Sunteți pe pagina 1din 35

CHAPTER 5: REACTION THERMODYNAMICS

Chemical reactions result in a rearrangement of atoms among molecules. Some


of the examples of chemical reaction are (1) Photosynthesis, (2) Metabolism, (3)
Combustion, (4) Petroleum refining, and (5) Plastics manufacturing. Energy
cannot be extracted from a system that is constrained unless the thermodynamic
constraint is removed. Fossil fuels such as coal, oil and natural gas are stored
below the earth’s surface and are in thermal, mechanical and chemical
equilibrium. If the chemical constraint is removed (allow the fuel to come in
physical contact with air), it is possible to release the chemical energy.

5.1 CHEMICAL REACTIONS AND COMBUSTION

5.1.1 STOICHIOMETRIC OR THEORETICAL REACTION

A stoichiometric or theoretical reaction results in the complete combustion of fuel.


For example, consider the reaction
CH 4 + aO2 → bCO2 + cH 2O (5.1)

involving one kmole of fuel and “a” kmoles of molecular oxygen. Species on the
left hand side of the reaction equation (Eq. (5.1)) are usually called the reactants
as they react and are consumed during the overall chemical reaction. Species on
the right hand side of Eq. (5.1), produced as a result of chemical reaction are
called products. In case of a steady flow reactor, the input and output streams
contain the following quantity in kmole of the elements.

Input Output
C 1 = b
H 4 = 2c
O 2a = 2b+c
We deduce from the carbon atom balance that b=1, from the H atom balance that
c=2, and from O balance that a=(2b+c)/2. Consequently, Eq. (5.1) can be written
in the form of the stoichiometric reaction
CH 4 + 2O2 → CO2 + 2 H 2O (5.2)

The coefficients (1, 2, 1 and 2) in Eq. (5.2) are called the stoichiometric
coefficients.

The minimum amount of air which provides sufficient oxygen for the complete
combustion of all the element like carbon, hydrogen, etc., which may oxidize is
called the theoretical or stoichiometric air. There is no oxygen in the products
when complete combustion (oxidation) is achieved with this theoretical air.

86
Let us consider again the combustion of 10 kmole of CH4, 25 kmoles of O2, 5
kmoles of CO2 and 3 kmole of H2O during which the fuel completely burns. Thus
only 10 kmole of CH4 are consumed in the reactor. Applying Eq. (5.2) we readily
determine that 20 kmole of O2 are utilized during the process, and hence 5 kmole
of O2, 15 (=10+5) kmole of CO2 and 23 (=20+3) kmole of H2O leave the
combustor. The stoichiometric coefficients do not necessarily represent the
amounts of species entering and leaving a reactor. From Eq. (5.2), the
stoichiometric air fuel ratio is

( A : F )mole = 2 + 2 × 3.76 ÷ 1 = 9.52 kmole of air to 1 kmole of fuel.


On mass basis, the A:F ratio is

( A : F )mole = 2 × 32 + 2 × 3.76 × 28 ÷ (1×16 ) = 17.16 kg of air/kg of fuel.


Atmospheric air contains 21% oxygen, 78% nitrogen, and 1% argon by volume.
In combustion calculations, however, the argon is usually neglected, and air is
assumed to consist of 21% oxygen and 79% nitrogen by volume (or molar basis).
On a mass basis, air contains 23% oxygen and 77% nitrogen. If we desire more
accuracy, air is a mixture of 78% N 2 , 1% Ar and 21% O2 . In this case,

( A : F )mole = 2 + 2 × 3.76 ÷ 1 = 9.52 kmole of air to 1 kmole of fuel.

( A : F )mass = 2 + 2 × 3.76 + 0.1× 40 ÷ (1×16 ) = 17.24 kg of air/kg of fuel.

5.1.2 REACTION WITH EXCESS AIR (LEAN COMBUSTION)

Fuel and air are often introduced separately (without premixing) into a
combustor, e.g., boiler, CI engine etc. Due to the large flow rates and short
residence time there is no assurance that each molecule of fuel is surrounded by
the appropriate number of oxygen molecules required for the stoichiometric
combustion. Therefore, it is customary to supply excess air in order to facilitate
better mixing and thereby ensure complete combustion. The excess oxygen
remains unburned and appears in the product.

5.1.3 REACTION WITH EXCESS FUEL (RICH MIXTURE)

Incomplete combustion occurs when the air supplied is less than the
stoichiometric amount required. For this condition, the products of incomplete
oxidation may contain a mixture of CO, CO2 , H 2 and H 2 O .

5.1.4 EQUIVALENT RATIO AND STOICHIOMETRIC RATIO

The equivalent ratio is defined as the actual fuel-air ratio to stoichimetric fuel-air
ratio.

87
F : Aactual A : Fstoichiometric ( O2 : F ) stoichiometric (5.3)
φ= = =
F : Astoichiometric A : Factual ( O2 : F )actual
The stoichiometric ratio is given by
1 (5.4)
SR =
φ
For example, if φ = 0.5 , for methane-air combustion, this implies that the
excess air is supplied for every kmole of fuel that is burned. In such case SR=2.
In otherwords, air supplied is two times as large as the stoichiometric or
theoretical air demand of the fuel. Eq. (5.2) can be written under such condition
as
⎛ 2⎞ ⎛1 ⎞ ⎛ 2⎞ (5.5)
CH 4 + ⎜ ⎟ ( O2 + 3.76 N 2 ) → CO2 + 2 H 2O + 2 ⎜ − 1⎟ O2 + ⎜ ⎟ × 3.76 N 2
⎝φ ⎠ ⎝φ ⎠ ⎝φ ⎠
5.2 DEGREE OF REACTION

Let us suppose that we have a mixture of four substances, A1, A2, A3 and
A4 capable of undergoing a reaction of the type
ZZX v3 A3 + v4 A4
v1 A1 + v2 A2 YZZ (5.6)
where the v ' are the stoichimetric coefficients.

Starting with arbitrary amounts of both initial and final constituents, let us imagine
that the reaction proceeds completely to the right with the disappearance of at
least one of the initial constituents, say, A1 . Then the original number of moles of
the initial constituents is given in the form
n1 (original) = n0v1 (5.7)

n2 (original) = n0v2 + N 2 (5.8)


where n0 is an arbitrary positive number, and N 2 is the residue(or excess) of
A2 , i.e., the number of moles of A2 which cannot combine. If the reaction is
assumed to proceed complete it to the left with the disappearance of the final
constituent A3 then
n3 (original) = n0′v3 (5.9)

n4 (original) = n0′v4 + N 4 (5.10)

