = yj Exp j = 1, . . . , J. (9.123)
Ru T

The mole fractions in scaled variables are

✓ ◆ J
Y
P
x1i = B̃1i (ỹj )A1ij
P
j=1
J (9.124)
Y Apij
xpi = B̃pi (ỹj ) p = 2, . . . , P.
j=1

Note that the governing equations (9.115) and (9.122) have exactly the same form and the
ỹj determine the mole fractions. The implication of this result is that the calculation of
mole fractions can be carried out either in terms of the standard Gibbs free energy of a
species or the Gibbs free energy of formation of the species.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-31

Generally the Gibbs free energy of formation of a substance is less than the Gibbs free
energy itself and so usually B̃pi < Bpi making the numerical solution of (9.122) simpler
than (9.115). This is especially important for gas calculations at low temperature where
g j /Ru T can be quite large producing values of the Bpi that can range over many orders
of magnitude. For many calculations, the Gibbs free energy of formation of a substance is
the only data tabulated. This is especially true for reactions in aqueous solution.

9.10 Example - combustion of carbon monoxide

If we mix carbon monoxide (CO) and oxygen (O2 ) at 105 P a and 298.15 K then ignite the
mixture, the result is a strongly exothermic reaction. The simplest model of such a reaction
takes one mole of CO plus half a mole of O2 to produce one mole of carbon dioxide.

1
CO + O2 ! CO2 (9.125)
2

But this model is not very meaningful without some information about the temperature
of the process. If the reaction occurs in an adiabatic system at constant pressure, the
final temperature is very high and at that temperature the hot gas consists of a mixture
of a number of species beside CO2 . A more realistic model assumes that the composition
includes virtually all of the combinations of carbon and oxygen that one can think of
including

C, CO, CO2 , O, O2 . (9.126)

Other more complex molecules are possible such as C2 and O3 but are only present in
extraordinarily low concentrations. For the composition (9.125) the temperature of the
mixture at one atmosphere turns out to be 2975.34 K. This is called the adiabatic flame
temperature.
Let’s use the minimization of the Gibbs free energy to determine the relative concentrations
of each molecular species for the mixture (9.126) at the equilibrium temperature 2975.34 K.
We will order the species as in (9.126). In this case all the species are in the gas phase and
the matrix of element coefficients Aij is shown in Figure 9.6.
For this system of molecular species, (9.107) and (9.108) lead to the following equations
governing the mole fractions and the total number of moles. Note that P = P in this
case.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-32

Figure 9.6: Matrix of element coefficients for the CO, O2 system.

✓ ✓ ◆ ✓ ◆
g1 g2
a1 = N A11 Exp + 1 A11
+ 2 A12 + A21 Exp + 1 A21
+ 2 A22 +
✓ Ru T ◆ ✓ Ru T ◆
g3 g4
A31 Exp + 1 A31 + 2 A32 + A41 Exp + 1 A41 + 2 A42 +
✓ Ru T ◆◆ Ru T
g5
A51 Exp + 1 A51 + 2 A52
Ru T
(9.127)

✓ ✓ ◆ ✓ ◆
g1 g2
a2 = N A12 Exp + 1 A11
+ 2 A12 + A22 Exp + 1 A21
+ 2 A22 +
✓ Ru T ◆ ✓ Ru T ◆
g3 g4
A32 Exp + 1 A31 + 2 A32 + A42 Exp + 1 A41 + 2 A42 +
✓ Ru T ◆◆ Ru T
g5
A52 Exp + 1 A51 + 2 A52
Ru T
(9.128)

✓ ◆ ✓ ◆
g1 g2
1 = Exp + 1 A11 + 2 A12
+ Exp + 1 A21 + 2 A22 +
✓ Ru T ◆
✓ Ru T ◆
g3 g4
Exp + 1 A31 + 2 A32 + Exp + 1 A41 + 2 A42 + (9.129)
✓ Ru T ◆ Ru T
g5
Exp + 1 A51 + 2 A52
Ru T

