Vous êtes sur la page 1sur 13

It would be convenient to manipulate this expression into a more useful form.

Consider the
general reaction equation for the carbon monoxide and oxygen reaction (eqn (12.27)):
1
a
CO+ - 0 2 ::>(1 - a)C0 2 + aCO + - 0

then
nco, p
1-a p
Prco 2 = - - - =
np Po
1 + a/2 Po

a
p
nco p
Prco = - - - =
np Po
1 + a/2 Po
no, p
Pro, = - np Po

(12.71)

a/2

1 + a/2 Po

Hence, the value of KP< can be related to the degree of dissociation, a, through the
following equation
K

=
P,

1-a l+a/2
1 + a/2
a

JPo=

V---;;/2 JP

1-a
a

JPo

V---;;/2 JP

(12.72a)

and
K = 1-a

~-1-=

KP,

(12.72b)

V---;;/2 JP JPo

The relationship between KP< and a allows the degree of dissociation, a, to be evaluated
if KP< is known. Equations (12.72) shows that, for this reaction, the value of KP< is a
function of the degree of dissociation and also the pressure of the mixture. This is because
the amount of substance of products is not equal to the amount of substance of reactants.
EXAMPLE 2

12.7.2

Consider the water-gas reaction


(12.73)

C0 2 + H2 ::>CO + HzO
By the law of mass action
K

PrcoPrH 2 0

p,

Prco, PrH 2
=

(12.74)

Xco .xH,o) 1+ 1_ 1_1 (XcoXH,o)


(

Xco,xH ,

=KP

Xco 2 XH 2

In this reaction the amount of substance in the products is equal to the amount of
substance in the reactants, and there is no effect of pressure in the dissociation equation.
Comparing the results for the carbon monoxide, (eqn (12.1)) and the water gas reactions
(eqn (12.73)), it can be seen that if a mixture of products of the reaction in eqn (12.1) was

subjected to a change in pressure, the chemical composition of the mixture would change,
whereas the chemical composition of the products of the water gas reaction would be the
same at any pressure.
These points are returned to later in this section.
12.7.3

EXAMPLE I: A ONE DEGREE OF DISSOCIATION EXAMPLE

A spark ignition engine operates on a 10% rich mixture of carbon monoxide (CO) and air.
The conditions at the end of compression are 8.5 bar and 600 K, and it can be assumed
that the combustion is adiabatic and at constant volume. Calculate the maximum pressure
and temperature achieved if dissociation occurs.
[There are two different approaches for solving this type of problem; both of these will
be outlined below. The first approach, which develops the chemical equations from the
degrees of dissociation is often the easier method for hand calculations because it is
usually possible to estimate the degree of dissociation with reasonable accuracy, and it
can also be assumed that the degree of dissociation of the water vapour is less than
that of the carbon dioxide. The second approach is more appropriate for computer
programs because it enables a set of simultaneous (usually non-linear) equations to be
defined.]
General considerations
The products of combustion with dissociation have to obey all of the laws which define the
conditions of the products of combustion without dissociation, described in Chapter 10,
plus the ratios of constituents defined by the equilibrium constant, K"'. This means that the
problem becomes one with two iterative loops: it is necessary to evaluate the degree of
dissociation from the chemical equation and the equilibrium constant, and to then ensure
that this obeys the energy equation (i.e. the First Law of Thermodynamics). The following
examples show how this can be achieved. At this stage the value of Kp, at different
temperatures will simply be stated. The method of calculation K"' is described in section
12.8, and K"' values are listed in Table 12.3.
Solution
Note that there is only one reaction involved in this problem, and that the carbon monoxide
converts to carbon dioxide in a single step. Most combustion processes have more than one
chemical reaction.
Stoichiometric combustion equation:
1
CO+ - (0i + 3.76N 2 ) => C0 2 + 1.88N 2
2
np = 2.88

(12.75)

" = 3.38

Rich combustion equation:


1

l.lCO + - (0i + 3.76N 2 ) => C0 2 + O.lCO + 1.88N 2


2
np = 2.98
" =3.48

(12.76)

Dissociation of C02 (eqn (12.29)):

Total combustion equation with dissociation


1
a
l.ICO + - (0i + 3.76N 2 ) - (1 - a) C0 2 + (0.1 + a)CO + - 0 2 + 1.88 N 2
2
2
np 2.98 + a/2