88
where n0′ is an arbitrary positive number and N 4 is the excess number of moles
of A4 left after the reaction is complete from right to left. For a reaction that
occurred completely to the left, there is a maximum possible amount of each
initial constituent, and a minimum possible amount of each final constituent, so
that
n1 (max) = n0v1 + n0′v1 = (n0 + n0′ )v1 (5.11)
where n0v1 = Original number of moles of A1
n0′v1 = Number of moles of A1 formed by chemical reaction
(n0′v3 A3 + n0′v4 A4 → n0′v1 A1 + n0′v2 A2 )

n2 (max) = (n0v2 + N 2 ) + n0′v2 = (n0 + n0′ )v2 + N 2 (5.12)


where ( n0v2 + N 2 ) = Original number of moles of A2
n0′v2 = Number of moles of A2 formed by chemical reaction
The constituent A3 completely disappears by reaction, hence
n3 (min)= 0 (5.13)
The excess number of moles of A4 that are left after the reaction is complete to
left

n4 (min)= 0 (5.14)
Similarly, if the reaction is imagined to proceed completely to the right, there is a
minimum amount of each initial constituent, and a maximum amount of each final
constituent, so that
n1 (min) = 0 (5.15)

n2 (min) = N 2 (5.16)

n3 (max) = n0′v3 + n0v3 = (n0 + n0′ )v3 (5.17)

where n0′v3 = original amount


n0v3 = amount formed by chemical reaction
(n0v1 A1 + n0v2 A2 → n0v3 A3 + n0v4 A4 )
n4 (max) = (n0 + n0′ )v4 + N 4 (5.18)

Let us suppose that the reaction proceeds partially either to the right or to the left
to the extent that there are n1 moles of A1 , n2 moles of A2, n3 moles of A3 and
n4 moles of A4 . The degree (or advancement) of reaction ε is defined in terms
of any one of the initial constituents, say, A1 as the fraction

89
n1 (max) − n1 (5.19)
ε=
n1 (max) − n1 (min)
It is seen that when n1 = n1 (max), ε = 0 , the reaction will start from left to right.
When n1 = n1 (min), ε = 1, reaction is complete from left to right.

The degree of reaction can thus be written in the form


(n0 + n0′ )v1 − n1 (5.20)
ε=
(n0 + n0′ )v1
Therefore
n1 = (n0 + n0′ )v1 − (n0 + n0′ )v0ε = n (at start) − n (consumed) (5.21)
or,

n1 = Number of moles of A1 at start − number of moles of A1 (5.22)


consumed in the reaction = ( n0 + n0′ )v1 (1 − ε )

n2 = n (at start) − n (consumed) (5.23)


= (n0 + n0′ )v2 + N 2 − (n0 + n0′ )v2ε
= (n0 + n0′ )v2 (1 − ε ) + N 2

n3 = n (at start) + n (formed) (5.24)


= 0 + (n0 + n0′ )v3ε
= (n0 + n0′ )v3ε

n4 = n (at start) + n (formed) (5.25)


= N 4 + (n0 + n0′ )v4ε
= (n0 + n0′ )v4ε + N 4

The number of moles of the constituents change during a chemical reaction, not
independently but restricted by the above relations (Eqs. (5.22-5.25)). These
equations are the equations of constraint. The n ' s are functions of ε only. In a
homogeneous system, in a given reaction, the mole fraction x ' s is also function
of ε only, as illustrated below.

Let us take the reaction


(5.26)
ZZX 1 N 2 + 1 O2
NO YZZ
2 2

90
n0 n
in which n0 moles of NO dissociates to produce moles of N 2 and 0
2 2
moles of O2 . The n ' s and x ' s as function of ε are shown in the Table-5.1
given below

Table 5.1 Values of ν , n and x (Eq. (5.26))


A v n x
n1 1 − ε
A1 = NO v1 = 1 n1 = n0 (1 − ε ) x1 = =
∑n 1+ ε
1 n0 ε
A3 = N 2 v3 = n3 = ε x3 =
2 2 1+ ε
1 n ε
A4 = O2 v4 = n4 = 0 ε x4 =
2 2 1+ ε
∑ n = n0 (1 + ε )

If the reaction is imagined to advanced to an infinitesimal extent, the degree


of reaction changes from ε to ε + d ε , and the various n ' s will change by the
amounts (Upon differentiation of Eqs. (5.22-5.24) w.r.t. ε )

dn1 = −(n0 + n0′ )v1d ε


dn2 = −(n0 + n0′ )v2 d ε
dn3 = (n0 + n0′ )v3d ε
dn4 = (n0 + n0′ )v4 d ε
or,

dn1 dn2 dn3 dn4 (5.27)


= = = = (n0 + n0′ )d ε
−v1 −v2 v3 v4
which shows that the dn ' s are proportional to the v ' s.

5.3 REACTION EQUILIBRIUM


Let us consider a homogeneous phase having arbitrary amounts of the
constituents, A1, A2, A3 and A4 , capable of undergoing the reaction
ZZX v3 A3 + v4 A4
v1 A1 + v2 A2 YZZ (5.28)

91
The phase is at uniform temperature T and pressure p . The Gibbs function of
the mixture is
G=µn +µ n +µn +µ n
1 1 2 2 3 3 4 4
(5.29)

where the n ' s are the number of moles of the constituents at ant moment, and
the µ ' s are the chemical potentials. Let us imagine that the reaction is allowed to
take place at constant T and p . The degree of reaction changes by an
infinitesimal amount from ε to ε + d ε . The change in the Gibbs function is
dGT , p = ∑ µk dnk = µ1dn1 + µ 2 dn2 + µ3dn3 + µ 4 dn4 (5.30)
The equations of constraint in differential form are

dn1 = −(n0 + n0′ )v1d ε (5.31)

dn2 = −(n0 + n0′ )v2 d ε (5.32)

dn3 = (n0 + n0′ )v3d ε (5.33)

dn4 = (n0 + n0′ )v4 d ε (5.34)

On substitution of Eqs. (5.31-5.34) in Eq. (5.30),

dGT , p = (n0 + n0′ )(−v1µ1 − v2 µ2 + v3 µ3 + v4 µ4 )d ε (5.35)

From Eq. (5.35), following interpretations can be made,

(1) When the reaction proceeds spontaneously to the right, d ε is positive, and
since dGT , p < 0
(v1µ1 + v2 µ 2 ) > (v3 µ3 + v4 µ 4 ) (5.36)

(2) When the reaction proceeds spontaneously to the left, d ε is negative


(v1µ1 + v2 µ 2 ) < (v3 µ3 + v4 µ 4 ) (5.37)
i.e., ∑ vk µ k is positive.
(3) At equilibrium, the Gibbs function will be minimum, hence
v1µ1 + v2 µ2 = v3 µ3 + v4 µ 4 (5.38)
which is called the equation of reaction equilibrium. Therefore, it is the value of
∑ vk µk which causes or forces the spontaneous reaction and is called the
“chemical affinity”.

92
5.4 LAW OF MASS ACTION

For a homogeneous phase chemical reaction at constant temperature and


pressure, when the constituents are ideal gases, the chemical potential are given
by the expressions of the type
µ = RT (φ + ln p + ln x )
k k
(5.39)
k

where the φ ' s are functions of temperature only.