The unknowns in this system are the two Lagrange multipliers 1 and 2 corresponding
to each element in the mixture and the total number of moles . The number of moles of
carbon atoms is a1 = 1 and the number of moles of oxygen atoms is a2 = 2. The great
advantage of this method is that the number of unknowns is limited to the number of
elements in the mixture not the number of molecular species. Use (9.111) and (9.113) to
rewrite (9.127), (9.128) and (9.129) in the form of (9.115).
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-33

a1 /N = A11 B1 y1 A11 y2 A12 + A21 B2 y1 A21 y2 A22 + A31 B3 y1 A31 y2 A32 +


(9.130)
A41 B4 y1 A41 y2 A42 + A51 B5 y1 A51 y2 A52

a2 /N = A12 B1 y1 A11 y2 A12 + A22 B2 y1 A21 y2 A22 + A32 B3 y1 A31 y2 A32 +


(9.131)
A42 B4 y1 A41 y2 A42 + A52 B5 y1 A51 y2 A52

1 = B1 y1 A11 y2 A12 + B2 y1 A21 y2 A22 + B3 y1 A31 y2 A32 +


(9.132)
B4 y1 A41 y2 A42 + B5 y1 A51 y2 A52

According to (9.116) these equations can be rescaled as follows. Let

B1 = B̃1 ↵1 A11 ↵2 A12


B2 = B̃2 ↵1 A21 ↵2 A22
B3 = B̃3 ↵1 A31 ↵2 A32 (9.133)
B4 = B̃4 ↵1 A41 ↵2 A42
B5 = B̃5 ↵1 A51 ↵2 A52

and

ỹ1
y1 =
↵1 (9.134)
ỹ2
y2 = .
↵2

If (9.133) and (9.134) are substituted into (9.130), (9.131) and (9.132) the resulting equa-
tions read exactly the same except in terms of tildaed variables. This invariance can be
exploited to reduce the range of magnitudes of the coefficients in the equation if needed.
Now

1 = N B 1 y1 + B 2 y1 y2 + B 3 y1 y2 2 (9.135)

2 = N B2 y1 y2 + 2B3 y1 y2 2 + B4 y2 + 2B5 y2 2 (9.136)

1 = B 1 y1 + B 2 y1 y2 + B 3 y1 y 2 2 + B 4 y2 + B 5 y2 2 (9.137)

where the coefficients are


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-34

gC
B1 = e Ru T
gCO
B2 = e Ru T
gCO
2
B3 = e Ru T (9.138)
gO
B4 = e Ru T
gO
2
B5 = e Ru T .

At this point we need to use tabulated thermodynamic data to evaluate the coefficients.
The Gibbs free energy of the ith molecular species is

gi (T ) = hf i (Tref ) + {hi (T ) hi (Tref )} T si (T ) . (9.139)

Data for each species is as follows (See appendix 2). In the same order as (9.126)

gC (2975.34) = 715.004 + 56.208 2975.34 (0.206054) = 158.131 kJ/mole


gCO (2975.34) = 110.541 + 92.705 2975.34 (0.273228) = 830.782 kJ/mole
gCO2 (2975.34) = 393.522 + 151.465 2975.34 (0.333615) = 1234.68 kJ/mole
gO (2975.34) = 249.195 + 56.1033 2975.34 (0.209443) = 317.866 kJ/mole
gO2 (2975.34) = 0.00 + 97.1985 2975.34 (0.284098) = 748.09 kJ/mole.
(9.140)
The universal gas constant in appropriate units is

3
Ru = 8.314472 ⇥ 10 kJ/mole K (9.141)

and Ru T = 24.7384 kJ/mole. Now the coefficients are


✓ ◆ ✓ ◆
gC 158.131
B1 = Exp = Exp = 1.67346 ⇥ 10 3
✓ Ru T ◆ ✓ 24.7382

gCO 830.782
B2 = Exp = Exp = 3.84498 ⇥ 1014
✓ Ru T ◆ ✓ 24.7382 ◆
gCO2 1234.68
B3 = Exp = Exp = 4.73778 ⇥ 1021 (9.142)
✓ Ru T ◆ ✓ 24.7382 ◆
gO 317.866
B4 = Exp = Exp = 3.80483 ⇥ 105
✓ Ru T ◆ ✓ 24.7382 ◆
g O2 748.090
B5 = Exp = Exp = 1.35889 ⇥ 1013 .
Ru T 24.7382
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-35