(12.77)
The equilibrium constant, KPr, is given by eqn (12.70):
(p/po)co 2

K =-----Pr
(p/po)co (p/po)g;

The values of partial pressures in the products are


Pco,
P2
1 - a P2
- - =Xco, - = - - - ;
Po
Po
np Po

Pco
P2
0.1 - a Pz
-=Xco-=
Po
Po
np
Po

Po,
P2
a P2
=Xo2-=-Po
Po
2np Po

(12.78)

Also, from the ideal gas law,


Pz V2 = nz~Tz, and P1 V1 = n1~T1

(12.79)

Thus
np

n1 T1

245.65

(12.80)

Hence, substituting these values in eqn (12.70) gives


K =
Pr

(1 - a) ( 2np p 0 )
-(0.1 +a) a Pz

112

~K

2
Pr

(1 - a) ( 2 x 245.65 )
p
(0.1 + a)2
aT2

(12.81)

This is an implicit equation in the product's temperature, T2 , because KPr = /(T2 ). Writing
as X gives

K;r

X(O.l + a) 2 aT2 = (1- a) 2 x 2 x 245.65p 0 = (1- a) 2 x 2 x 245.65 x 1.013 25


(12.82)
Expanding eqn (12.82) gives
0 = 497.77 - a(995.54 + O.OIXT2 ) + a 2 {497.77 - 0.2XT2 ) - a 3XT2

(12.83)

The term XT2 can be evaluated for various temperatures because XT2 = Ki,T2 :
T2 (K)

KP.

XT2

2800
2900
3000

6.582
4.392
3.013

121 303
55 940
27 234

Solving for a at each temperature gives

2800
2900
3000

0.090625
0.1218125
0.171875

These values of a all obey the chemical, i.e. dissociation, equation but they do not all
obey the energy equation. It is necessary to consider the energy equation now to check
which value of T2 balances an equation of the form
Up(Tp) = -(Qv)s + [UR(TR) - UR(T.)] + Up(T.)

This energy equation, based on the internal energy of reaction at T

0, may be rewritten

0 = -!1U0 - Ur(Tp) + UR(TR)

(12.84)

Now UR(TR) is constant and is given by


Constituent

co

u(J()()

12 626.2

1.1

12 939.6
0.5

12 571.7
1.88

Hence,
Ur(Tr) = (1 - a) x 276 960 + 43 993 kJ

(12.85)

It can be seen that the value of Ur is a function of a; the reason for this is because the
combustion of the fuel (CO) is not complete when dissociation occurs. In a simple, single
degree of freedom reaction like this, the reduction in energy released is directly related to
the progress of the reaction:

a
2800 0.090625
2900 0.1218125
3000 0.171875

295 853
287 216
273 351

Evaluating the energy which is contained in the products at 2900 K and 3000 K,
allowing for the variation in a as the temperature changes, gives

Tr= 2900 K
Constituent

131 933.3
0.871875

U2900
n

co

02

N2

74 388.9
0.2281

78 706.6
0.0641

73 597.5
1.88

C0 2

Up(Tp) = 275 405 kJ

Tr= 3000 K
Constituent

co

C0 2

137 320 77 277


0.8281
0.2719

U31XXJ

02

N2

81 863
76 468
0.085 94 1.88

Up(Tp) = 285 521 kJ

These values are plotted in Fig 12.3, and it can be seen that, if the variation of the
energy terms was linear with temperature, the temperature of the products after
dissociation would be 2949 K. The calculation will be repeated to show how well this
result satisfies both the energy and dissociation equations.
288000

286000

284000

282000

""'2"'

Energy released from


dissociation equation

280000

Q.