Substituting in the equation of reaction equilibrium (Eq. (5.38))
v1 (φ1 + ln p + ln x1 ) + v2 (φ2 + ln p + ln x2 ) (5.40)
= v3 (φ3 + ln p + ln x3 ) + v4 (φ4 + ln p + ln x4 )

On rearranging
v3 ln x3 + v4 ln x4 − v1 ln x1 − v2 ln x2 + (v3 + v4 − v1 − v2 )ln p (5.41)
= −(v3φ3 + v4φ4 − v1φ1 − v2φ2 )


x3v .x4v v + v −v −v
3 4 (5.42)
ln v v p 3
= −(v3φ3 + v4φ4 − v1φ1 − v2φ2 )
4 1 2

x1 .x2
1 2

Denoting
ln K = −(v3φ3 + v4φ4 − v1φ1 − v2φ2 ) (5.43)
where K , is known as the equilibrium constant, is a function of temperature only
⎡ x v3 x v4 ⎤ (5.44)
⎢ x v .x v ⎥ p
3 4 v + v −v −v
=K 3 4 1 2

⎣ 1 2 ⎦ ε =ε
1 2

This equation is called the law of mass action. K has the dimension of pressure
raised to the (v3 + v4 − v1 − v2 ) th power. Here the x ' s is the values of mole
fractions at equilibrium when the degree of reaction is ε e .
The law of mass action can also be written in this form
p3v . p4v
3 4 (5.45)
=K
p1v . p2v
1 2

where the p ' s are the partial pressures.

5.5 HEAT OF REACTION


The equilibrium constant K is defined by the expression
ln K = −(v3φ3 + v4φ4 − v1φ1 − v2φ2 ) (5.46)

Differentiating ln K with respect to T

93
d dφ dφ dφ dφ (5.47)
ln K = −(v3 3 + v4 4 − v1 1 − v2 2 )
dT dT dT dT dT

Let,

φ = 0 − ∫ ∫ p2 − 0
h 1 c dT s (5.48)

RT R T R
Therefore,

= − 02 − ∫ p 2
dφ h c dT (5.49)

dT RT RT
1 h
=
RT 2
( h0
+ ∫ c p
dT ) = −
RT 2
Therefore,
d 1 (5.50)
ln K = ( v h
3 3
+ v h
4 4
− v h
1 1
− v h
2 2
)
dT RT 2
where the h ' s refer to the same temperature T and the same pressure p . If v1
moles of A1 and v2 moles of A2 combine to form v3 moles of A3 and v4 moles of
A4 at constant temperature and pressure, the heat transferred would be, as
shown in Fig.5.1, equal to the final enthalpy (v3h3 + v4 h4 ) minus the initial
enthalpy (v1h1 + v2 h2 ). This is known as the heat of reaction, and is denoted by
∆H .
H

A1
V1 A3
V3

V2 A2 T,P V4 A4

Fig.5.1 Heat of Reaction

Hence, heat of reaction is


∆H = v3h3 + v4 h4 − v1h1 − v2 h2 (5.51)

94
Therefore,
d ∆H (5.52)
ln K =
dT RT 2
This is known as the van’t Hoff equation. This equation can be used to calculate
the heat reaction at any desired temperature or within a certain temperature
range. By rearranging Eq. (5.52)
d ln K ∆H (5.53)
=
dT R
T2
or,
d ln K ∆H (5.54)
=−
d (1/ T ) R
Therefore,
d log K d log K (5.55)
∆H = −2.303R = −19.148 kJ / kgmol
1 1
d( ) d( )
T T
If K1 and K 2 are the equilibrium constants evaluated at temperatures T1 and T2
respectively
log K1 − log K 2 (5.56)
∆H = −19.148
1 1

T1 T2
or,
TT K (5.57)
∆H = −19.148 1 2
log 1
T1 − T2 K2
If ∆H is positive, the reaction is said to be endothermic. If ∆H is negative, the
reaction is exothermic.

5.6 TEMPERATURE DEPENDENCE OF THE HEAT OF REACTION

∆H = v3h3 + v4 h4 − v1h1 − v2 h2 (5.58)

h = h0 + ∫ c p dT (5.59)
Therefore,

∆H = v3h03 + v4 h04 − v1h01 − v2 h02 + ∫ (v3c p 3 + v4c p 4 − v1c p1 − v2c p 2 )dT (5.60)

Denoting ∆H = v3h03 + v4 h04 − v1h01 − v2 h02


∆H = ∆H 0 + ∫ (v3c3 + v4c4 − v1c1 − v2c2 )dT (5.61)

95
If c p is known as a function of temperature and if at any temperature ∆H is
known, then at any other temperature, ∆H can be determined for a certain
chemical reaction from the above relation.
Some chemical reactions may be expressed as the result of two or more
reactions. If ∆H 0 is known for each of the separate reactions, then ∆H 0 of the
resultant reaction may be calculated.

For example,
ZZX H 2 + 1 O2
H 2O YZZ ∆H 0 = 239,250 J / gmol
2
ZZX CO + 1 O2
CO2 YZZ ∆H 0 = 279,890 J / gmol
2
ZZX CO2 + H 2
CO + H 2O YZZ ∆H 0 = −40640 J / gmol

5.7 TEMPERATURE DEPENDENCE OF THE EQUILIBRIUM CONSTANT

ln K = −(v3φ3 + v4φ4 − v1φ1 − v2φ2 ) (5.62)


where

φ = 0 − ∫ ∫ p 2 dT − 0
h 1 c dT s (5.63)

RT R T R
On substitution
1 (5.64)
ln K = (v3h03 + v4 h04 − v1h01 − v2 h02 )
RT
+ ∫ ∫ 3 p 3 4 p 4 2 1 p1 2 p 2
1 (v c + v c − v c − v c )dT
.dT
R T
1
+ (v3 s03 + v4 s04 − v1s01 − v2 s02 )
R
If
∆H 0 = v3h03 + v4 h04 − v1h01 − v2 h02 (5.65)
∆S0 = v3 s03 + v4 s04 − v1s01 − v2 s02

Then,

ln K = −
∆H 0 1
+ ∫ ∫ (v c
3 p3
+ v4c p 4 − v1c p1 − v2c p 2 )dT
dT +
∆S 0 (5.66)

RT R T2 R
This equation is sometimes called as the Nernst’s equation.

96
5.8 THERMAL IONIZATION OF A MONATOMIC GAS
One interesting application of Nernst’s equation was made by Dr. M.N. Saha to
the thermal ionization of a monatomic gas. If a monatomic gas is heated to a high
enough temperature (5000 K and above), some ionization occurs, with the
electrons in the outermost orbit being shed off, and the atoms, ions, and
electrons may be regarded as a mixture of three ideal monatomic gases,
undergoing the reaction
ZZX A+ + e
A YZZ (5.67)
Starting with n0 moles of atoms,
A v n x
1− εe
A1 = A v1 = 1 n1 = n0 (1 − ε e ) x1 =
1+ εe
εe
A3 = A+ v3 = 1 n3 = n0ε e x3 =
1+ εe
ε
A4 = e v4 = 1 n4 = n0ε 0 x4 = e
1 + εe
∑ n = n (1 + ε 0 e
)

The equilibrium constant is given by


⎧ x3v .x4v ⎫
3 4 (5.68)
ln K = ln ⎨ v v ⎬ . p v + v −v −v 3 4 1 2

⎩ x1 .x2 ⎭ε
1 2

or,

εe ε (5.69)
. e
1 + ε e 1 + ε e v + v −v ε e2
ln K = .p 3
= ln
4 1
.p
1− εe 1 − ε e2
1+ εe
5
Since the three gases are monatomic, c p = R which, on being substituted in the
2
Nernst’s equation, gives
ε e2 ∆H 0 5 (5.70)
ln . p = − + ln T + ln B
1 − ε e2 RT 2