Notice the ill conditioned nature of this problem. The constants in (9.142) vary over many
orders of magnitude requiring high accuracy and very careful numerical analysis. The
scaling (9.133) is

B1 = B̃1 ↵1
B2 = B̃2 ↵1 ↵2
B3 = B̃3 ↵1 ↵2 2 (9.143)
B4 = B̃4 ↵2
B5 = B̃5 ↵2 2 .
p
3
Choose ↵2 = 1.35889 ⇥ 1013 so that B̃5 = 1 and ↵1 = 1.67346 ⇥ 10 so that B̃1 = 1.
The scaled coefficients are

3
B̃1 = B1 /↵1 = 1.67346 ⇥ 10 /1.67346
⇣ ⇥ 10 3 = 1p ⌘
B̃2 = B2 /↵1 ↵2 = 3.84498 ⇥ 1014 / 1.67346 ⇥ 10 3 1.35889 ⇥ 1013 = 6.23285 ⇥ 1010
B̃3 = B3 /↵1 ↵2 2 = 4.73778 ⇥ 10 21
p / 1.67346 ⇥ 10
3
1.35889 ⇥ 1013 = 2.08341 ⇥ 1011
B̃4 = B4 /↵2 = 3.80483 ⇥ 105 / 1.35889 ⇥ 1013 = 0.103215
B̃5 = B5 /↵2 2 = 1.35889 ⇥ 1013 /1.35889 ⇥ 1013 = 1.
(9.144)
Using this procedure we have reduced the range of the coefficients from 24 down to 11
orders of magnitude, a very significant reduction. The equations we need to solve are as
follows.
⇣ ⌘
1 = N B̃1 ỹ1 + B̃2 ỹ1 ỹ2 + B̃3 ỹ1 ỹ22 (9.145)

⇣ ⌘
2 = N B̃2 ỹ1 ỹ2 + 2B̃3 ỹ1 ỹ22 + B̃4 ỹ2 + 2B̃5 ỹ22 (9.146)

1 = B̃1 ỹ1 + B̃2 ỹ1 ỹ2 + B̃3 ỹ1 ỹ22 + B̃4 ỹ2 + B̃5 ỹ22 (9.147)

I used Mathematica to solve the system, (9.145), (9.146), and (9.147). The result is

11
ỹ1 = 1.42474 ⇥ 10
ỹ2 = 0.392402 (9.148)
n = 1.24144
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-36

At the mixture temperature T = 2975.34 K, the mole fractions of the various species
are

xC = B̃1 ỹ1 = 1.42474 ⇥ 10 11


xCO = B̃2 ỹ1 ỹ2 = 6.23285 ⇥ 1010 ⇥ 1.42474 ⇥ 10 11 ⇥ 0.392402 = 0.34846
xCO2 = B̃3 ỹ1 ỹ22 = 2.08341 ⇥ 1011 ⇥ 1.42474 ⇥ 10 11 ⇥ 0.3924022 = 0.45706 (9.149)
xO = B̃4 ỹ2 = 0.103215 ⇥ 0.392402 = 0.0405018
xO2 = B̃5 ỹ22 = 0.3924022 = 0.153979.

Note that there is almost no free carbon at this temperature. We could have dropped C
from the mixture (9.126) and still gotten practically the same result.

9.10.1 CO Combustion at 2975.34K using Gibbs free energy of forma-


tion.

The Gibbs free energies of formation of the various species are

gC (2975.34) = 250.012 kJ/mole


gCO (2975.34) = 365.761 kJ/mole
gCO2 (2975.34) = 395.141 kJ/mole (9.150)
gO (2975.34) = 55.8281 kJ/mole
gO2 (2975.34) = 0.0 kJ/mole.