∆S 0
where = ln B
R

97
Now,
ε e2 p ∆H 0 (5.71)
ln . 5/ 2 = −
1 − ε e T .B
2
RT

ε e2 T 5/ 2 (5.72)
∴ = Be −∆H 0 / RT
.
1 − εe 2
p
where ε e is the equilibrium value of the degree of ionization. This is known as
the Saha’s equation. For a particular gas the degree of ionization increases with
an increase in temperature and a decrease in pressure. It can be shown that
∆H 0 is the amount of energy necessary to ionize 1 gmole of atoms. If we denote
the ionization potential ∆H 0 of the atom in volts by E , then
Coulomb electron (5.73)
∆H 0 = E (volt ) ×1.59 × 10−19 × 6.06 × 1023
electron gmol
= 9.6354 × 10 E J / gmol
4

Equation (5.70) becomes


εe 96354 E 5 (5.74)
ln p = − + ln T + ln B
1 − ε e2 RT 2
Expressing p in atmospheres, changing to common logarithms and introducing
the values of B from statistical mechanics, Saha finally obtained the equation
ε2 96,354 E 5 ωω (5.75)
log e
p (atm) = − + log T + log i e − 6.491
1− ε 2
e
19.148T 2 ωa

where ωi , ωe and ωa are constants that refer to the ion, electron and atom
respectively. The value ωe for an electron is 2 . The value of E , ωi andωa for a
few elements are given below

Table 5.2 Values for E and ω

Element E , volts ωa ωi
5.12 2 1
Na
3.87 2 1
Ca
6.09 1 2
Cd
8.96 1 2
Zn
9.36 1 2

98
For alkali metals like Cs, Na, K etc., the ionization potential is less. It means less
energy is required to ionize one gmole of atoms. So these are used as seed for
magnetohydrodynamic power generation. Saha applied Eq. (5.70) to the
determination of temperature of stellar atmosphere. The spectrum of a star
contains lines which originate from atoms (arc lines) and those which originate
from ions (spark lines). A comparison of the intensities of an arc line and a spark
line from the same element gives a measure of ε e . Considering a star as a
sphere of ideal gas, the pressure of a star can be estimated. Thus, knowing
ε e , p, E and the ω ' s, temperature of star can be calculated.

5.9 GIBBS FUNCTION CHANGE

Molar Gibbs function of an ideal gas at temperature T and pressure p is given


by
g = RT (φ + ln p ) (5.76)
For the reaction of the type
ZZX v3 A3 + v4 A4
v1 A1 + v2 A2 YZZ (5.77)
The Gibbs function change of the reaction ∆G is defined by the expression

∆G = v3 g 3 + v4 g 4 − v1 g1 − v2 g 2 (5.78)
where the g ' s refer to the gases completely separated at T , p, ∆G is also
known as the free energy change. Substituting the values of the g ' s
∆ G = RT (v φ + v φ − v φ − v φ ) + RT ln p v3 + v4 −v1 −v2
3 3 4 4 1 1 2 2
(5.79)
But
ln K = −(v3φ3 + v4φ4 − v1φ1 − v2φ2 ) (5.80)

∴ ∆G = − RT ln K + RT ln p v + v −v −v
3 4 1 2 (5.81)
If p is expressed in atmospheres and ∆G is calculated from each gas is at a
pressure of 1 atm, the second term on the right drops out. Under these
conditions ∆G is known as the standard Gibbs function change and denoted by
∆G 0
(5.82)
∆G 0 = RT ln K
This is an important equation which relates the standard Gibbs function change
with temperature and the equilibrium constant. From this the equilibrium constant
can be calculated from changes in the changes in the standard Gibbs function, or
vice versa. For dissociation of water vapour,
1 (5.83)
ZZX H 2 + O2
H 2O YZZ
2
ln K 298 = −93.7.

99
Therefore,
∆G298
0
= −8.3143 × 298 × (−93.7) = 232,157 J / gmol (5.84)

Substituting ln K from Nernst’s equation


(v3c p 3 + v4c p 4 − v1c p1 − v2 c p 2 ) (5.85)
∆G 0 = ∆H 0 − T dT − T ∆S0
T2
from which also ∆G may be calculated directly. Values of ∆H 0 , ∆S0 and ∆G298
0 0

for fundamental ideal gas reactions are given in Table 5.3


+
Table 5.3 Fundamental ideal gas reactions

Re action ∆H 0, J / gmol ∆S0 J / gmol − K ∆G298


0
J / gmol
ZZX 2 H
H 2 YZZ 427,380 4.90 404,335

ZZX 1 H 2 + 1 Cl2 91,760


HCL YZZ −22.25 95,110
2 2
ZZX 1 H 2 + 1 Br2 50,280
HBr YZZ −24.05 54,050
2 2
1
ZZX H 2 + I 21
HI YZZ 5, 320 −21.00 8, 380
2 2
ZZX 1 N 2 + 1 O2
NO YZZ −90,500 −10.48 −87,570
2 2
ZZX H 2 + 1 O2
H 2O YZZ 239,250 14.70 232,545
2
ZZX H 2 + 1 S 2
H 2 S YZZ 80,450 6.90 71,230
2
ZZX CO + 1 O2
CO2 YZZ 279,890 18.69 258,940
2
ZZX NO + 1 O2
NO2 YZZ 59,500 11.44 37,290
2
ZZX 1 S 2 + O2
SO2 YZZ 349,025 3.77 329,330
2
ZZX 1 N 2 + 3 H 2
NH 3 YZZ 39,800 40.27 16,380
2 2
ZZX SO2 + 1 O2
SO3 YZZ 94,690 89.50 67,880
2
ZZX 2 NO2
N 2O4 YZZ 56,980 174.30 5, 030

100
Both ∆S0 and ∆G298 may be added and subtracted in the same manner as ∆H 0 .
0

For example,
ZZX H 2 + 1 O2
H 2O YZZ ∆G298
0
= 232,545 J / gmol
2
ZZX CO + 1 O2
CO2 YZZ ∆G0 = 258,940 J / gmol
2

ZZX CO2 + H 2
CO + H 2O YZZ ∆G298
0
= −26,395 J / gmol

∆G298
0
26,395
ln K 298 = = = 10.653
RT 8.3143 × 298
K 298 = 42,330

From this, the value of the degree of reaction at equilibrium ε e may be


calculated. From Eq.(5.35), if n0 = 1 and n0′ = 0

⎛ ∂G ⎞
⎜ ⎟ = (v3 µ3 + v4 µ4 − v1µ1 − v2 µ2 )
⎝ ∂ε ⎠T , p

Since µ k = RT (φk + ln p + ln xk )
And g k = RT (φk + ln p )
So µ k = g k + RT ln xk
Therefore
⎛ ∂G ⎞ x3v .x4v
3 4

⎜ ⎟ = v3 g 3 + v4 g 4 − v1 g1 − v2 g 2 + RT ln v v
⎝ ∂ε ⎠T , p x1 .x2
1 2

x3v .x4v
3 4

= ∆G + RT ln
x1v .x2v
1 2

At ε = 0, x3 = 0, x4 = 0
⎛ ∂G ⎞
⎜ ⎟ = −∞
⎝ ∂ε ⎠T , p
and at ε = 1, x1 = 0, x2 = 0