These values can be found in the JANAF tables in Appendix 2. Using the Gibbs free
energies of formation, the coefficients of the element potential equations are
 
gC (2975.34) 250.012 5
B1 = Exp = Exp = 4.0821 ⇥ 10
 Ru T  24.7382
gCO (2975.34) 365.761
B2 = Exp = Exp = 2.63731 ⇥ 106
 Ru T 24.7382
gCO2 (2975.34) 395.141
B3 = Exp = Exp = 8.6486 ⇥ 106 (9.151)
 Ru T  24.7382
gO (2975.34) 55.8281
B4 = Exp = Exp = 0.104689
 Ru T  24.7382
gO2 (2975.34) 0.0
B5 = Exp = Exp = 1.00.
Ru T 24.7382

The results using Mathematica to solve this system are


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-37

7
y1 = 3.382049 ⇥ 10
y2 = 0.3935536 (9.152)
n = 1.24391

and the mole fractions are

xC = B1 y1 = 1.380743 ⇥ 10 11
xCO = B2 y1 y2 = 0.3509709
xCO2 = B3 y1 y2 2 = 0.4529426 (9.153)
xO = B4 y2 = 0.04120202
xO2 = B5 y2 2 = 0.15488445.

These results agree closely with the results in (9.149) and with calculations using CEA.
Now transfer heat out of the gas mixture to bring the temperature back to 298.15 K at one
atmosphere. The standard Gibbs free energies of the species at this temperature are

gC (298.15) = 669.54219 kJ/mole


gCO (298.15) = 169.46747 kJ/mole
gCO2 (298.15) = 457.25071 kJ/mole (9.154)
gO (298.15) = 201.15482 kJ/mole
gO2 (298.15) = 61.16531 kJ/mole

and Ru T = 2.47897 kJ/mole. The coefficients are


✓ ◆ ✓ ◆
gC 669.54219
B1 = Exp = Exp = 5.03442 ⇥ 10 118
✓ Ru T ◆ ✓ 2.47897◆
gCO 169.46747
B2 = Exp = Exp = 4.88930 ⇥ 1029
✓ Ru T ◆ ✓ 2.47897 ◆
gCO2 457.25071
B3 = Exp = Exp = 1.277622 ⇥ 1080 (9.155)
✓ Ru T ◆ ✓ 2.47897 ◆
gO 201.15482
B4 = Exp = Exp = 5.74645 ⇥ 10 36
✓ Ru T ◆ ✓ 2.47897

g O2 61.16531
B5 = Exp = Exp = 5.19563 ⇥ 1010 .
Ru T 2.47897

At this relatively low temperature the coefficients range over 198 orders of magnitude! De-
spite this a good modern solver should be able to find the solution of (9.135), (9.136) and
(9.137) for this difficult case even without the scaling used earlier. For example, Mathemat-
ica has a feature where the user can specify an arbitrary number of digits of precision for
the calculation. The solution of (9.135), (9.136) and (9.137) at this temperature is
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-38

40
y1 = 7.07098 ⇥ 10
21
ỹ2 = 3.32704 ⇥ 10 (9.156)
n = 1.0000.

The resulting mole fractions at T = 298.15 K are

xC = B1 y1 = 3.55983 ⇥ 10 157
xCO = B2 y1 y2 = 1.15023 ⇥ 10 30
xCO2 = B3 y1 y2 2 = 1.00000 (9.157)
xO = B4 y2 = 1.91187 ⇥ 10 56
xO2 = B5 y2 2 = 5.75116 ⇥ 10 31 .

Only when the mixture is brought back to low temperature is the reaction model (9.125)
valid. The enthalpy change for this last step is

h |mixture at 298.15K h |mixture at 2975.34K = 8.94 ⇥ 106 + 2.51 ⇥ 106 = 6.43 ⇥ 106 J/kg.
(9.158)
This is the chemical energy released by the reaction (9.125) and is called the heat of
reaction.

9.10.2 Adiabatic flame temperature

In the example in the previous section the products of combustion were evaluated at the
adiabatic flame temperature. This can be defined at constant volume or constant pressure.
For our purposes we will use the adiabatic flame temperature at constant pressure. Imagine
the reactants brought together in a piston-cylinder combination permitting the volume to
be adjusted to keep the pressure constant as the reaction proceeds. A source of ignition is
used to start the reaction that evolves to the equilibrium state defined by the equilibrium
species concentrations at the original pressure and at an elevated temperature called the
adiabatic flame temperature. In the process the Gibbs function is minimized and since the
process is adiabatic, the enthalpy before and after the reaction is the same.
The general enthalpy balance for a reaction is given in (9.67). Fully written out the balance
is
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-39

Iproduct n ⇣ ⌘o
X
h (Tf inal ) = niproduct hf iproduct (298.15) + hiproduct (Tf inal ) hiproduct (298.15)
iproduct
Ireactant
X
nireactant hf ireactant (298.15) + hireactant (Tireactant ) hireactant (298.15) .
ireactant
(9.159)
If the reaction takes place adiabatically then h (Tf inal ) = 0 and

Iproduct n ⇣ ⌘o
X
niproduct hf iproduct (298.15) + hiproduct (Tf inal ) hiproduct (298.15) =
iproduct
Ireactant
X
nireactant hf ireactant (298.15) + hireactant (Tireactant ) hireactant (298.15) .
ireactant
(9.160)
Equation (9.160) can be solved along with (9.135), (9.136) and (9.137) to determine the
final temperature of the mixture along with the mole fractions and total number of moles.
In the carbon monoxide combustion example of the previous section we would write

nC hf C (298.15) + (hC (Tf inal ) hC (298.15)) +


nCO hf CO (298.15) + (hCO (Tf inal ) hCO (298.15)) +
nCO2 hf CO2 (298.15) + hCO2 (Tf inal ) hCO2 (298.15) +
(9.161)
nO hf O (298.15) + (hO (Tf inal ) hO (298.15)) +
n O2 hf O2 (298.15) + hO2 (Tf inal ) hO2 (298.15) =
nCO hf CO (298.15) + nO2 hf O2 (298.15) .

The enthalpy of the reactants is

nCO hf CO (298.15) + nO2 hf O2 (298.15) =


3
1.0 kgmole 110.527 ⇥ 10 kJ/kgmole + 0.5 kgmole {0 kJ/kgmole} = (9.162)
110.527 ⇥ 103 kJ.

On a per unit mass basis the enthalpy of the reactant mixture is

110.527 ⇥ 106 J
= 2.5117 ⇥ 106 J/kg. (9.163)
1 ⇥ 28.014 + 0.5 ⇥ 31.98
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-40

The enthalpy per unit mass of the product mixture at various temperatures is plotted in
Figure 9.7.

Figure 9.7: Enthalpy of the product mixture as a function of temperature.

As the products of combustion are cooled from 4000 K the enthalpy decreases monitonically.
The only temperature where the enthalpy of the product mixture matches that of the
original reactants is the adiabatic flame temperature, 2975.34 K.

9.10.3 Isentropic expansion

Now consider an isentropic expansion from a known initial state, (Tinitial , Pinitial ) to a final
state (Tf inal , Pf inal ) with the final pressure known. The condition that determines the
temperature of the final state is

S(Tf inal , Pf inal , n1f inal , n2f inal , n3f inal , ..., nIf inal ) =
(9.164)
S(Tinitial , Pinitial , n1initial , n2initial , n3initial , ..., nIinitial )

or

I
X I
X ✓ ◆
Pf inal
nif inal si (Tf inal ) Nf inal Ru xif inal ln xif inal Nf inal Ru ln =
P
i=1 i=1
I
X I
X ✓ ◆
Pinitial
niinitial si (Tinitial ) Ninitial Ru xiinitial ln (xiinitial ) Ninitial Ru ln .
P
i=1 i=1
(9.165)
Equation (9.165) can be solved along with (9.135), (9.136) and (9.137) to determine the
final temperature of the mixture after isentropic expansion along with the mole fractions
and total number of moles.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-41

For example, take the mixture from the previous section at the initial state, Tinitial = 2975.34 K
and Pinitial = 1 bar. The entropy of the system on a per unit mass basis is 8.7357kJ/kg K
which is essentially equivalent to the extensive entropy. Now expand the mixture to
Pf inal = 0.1 bar. If we calculate the entropy of the system at this pressure and various
temperatures, the result is the following plot.