101
⎛ ∂G ⎞
⎜ ⎟ = +∞
⎝ ∂ε ⎠T , p
1 v v
At ε = , x1 = 1 , x2 = 2
2 ∑v ∑v
v v
x3 = 3 , x4 = 4
∑v ∑v
where ∑ v = v3 + v4 + v1 + v2
v3 v4
⎛ v3 ⎞ ⎛ v4 ⎞
⎜ ⎟ ⎜ ⎟
⎛ ∂G ⎞ ⎝ ∑ v⎠ ⎝ ∑v ⎠
⎜ ⎟ = ∆G + RT ln
⎝ ∂ε ⎠Tε =, p1
v v
⎛ v1 ⎞ ⎛ v2 ⎞
1 2

⎜ ⎟ ⎜ ⎟
⎝ ∑v ⎠ ⎝ ∑v ⎠
2

If p = 1 atm, T = 298 K
⎛ ∂G ⎞
⎟ p =1atm = ∆G298
0

⎝ ∂ε ⎠T = 298 K
1
ε=
2

⎛ ∂G ⎞ 1
The slope ⎜ ⎟ at ε = is called the “affinity” of the reaction, and it is equal
⎝ ∂ε ⎠T , p 2
to ∆G
0
at the standard reference state. The magnitude of the slope at
1
ε = (Fig.5.2) indicates the direction in which the reaction will proceed.
2
( δ G /δ ε ) = +α
T ,p

G ( δ G /δ ε ) = −α
T ,p

T,p

ε = 0 ε = εe ε = 1 /2 ε =1

Fig. 5.2 Plot of G against ε at constants T and p

102
For water vapour reaction, ∆G298 is a large positive number, which indicates the
0

1
equilibrium point is far to the left of ε = , and therefore, ε e is very small. Again
2
ZZX 1 1
for the reaction, NO YZZ N 2 + O2 , ∆G298 0
is a large negative value, which
2 2
1
shows that the equilibrium point is far to the right of ε = , and so ε e is close to
2
unity.

5.10 FUGACITY AND ACTIVITY

The differential of Gibbs function of an ideal gas undergoing an isothermal


process is

nRT (5.86)
dG = Vdp = dp = nRTd (ln p )
p

Analogously, the differential of Gibbs function for a real gas is


dG = nRTd (ln f ) (5.87)
where f is called the fugacity, first used by Lewis. The value of fugacity
approaches the value of pressure, as the letter tends to zero, i.e., when ideal gas
conditions apply. Therefore,

f (5.88)
lim
p →0
=1
p
For an ideal gas f = p. Fugacity has the same dimension as pressure.
Integrating Eq. (5.87)

f (5.89)
G − G 0 = nRT ln
f0

refer to the reference state when p0 = 1 atm. The ratio


0 0
where G and f
f / f 0 is called the activity.
Therefore,
(5.90)
G − G 0 = nRT ln a
For ideal gases, the equilibrium constant is given by
P3v .P4v 3 4 (5.91)
K= v v
P1 .P2 1 2

For real gases

103
3 4
f 3v . f 4v (5.92)
K real = v v 1 2
f1 . f 2

Similarly, it can show that


∆G 0 = −nRT ln K real (5.93)

and
d ln K real ∆H 0 (5.94)
=
dT RT 2

5. 11 ENTHALPY OF FORMATION

Let us consider the steady state steady flow combustion of carbon and oxygen to
from CO2 (Fig.5.3). Let the carbon and oxygen each enter the control volume at
250 C and 1 atm. Pressure, and the heat transfer be such that the product CO2
0
leaves at 25 C ,1 atm Pressure. The measured value of heat transfer is
−393,522 kJ / kgmol of CO2 formed. If H R and H P refer to the total enthalpy
of the reactants and products respectively, then the first law applied to the
reaction C + O2 → CO2 gives

C .V .

C
25 C ,
1 a tm 2 5 C , 1 a tm

C O 2
O 2

Q C .V . = -3 9 3 ,5 2 2 k j/k g m o l o f C O 2

Fig.5.3 Enthalpy of formation

For all the reactants and products in a reaction, the equation may be written as

n h + Q = n he
R
i C .V ∑
P
e
(5.95)
i

104
where R and P refer to the reactants and products respectively. The enthalpy of
0
all the elements at the standard reference state of 25 C , 1 atm is assigned the
value of zero. In the carbon-oxygen reaction, H R = 0. Hence, the energy
equation gives
Q = H = −393,522kJ / kg mol
C .V P
(5.96)
0
This is known as the enthalpy of formation of CO2 at 25 C , 1atm., and

designated by the symbol, h f ( ) 0

CO2
= −393,522 kJ / kgmol . In most cases,
0
however, the reactants and products are not at 25 C , 1atm , therefore, the
change of enthalpy (in case of constant pressure or S.S.S.F. process) between
250 C , 1atm. and the given state must be known. Thus the enthalpy at any
temperature and pressure, hT , p is

( )
h T , p = h 0f
298 K ,1atm
( )
+ ∆h
298 K ,1atm →T , p
(5.97)

For convenience, the subscripts are usually dropped, and


hT , p = h 0f + ∆ h (5.98)

where ∆ h represents the difference in enthalpy between any given state and the
enthalpy at 298.15 K , 1atm . Table 5.4 gives the values of the enthalpy of
formation of a number of substances in kJ / kgmol .

105
Table 5.4 Enthalpy of Formation, Gibbs Function of Formation, and Absolute
0
Entropy of Various Substance at 25 C , 1atm pressure

Molecular h 0f g 0f s0
Substance
Weight , M kJ / kgmol kJ / kgmol kJ / kgmolK
CO( g ) 28.011 −110529 −137150 197.653
CO2 ( g ) 44.001 −393522 −394374 213.795
H 2O ( g ) 18.015 −241827 −228583 188.833
H 2O(l ) 18.015 −285838 −237178 70.049
CH 4 ( g ) 16.043 −74873 −50751 186.256
C2 H 2 ( g ) 26.038 +226731 +209234 200.956
C2 H 4 ( g ) 28.054 +52283 +68207 219.548
C2 H 6 ( g ) 30.070 −84667 −32777 229.602
C3 H 8 ( g ) 44.097 −103847 −23316 270.019
C4 H10 ( g ) 58.124 −126148 −16914 310.227
C8 H18 ( g ) 114.23 −208447 +16859 466.835
C8 H18 (l ) 114.23 −249952 +6940 360.896

Table 5.5 give the values of ∆ h = h − h298 kJ / kg mol for various substances at
0 0

different temperature.