Figure 9.8: Entropy of the product mixture as a function of temperature.

The temperature at which the entropy of the final state is the same as the initial state is
Tf inal = 2566.13 K.

9.10.4 Nozzle expansion

If we interpret the expansion just described as an adiabatic, isentropic expansion in a


nozzle we can use the conservation of stagnation enthalpy to determine the speed of the
gas mixture at the end of the expansion.

1
Hinitial = Hf inal + U 2 (9.166)
2
The initial enthalpy is taken to be the reservoir value. For this example the numbers
are

q q
U = 2 (Hinitial Hf inal ) = 2 ( 2.51162 + 3.94733) ⇥ 106 J/kg = 1694.53 m/ sec .
(9.167)
Ordinarily we are given the geometric area ratio of the nozzle rather than the pressure
ratio. Determining the exit velocity in this case is a little more involved. Here we need
to carry out a series of calculations at constant entropy and varying final pressure. For
each calculation we need to determine the density and velocity of the mixture and plot the
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-42

product ⇢U as a function of the pressure ratio. Beginning with the mixture at the adiabatic
flame temperature as the reservoir condition, the results are plotted below.

Figure 9.9: Mass flux in a converging-diverging nozzle as a function of nozzle static pressure
ratio.

The maximum mass flux occurs at the nozzle throat. Equate the mass flow at the throat
and the nozzle exit. For Pinitial /Pf inal = 10.0 the nozzle area ratio is

Ae ⇢t Ut 73.6768
= = = 2.46944. (9.168)
At ⇢e Ue 29.8354

This completes the specification of the nozzle flow. The case we have considered here is
called the shifting equilibrium case where the gas mixture is at equilibrium at every point
in the nozzle. One can also consider the case of frozen flow where the composition of the
gas mixture is held fixed at the reservoir condition.

9.10.5 Fuel-rich combustion, multiple phases

Suppose we choose a mixture of CO and O2 that has an excess of CO. In this case the
overall balance of carbon and oxygen can be satisfied in more than one way when the
products of the reaction are brought back to low temperature. For example, if we mix 2
moles of CO with 0.5 moles of O2 at a pressure of 105 N/m2 we could end up with either
of the following.

1
2CO + O2 ! CO2 + CO
2
(9.169)
1 3 1
2CO + O2 ! CO2 + C(gr)
2 2 2
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-43

Which balance actually occurs is of course determined by which one minimizes the Gibbs
free energy at the given temperature. The set of species in the mixture is now

C, CO, CO2 , O, O2 , C(gr) (9.170)

where we have allowed for the possible condensation of solid carbon. The matrices of
element coefficients Apij for the two phases are shown in Figure 9.10.

Figure 9.10: Matrix of element coefficients for the CO, O2 system with condensed species
(graphite).

The governing equations (9.122) become the following.

2 = N1 B11 y1 + B12 y1 y2 + B13 y1 y2 2 + N2 (B21 y1 ) (9.171)

3 = N1 B12 y1 y2 + 2B13 y1 y2 2 + B14 y2 + 2B15 y2 2 (9.172)

1 = B11 y1 + B12 y1 y2 + B13 y1 y2 2 + B14 y2 + B15 y2 2 (9.173)

1 = B21 y1 (9.174)

N = N1 + N2 (9.175)

The coefficients are


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-44

gC
B11 = e Ru T
gCO
B12 = e Ru T
gCO
2
B13 = e Ru T (9.176)
gO
B14 = e Ru T
gO
2
B15 = e Ru T

and

gC
(gr)
B21 = e Ru T . (9.177)

As an example, let’s solve for the various mole fractions at a mixture temperature of 940 K.
From tables, the standard Gibbs free energies at this temperature are

gC (940) = 558.95895 kJ/mole


gCO (940) = 309.37696 kJ/mole
gCO2 (940) = 613.34861 kJ/mole
(9.178)
gO (940) = 88.41203 kJ/mole
gO2 (940) = 206.32876 kJ/mole
gC(gr) (940) = 11.22955 kJ/mole.

The number of moles of each species in the mixture at T = 940 K are

nC = N1 B11 y1 = 4.11775 ⇥ 10 32
nCO = N1 B12 y1 y2 = 0.977065
nCO2 = N1 B13 y1 y2 2 = 1.011467
(9.179)
nO = N1 B14 y2 = 3.23469 ⇥ 10 22
nO2 = N1 B15 y2 2 = 1.027845 ⇥ 10 22

nC(gr) = N2 B21 y1 = 0.011467.

At this temperature, the mixture is predominately CO and CO2 with some C(gr) . Figure
9.11 shows the mole fraction of CO and C(gr) at a pressure of 105 N/m2 based on the total
number of moles in the mixture for a variety of mixture temperatures.
The excess CO begins to react and solid carbon begins to condense out at about T =
942.2 K. Below 600 K there is virtually no CO in the mixture.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-45

Figure 9.11: Condensation of graphite in the CO, O2 system.

9.11 Rocket performance using CEA

The equilibrium combustion package CEA (Chemical Equilibrium with Applications) from
NASA Glenn can also be used to perform equilibrium chemistry calculations and has a
capability similar to STANJAN but with a much wider range of chemicals with data based
on the current standard pressure. Some typical performance parameters for several pro-
pellant combinations at two chamber pressures are shown in Figure 9.12. The propellants
are taken to be at an equivalence ratio of one (complete consumption of fuel and oxidizer)
and so the exhaust velocity is not optimized. The numbers correspond to the e↵ective
exhaust velocity for the given chamber pressure and area ratio assuming vacuum ambient
pressure. The maximum e↵ective exhaust velocity generally occurs with a somewhat fuel
rich mixture that produces more low molecular weight species in the exhaust stream.

9.12 Problems

Problem 1 - A nuclear reactor is used to heat hydrogen gas in a rocket chamber to a


temperature of 4000 K. The pressure in the chamber is 100 atm. At these conditions a
significant fraction of the H2 is dissociated to form atomic H. What are the mole fractions
of H, H2 in the mixture? Relevant thermochemical data is provided in Figure 9.13 . The
reference temperature is 298.15 K.
Work the problem by hand and compare with CEA.
Problem 2 - Hydrogen gas heated to 4000 K is fed at a mass flow rate of 200 kg/sec into
a rocket chamber with a throat area of 0.2 m2 . The gas is exhausted adiabatically and
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-46

Figure 9.12: Some typical performance parameters for several propellant combinations at
two chamber pressures.

Figure 9.13: Hydrogen dissociation at 4000 K.


CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-47

isentropically through a nozzle with an area ratio of 40. Determine the exhaust velocity
for the following cases.
i) Expansion of undissociated H2 .
ii) Frozen flow expansion of dissociated H2 at the rocket chamber composition of H and
H2 .
(III) Shifting equilibrium expansion.
Problem 3 - An exotic concept for chemical rocket propulsion is to try to harness the
energy released when two atoms of hydrogen combine to form H2 . The idea is to store
the hydrogen atoms in solid helium at extremely low temperatures. Suppose a space en-
gine is designed with a very large area ratio nozzle. Let a 50-50 mixture by mass of
atomic hydrogen and helium be introduced into the combustion chamber at low temper-
ature. The propellant vaporizes and the hydrogen atoms react releasing heat. The gas
exhausts through the nozzle to the vacuum of space. Estimate the exhaust velocity of this
rocket.
Problem 4 - I would like you to consider the Space Shuttle Main Engine (SSME). Use
the thermochemical calculator CEA or an equivalent application to help solve the problem.
The specifications of the SSME are