0
Table 5.5 Enthalpy of formation at 25 C , ideal gas enthalpy, and Absolute
Entropy at 0.1 MPa (1bar) Pressure

Nitrogen, Diatomic ( N 2 ) Oxygen, Diatomic ( O2 )

(h )
0
f 298
= 0kJ / kmol (h )
0
f 298
= 0kJ / kmol
M = 28.013 M = 31.999

Temperature ( 0
h − h 298
0
) s
0
( 0 0
h − h 298 ) s
0

K kJ / kmol kJ / kmolK kJ / kmol kJ / kmolK

0 -8669 0 -8682 0
100 -5770 159.813 -5778 173.306
200 -2858 179.988 -2866 193.486

106
298 0 191.611 0 205.142
300 54 191.791 54 205.322
400 2971 200.180 3029 213.874
500 5912 206.740 6088 220.698
600 8891 212.175 9247 226.455
700 11937 216.866 12502 231.272
800 15046 221.016 15841 235.924
900 18221 224.757 19246 239.936
1000 21406 288.167 22707 243.585
1100 24757 231.309 26217 246.928
1200 28108 234.225 29765 250.016
1300 31501 236.941 33351 252.886
1400 34936 239.484 36966 255.564
1500 38405 241.878 40610 258.078
1600 41903 244.137 44279 260.446
1700 45430 246.275 47970 262.685
1800 48982 248.304 51689 264.810
1900 52551 250.237 55434 266.835
2000 56141 252.078 59199 268.764

Carbon Dioxide ( CO ) 2
Carbon Monoxide ( CO )
(h ) 0
f 298
= −393522kJ / kmol (h )
0
f 298
= −110529kJ / kmol
M = 44.01 M = 28.01

Temperature ( 0
h − h 298
0
) s
0
( 0
h − h 298
0
) s
0

K kJ / kmol kJ / kmolK kJ / kmol kJ / kmolK

0 -9364 0 -8669 0
100 -6456 179.109 -5770 165.850
200 -3414 199.975 -2858 186.025
298 0 213.795 0 197.653
300 67 214.025 54 197.833
400 4008 225.334 2975 206.234
500 8314 234.924 5929 212.828
600 12916 243.309 8941 218.313
700 17765 250.773 12021 223.062
800 22815 257.517 15175 227.271
900 28041 263.668 18397 231.006
1000 33405 269.325 21686 234.531
1100 38894 274.555 25033 237.719
1200 44484 279.417 28426 240.673
1300 50158 283.956 31865 243.426

107
1400 55907 288.216 35338 245.999
1500 61714 292.224 38848 248.421
1600 67580 296.010 42384 250.702
1700 73492 299.592 45940 252.861
1800 79442 302.993 49522 254.907
1900 85429 306.232 53124 256.852
2000 91450 309.320 56739 258.710

Water ( H O)
2
Hydrogen,Diatomic (H ) 2

(h )
0
f
298
= −241827 kJ / kmol (h )
0
f
298
= 0kJ / kmol
M = 18.015 M = 2.016

Temperature (h − h )
0 0
298 s
0
(h − h )
0 0
298 s
0

K kJ / kmol kJ / kmolK kJ / kmol kJ / kmolK

0 -9904 0 -8468 0
100 -6615 152.390 -5293 102.145
200 -3280 175.486 -2770 119.437
298 0 188.833 0 130.684
300 63 189.038 54 130.864
400 3452 198.783 2958 139.215
500 6920 206.523 5883 145.738
600 10498 213.037 8812 151.077
700 14184 218.719 11749 155.608
800 17991 223.803 14703 159.549
900 21924 228.430 17628 163.060
1000 25978 232.706 20686 166.223
1100 30167 236.694 23723 169.118
1200 34476 240.443 26794 171.792
1300 38903 243.986 29907 174.281
1400 43447 247.350 33062 176.620
1500 48095 250.560 36267 178.833
1600 52844 253.622 39522 180.929
1700 57685 256.559 42815 182.929
1800 62609 259.372 46150 184.833
1900 67613 262.078 49522 186.657
2000 72689 264.681 52932 188.406

5.12 FIRST LAW FOR REACTIVE SYSTEMS

For the S.S.S.F. process as shown in Fig.5.4 the first law gives

108
H R + QC .V . = H P + WC .V . (5.99)
or,
∑n h + Q
R
i i C .V .
= ∑ ne he + WC .V .
p
(5.100)

or,

∑ n ⎡⎣ h
R
i
0
f
+ ∆ h ⎤⎦ + QC .V . = ∑ ne ⎡⎣ h 0f + ∆ h ⎤⎦ + WC .V .
p e
(5.101)
i

C .V

n1 n3
Reactants

Products
n2 n4

Q c.v W c.v

Fig.5.4 First law for a reaction system

When the states of reactants and products are not in the standard reference
state (298 K , 1atm), then, as shown in Fig.5.5,

109
R

R0

h0 P
RP
H
P0

298 K T

Fig.5.5 Enthalpy of reactants and products varying with temperature

QC .V . = H P − H R = ( H P − H P ) + ( H P − H R ) + ( H R − H R )
0 0 0 0
(5.102)

= ∑ ne (h p − h p ) + ∆ h RP − ∑ ni (h R − h R ) (5.103)
0
0 0
P R

0
where ∆ h PR is the enthalpy of reaction at the standard temperature (298 K ) . The
variation of enthalpy with pressure is not significant.
QC .V . = U P − U R = (U P − U P ) + (U P − U R ) + (U R − U R )
0 0
(5.104)
0 0

= ∑ ne (u p − u p ) + ∆u − ∑ ni (u R − u R )
0
0 RP 0
p R

0
where ∆u RP is the internal energy of reaction at 298 K .

5.13 ADIABATIC FLAME TEMPERATURE

If a combustion process occurs adiabatically in the absence of work transfer or


changes in K.E. and P.E., then the energy equation becomes
H =H R P
(5.105)
or,
∑n h = ∑n h
R
i i
p
e e
(5.106)

or,
∑n h = ∑n h
R
i i
p
e e
(5.107)

or,

110
∑ n ⎡⎣ h + ∆ h ⎤ = ∑ ne ⎡ h f + ∆ h ⎤
0 0 (5.108)
⎦i P ⎣ ⎦e
i f
R

For such a process, the temperature of the products is called the adiabatic flame
temperature which is the maximum temperature achieved for the given reactants.
The adiabatic flame temperature can be controlled by the amount of excess air
supplied; it is the maximum with a stoichiometric mixture. Since the maximum
permissible temperature in a gas turbine is fixed from metallurgical
considerations, close control of the temperature of the products is achieved by
controlling the excess air. For a given reaction the adiabatic flame temperature is
computed by trial and error. The energy of the reactants H R being known, a
suitable temperature is chosen for the products so that the energy of products at
that temperature becomes equal to the energy of the reactants.

5.14 ENTHALPY AND INTERNAL ENERGY OF COMBUSTION: HEATING


VALUE

The enthalpy of combustion is defined as the difference between the enthalpy of


the product and the enthalpy of the reactants when complete combustion occurs
at a given temperature and pressure. Therefore,

h RP = H P − H R (5.109)

or,
h RP = ∑ ne ⎡ h f + ∆ h ⎤ − ∑ ni ⎡ h f + ∆ h ⎤
0 0 (5.110)
P ⎣ ⎦e R ⎣ ⎦i
where h RP is the enthalpy of combustion ( kJ / kg or kJ / kgmol ) of fuel. The
value of the enthalpy of combustion of different hydrocarbon fuels at
250 C , 1atm are given in Table 5.6.The internal energy of combustion, µ RP , is
defined similar way.

µ RP = U P − U R (5.111)

= ∑ ne ⎡ h f + ∆ h − pv ⎤ − ∑ ni ⎡ h f + ∆ h − pv ⎤
0 0 (5.112)
p ⎣ ⎦e R ⎣ ⎦i
If all the gaseous constituents are considered ideal gases and the volume of
liquid and solid considered is assumed to be negligible compared to gaseous
volume.
u RP = u RP − RT (n −n
gaseous products
) (5.113)
gaseous reac tan ts

111
In the case of a constant pressure or steady flow process, the negative of the
enthalpy of combustion is frequently called the heating value at constant
pressure, which represents the heat transferred from the chamber during
combustion at constant pressure. Similarly, the negative of the internal energy of
combustion is sometimes designated as the heating value at constant volume in
the case of combustion, because it represents the amount of heat transfer in the
constant volume process. The higher heating value (HHV) or higher calorific
value (HCV) is the heat transferred when H 2O in the products is in the liquid
state. The lower heating value (LHV) or lower calorific value (LCV) is the heat
transferred in the reaction when H 2O in the products is in the vapour state.
Therefore,
LHV = HHV − mH O .h fg 2
(5.114)
where mH O is the mass of water formed in the reaction.
2

5.15 ABSOLUTE ENTROPY AND THE THIRD LAW OF THERMODYNAMICS

So far only first law aspects of chemical reactions have been discussed. The
second law analysis of chemical reactions needs a base for the entropy of
various substances. The entropy of substances at the absolute zero of
temperature, called absolute entropy, is dealt with by the third law of
thermodynamics formulated in the early twentieth century primarily by Nernst
(1864-1941) and Max Planck (1858-1947). The third law states that the entropy
of a perfect crystal is zero at the absolute zero of temperature ant it represents
the maximum degree of order. A substance not having a perfect crystalline
structure and processing a degree of randomness such as a solid solution or a
glassy solid has a finite value of entropy at absolute zero. The third law provides
an absolute base from which the entropy of each substance can be measured.
The entropy relative to this base is referred to as the absolute entropy. Table 5.5
gives the absolute entropy of various substances at the standard state
250 C , 1atm. For any other state,
( ) (5.115)
0
sT , p = sT + ∆ s
T ,1atm ,→T , p
0
where s T refers to the absolute entropy at 1 atm, and temperature T, and
(s)T ,1atm ,→T , p
refers to the change of entropy for an isothermal change of pressure
from 1atm to pressure p (Fig.5.6).

112
A b s o lu te e n tro p y is

tm
k n o w a lo n g th is

1a
is o b a r, a s s u m in g id e a l g a s b e h a v io u r a t
1 a tm

T
( s ) T ,1 a tm T ,p

T .p

s
Fig. 5.6 Absolute entropy

0
Table 5.5 give the value of s for various substances at 1 atm. and at different
temperatures. Assuming ideal gas behaviour ( ∆s ) T ,1atm ,→T , p
can be determined
(Fig 5.6).
p2 (5.116)
s 2 − s1 = − R ln
p1
or,

( ∆s ) T ,1atm , →T , p
= − R ln p (5.117)

where p is in atm.

0
Table 5.6 Enthalpy of Combustion of Some Hydrocarbons at 25 C

Liquid H 2O in Vapour H 2O in
Pr oducts Pr oducts
( Negative of ( Negative of
Higher Heating Lower Heating
Value) Value
Liquid Liquid
Gaseous Gaseous
Hydro − Hydro −
Hydro − Hydro −
Hydrocarbon carbon carbon
Formula carbon carbon
kJ / kg kJ / kg
kJ / kg kJ / kg
fuel fuel
fuel fuel

113
(1) (2) (3) (4) (5) (6)
Paraffin Family
−55496
CH 4 −50010
Methane −51875
C2 H 6 −47484
Ethane −49975 −50345
C3 H 8 −45983 −46353
Propane −49130 −49500
C4 H10 −45344 −45714
Butane −48643 −49011
C5 H12 −44983 −45951
Pentane −48308 −48676
C6 H14 −44733 −45101
Hexane −48071 −48436
C7 H16 −44557 −44922
Hexane −47893 −48256
C8 H18 −44425 −44788
Octane −47641 −48000
C10 H 22 −44239 −44598
Decane −47470 −47828
C12 H 26 −44100 −44467
Olefin Family
Ethene C2 H 4 −50296 −47158
Propene C3 H 6 −48917 −45780
Butene C4 H 8 −48453 −45316
Pentene C5 H10 −48134 −44996
Hexene C6 H12 −47937 −44800
Heptene −47800 −44662
C7 H14
Octene −47693 −44556
C8 H16
Nonene −47612
C9 H18 −44475
Decene −47547
C16 H 20 −44410
Alkylbenzene Family
Benzene C6 H 6 −41831 −42266 −40141 −40576
Methylbenezene C7 H 8 −42473 −42847 −40527 −40937
Ethylbenzene C8 H10 −42997 −43395 −40924 −41322
Propylbenzene C9 H12 −43416 −43800 −41219 −41603
Butylbenzene C10 H14 −43748 −44123 −41453 −41828

114
5.16 SECOND LAW ANALYSIS OF REACTIVE SYSTEMS

The reversible work for a steady state flow process, in the absence of changes in
K.E. and P.E., is given by

Wrev = ∑ ni (hi − T0 si ) − ∑ ne (he − T0 se ) (5.118)

For an S.S.S.F. process involving a chemical reaction


Wrev = ∑ ni ⎡ h f + ∆ h − T0 s ⎤
0 (5.119)
R ⎣ ⎦i
−∑ ne ⎡ h f + ∆ h − T0 s ⎤
0

P ⎣ ⎦e

The irreversibility for such a process is


I = ∑ neT0 s e − ∑ niT0 s i − QC .V . (5.120)
P R
The availability,ψ , in the absence of K.E. and P.E. changes, for an S.S.S.F.
process is
ψ = (h − T s ) − (h − T s )
0 0 0 0
(5.121)
When an S.S.S.F. chemical reaction takes place in such a way that both the
reactants and products are in temperature equilibrium with the surroundings, the
reversible work is given by
W = ng − ng
rev ∑R
i i ∑ P
e e
(5.122)

where the g ′ s refer to the Gibbs function. The Gibbs function for formation, g f , is
0

0
defined similar to the enthalpy of formation, h f . The Gibbs function of each of the
0
elements at 25 C and 1atm. is assumed to be zero, and the Gibbs function of
0
each substance is found relative to this base. Table 5.4 gives g f for some
0
substance at 25 C , 1atm.

5.17 SECOND LAW EFFICIENCY OF A REACTIVE SYSTEM

For a fuel at T0, p0, the chemical exergy is the maximum theoretical work that
could be obtained through reaction with environmental substances. However,
due to various irreversibilities like friction and heat loss, the actual work obtained
is only a fraction of this maximum theoretical work. The second law efficiency
may thus be defined as the ratio of

115
Actual work done (5.143)
η II =
Maximum theoretical work

WcV (5.144)
η II =
m fuel xach
The associated irreversibility and the consequent exergy loses require to be
reduced to enhance the second law efficiency, which in turn, reduce the fuel
consumption and also increases the cost. The trade off between the fuel savings
and the additional costs must be carefully weighed.

5.18 CHEMICAL EXERGY

A system is in thermal and mechanical equilibrium with the environment when


the system attains the dead state. Under such conditions, exergy of the system is
zero. However, the contents of a system even at the dead state may undergo
chemical reaction with environmental components and produce additional work.