Pt2 = 3000 psia


Tt2 = 3250 K (9.180)
Ae /At = 77.5

and

ṁH2 = 69 kg/ sec


(9.181)
ṁO2 = 400 kg/ sec .

i) Determine the adiabatic flame temperature and equilibrium mass fractions of the system
(H, H2 , HO, H2 O, O, O2 ) at the given chamber pressure.
ii) You will find that the temperature in (1) is higher than the specified chamber temper-
ature. Bring your mixture to the specified chamber pressure and temperature (in e↵ect
accounting for some heat loss from the engine).
Determine the exhaust velocity and sea level thrust assuming that the mixture remains at
equilibrium during the expansion process. Suppose the mass flow rates are changed to the
stoichiometric ratio
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-48

ṁH2 = 52.5 kg/ sec


(9.182)
ṁO2 = 416.5 kg/ sec.

How much does the exhaust velocity change?


Problem 5 - A monopropellant thruster for space applications uses nitrous oxide N2 O at
a low initial temperature as a monopropellant. The gas is passed through a catalyst bed
where it is decomposed releasing heat. The hot gas is expelled through a large area ratio
nozzle to the vacuum of space. The exhaust gas is composed of O2 and N2 . The enthalpy
of formation of N2 O at the initial temperature is 1.86 ⇥ 106 J/kg. Estimate the exhaust
velocity of this rocket.
Problem 6 - Element number three in the periodic table is Lithium which is a soft,
silvery white, highly reactive metal at room temperature. Because of its low atomic weight
it has been considered as a propellant for propulsion applications. Above 1615 K it is a
monatomic gas. At high temperatures and pressures the diatomic gas begins to form. At
the super extreme conditions of 7000 K and 105 bar almost 92% of the mixture is Li2 and
only 8% is Li. By the way, at these conditions the density of the mixture is 2.3⇥104 kg/m3 ,
greater than Uranium. Perhaps this is the di-Lithium crystal propellant that, according to
Gene Roddenberry, will in the future power generations of starships. Suppose this mixture
at an enthalpy hinitial = 38.4 ⇥ 106 J/kg expands isentropically from a rocket chamber to a
nozzle exit where the enthalpy is hf inal = 9.3 ⇥ 106 J/kg. What is the nozzle exit velocity?
What is the composition of the mixture at the nozzle exit?
Problem 7 - One mole per second of methane CH4 reacts with 2 moles per second of
O2 in a rocket chamber. The reaction produces products at 3000 K and 100 bar. Assume
the mixture contains (CO, CO2 , H, H2 , H2 O, O, OH, O2 ). Set up the system of equations
(9.115) for this problem. Use data from Appendix 2 to evaluate the coefficients in these
equations and solve for the mole fractions. Compare your results with CEA. The gas
exhausts isentropically to the vacuum of space through a nozzle with an area ratio of 80.
Determine the thrust.
Problem 8 - Ozone (O3 ) releases energy when it decomposes and can be used as a mono-
propellant for a space thruster. Let ozone be decomposed across a manganese oxide catalyst
bed and introduced into a thrust chamber. The pressure and temperature in the cham-
ber are 105 N/m2 and 2688 K. At these conditions the mixture is composed of only two
constituents, O2 and O. Relevant thermo-chemical data is given in Figure 9.14.
In the units used in this table the ideal gas constant is Ru = 0.00831451 kJ/ (mole K).
a) Determine the mole fractions of O2 and O in the mixture.
b) Determine the total number of moles in the mixture.
CHAPTER 9. THERMODYNAMICS OF REACTING MIXTURES 9-49

Figure 9.14: Ozone decomposition at 2688 K.

c) Determine the enthalpy per unit mass of the mixture.


d) If the gas is exhausted to the vacuum of space what is the maximum gas speed that
could be reached.

Vous aimerez peut-être aussi