Consider a combined system formed by an environment and a system having a


certain amount of fuel at T0 , p0 . Work is obtainable by allowing the fuel to react
with oxygen from the environment to produce the components of CO2 and H 2O .
The chemical exergy is thus defined as the maximum theoretical work that could
be developed by the combined system. Thus for a given system at a specific
state

Total exergy = Thermomechanical exergy + Chemical exergy

Let us consider a hydrocarbon fuel (Ca H b ) at T0 , p0 reacting with oxygen from


the environment (Fig.5.7) which is assumed to be consisting of an ideal gas
mixture at T0, p0.

116
Work
p0, T0
Ca H b

CO 2 at T0
O2 at T0 XCO2, p0
XO2 p0
H2O at T0,XH20 p 0
T0

Heat transfer environment at p0 , T0

Boundary of the combined system

Fig. 5.7 Fuel exergy concept

The oxygen that reacts with the fuel is at a partial pressure of x02 p0, where x02 is
the mole fraction of oxygen in the environment. The fuel and oxygen react
completely to produce CO2 and H 2O, which exit in separate steams at τ o and
respective partial pressures of χ co ρo and χ H ρo . The reaction is given by
2
.
2O

⎛ b⎞ b (5.123)
Ca H b + ⎜ a + ⎟ O2 → aCO2 + H 2O
⎝ 4⎠ 2
At steady state, the energy balance gives
⋅ (5.124)
Q cV
 R = WcV + nH
+ nH  P
or,
WcV Q cV Q 0 ⎛ b⎞ b (5.125)
= + H R − H P = cV + (h f + ∆ h)Ca H b + ⎜ a + ⎟ h 0 − ahCO − h H O
n n ⎝ 4⎠
2 2 2
n 2
where n is the rate of fuel flow in moles, and K.E. and P.E. effects are neglected.
An entropy balance for the control volume gives

Q cV / n ⎛ b⎞ (5.126)
0= + sC H + ⎜ a + ⎟ so
⎝ 4⎠
a b 2
T0
b s
− as CO − s H O + gen
n
2 2
2
Eliminating Q cV between Eq. (5.125) and Eq. (5.126)

117
WcV ⎡ ⎛ b⎞ b ⎤ (5.127)
= ⎢ hC H + ⎜ a + ⎟ h o − ahCO − h H O ⎥
n ⎝ 4⎠
a b 2 2 2

⎣ 2 ⎦
⎡ ⎛ b⎞ b ⎤ T s
−T0 ⎢ s C H + ⎜ a + ⎟ s o − as CO − s H O ⎥ − 0 gen
⎝ 4⎠ n
a b 2 2 2

⎣ 2 ⎦

The specific enthalpies in Eq. (5.127) can be determined knowing only the
temperature T0 and the specific entropies can be determined knowing T0 , p0 and
the composition of the environment. For maximum work,
(5.128)
T s
I = 0 gen = 0
n

Therefore, the chemical exergy a ch can be expressed as

⎡ ⎛ b⎞ b ⎤ (5.129)
a ch = ⎢ hC H + ⎜ a + ⎟ hO − ahCO − h H O ⎥
⎝ 4⎠
a b 2 2 2

⎣ 2 ⎦
⎡ ⎛ b⎞ b ⎤
−T0 ⎢ s C H + ⎜ a + ⎟ s O − as CO − s H O ⎥
⎝ 4⎠
a b 2 2 2

⎣ 2 ⎦
The specific entropies for O2 , CO2 and H 2O are written from Eq. (5.104),
s i (T0, xi p0 ) = s i (T0, p0 ) − R ln xi (5.130)

where the first term on the right is the absolute entropy at T0 and p0, and xi is the
mole fraction of component in the environment .Therefore ,Eq.(5.129) becomes,
(5.131)
⎡ ⎛ b⎞ b ⎤
a ch = ⎢ hC H + ⎜ a + ⎟ h o − ahCO − h H O ⎥ (at T0, p0 )
⎝ 4⎠
a b 2 2 2

⎣ 2 ⎦
⎡ ⎛ b⎞ ⎤
−T0 ⎢ s C H + ⎜ a + ⎟ s O − as CO − s H O ⎥ ( at T0, p0 )
b
⎝ 4⎠
a b 2 2 2

⎣ 2 ⎦
(x )
a +b / 4

+ RT ln
O2

(x ) (x )
0 a b/2
CO2 H 2O

In terms of Gibbs function of respective substances,

⎡ ⎛ b⎞ b ⎤ (5.132)
a ch = ⎢ g C H + ⎜ a + ⎟ g O − ag CO − g H O ⎥ (at T0, p0 )
⎣ ⎝ 4⎠ 2 ⎦
a b 2 2 2

118
(x )
a +b / 4
O2
+ RT ln
(x ) (x )
0 a b/2
CO2 H 2O

where
g (T0, p0 ) = g f + ∆ g T (5.133)
0

0 , p0→Tref , pref

For the special case when T0 and p0 are the same as Tref and pref , ∆ g will be
zero. The chemical exergy of pure CO at T0 , p0 where the reaction is given by
1 (5.134)
CO + O2 → CO2
2
⎡ ⎤
(a )
(5.135)
= ⎢ g CO + g O − g CO ⎥ ( at T0, p0 )
1
ch
⎣ 2 ⎦
CO 2 2

(x )
1/ 2

+ RT ln
O2
0
xCO 2

Water is present as a vapour within the environment, but normally is a liquid at


T0, p0. The chemical exergy of liquid water is,

H 2O(l ) → H 2O( g ) (5.136)

(a ) ch
H 2O ( l )
= ⎡⎣ g H O ( l ) − g H O ( g ) ⎤⎦ ( atT0, p0 )
2 2
(5.137)

The specific exergy of a system is

a = athermo −mech + achem (5.138)


V2
= (u − u0 ) + p0 (v − v0 ) − T0 ( s − s0 ) + + gz + ach
2
and the specific flow exergy is given by
V2 (5.139)
a = (h − h0 ) − T0 ( s − s0 ) +
+ gz + ach
2
Between two states of system a constant composition, ach cancels, leaving just
the thermo-mechanical contribution. For example, to find the chemical exergy of
liquid octane, the reaction is
C H (l ) + 12.5O → 8CO + 9 H O( g )
8 18 2
(5.140)
2 2

Assume the composition of the environment (on molar basis):


N 2 75.67%, O2 20.35%, H 2O3.12%, CO2 0.03%, others 0.83%. Then,
a ch = ⎡⎣ g C H8 18 ( l )
+ 12.5 g O − 8 g CO − 9 g H O ⎤⎦ ( atT0, p0 )
2 2 2
(5.141)

119
(x )
12.5

+ RT ln
O2

(x ) (x )
0 8 9
CO2 H 2O

Using the values given in Table 5.4,


(a )C H
ch 8 18
(l ) = 6940 + 12.45(0) − 8(−394,374) − 9(−228,583) (5.142)

( 0.2035)
12.5

+8.3143(298.15)ln
( 0.003) ( 0.0312 )
8 9

= 5,408,068.72kJ / kgmol
= 47,346kJ / kg

120

S-ar putea să vă placă și