.5

w"
c

Energy absorbed
by gases

278000

276000

274000

272000
2900

2910

2920

2930

2940

2950

2960

2970

2980

2990

3000

Temperature I (K)

Fig. 12.3 Energy contained in products based on eqn (12.80) and the tables of energies

First, it is necessary to evaluate the degree of dissociation that will occur at this product
temperature. At Tr= 2950 K, KP<= 3.62613 (see Table 12.3), and this can be substituted
into eqn (12.78) to give a= 0.1494.
Hence, the chemical equation, taking account of dissociation, is
1
1.lCO + - (0 2 + 3.76 N 2 ) => 0.8506 C0 2 + 0.2494 CO+ 0.0747 0 2 + 1.88 N 2
2

Applying the energy equation: the energy released by the combustion process gives a

product energy of
Ur(Tr) = (1 - a) x 276 960 + 43 993 = 279 575 kJ

This energy is contained in the products as shown in the following table:

co

Constituent
U2949

135 182 75 820


0.8506
0.2494

80 205
0.0747

75 040
1.88

The equations are balanced to within 0.5%, and this is close enough for this example.
Alternative method
In this approach the degree of dissociation, a, will not be introduced explicitly. The
chemical equation which was written in terms of a in eqn (12.77) can be written as
(12.86)
np=a + b+ c + d

" = 3.48

Considering the atomic balances,


Carbon:

l.l=a+b

giving b = 1.1 - a

(12.87a)

Oxygen:

2.1 =2a+b+2c
1.88 = d

giving c = 0.5 (1 - a)

(12.87b)

Nitrogen:

(12.87c)

Total amount of substance in the products,


nr= a+ b + c + d

(12.88)

The ratios of the amounts of substance in the equilibrium products are defined by the
equilibrium constant
(p/po)co,
KP,= - - - - - - -2
(p/po)co (p/po)Y,

1/2

1/2

anr Po

b c 1/2Pr1/2

(12.89)

Hence
Pr
b2c Po nr

K2

P,=a

(12.90)

and, from the atomic balances, this can be written in terms of a as


2 1
Pr
2
2
(1.1 - a) - (1 - a) - - KP =a
2
Po nr
'

(12.91)

The previous calculations showed that the temperature of the products which satisfies the
governing equations is Tr= 2950 K, which gives a value of K~, = 13.1488. These values

will be used to demonstrate this example. Then


2
p
13.1488
{1.21--2.2a+a 2 -(l.21a+2.2a 2 +a 3 )} _r_ x
=a
2
Ponr

(12.92)

It is possible to evaluate the ratio Pr/ p 0 nr from the perfect gas relationship, giving
8.5 x 2950
---=12.009
3.48 x 600
This enables a cubic equation in a to be obtained:
95.532 -- 269.23a + 251.65a 2 - 78.953a 3 = 0

(12.93)

The solution to this equation is a = 0.8506, and hence the chemical equation becomes
1

l.lCO + - (0 2 + 3.76N 2 )--0.8506C0 2 + 0.2494CO + 0.07470 2 + l.88N 2


2
(12.94)
which is the same as that obtained previously. The advantage of this approach is that it is
possible to derive a set of simultaneous equations which define the equilibrium state, and
these can be easily solved by a computer program. The disadvantage is that it is not possible
to use the intuition that most engineers can adopt to simplify the solution technique. It must
be recognised that the full range of iteration was not used in this demonstration, and the
solutions obtained from the original method were simply used for the first 'iteration'.

12.8 The Van't Hoff relationship between equilibrium constant and


heat of reaction
It has been shown that (eqn (12.59))
(12.95)
Thus

(12.96)
Consider the definition of 0 ,
0

= h 0 +h(T)-T{s 0 +s(T))
=h 0 +h(T)-Ts 0 -T

Consider the terms


h(T)-T

I dh

dh
T
I

(12.97)

I--y- I T I

y:=T h(T)

dh

(12.98)

Let
v=h(T),

dv=dg
dT
du= - -2

1
u=-,
T

Integrating by parts gives


T!

h~) - I dh~T) I= -TI

h(T)dT
T2

(12.99)

Thus
~ lnK
dT
p,

= _ 'V _2:'.._

L., 9t

h(T)
T2

= _ 'V

vh(T)

L., 9tT 2

(12.100)

Although L has been used as a shorthand form it does include both +ve and -ve signs;
these must be taken into account when evaluating the significance of the term.
Thus

(12.101)
But, by definition

Qp= (vjzc+ vdhd)- (vaha+ vbhb)

(12.102)

~(lnK )=~2

(12.103)

hence
dT

p,

9tT

Equation (12.103) is known as the Van't Hoff equation. It is useful for evaluating the
heat of reaction for any particular reaction because

Q
P

=-

9td(lnKP)
d(l /T)

The value of d(ln KP)/d(l/T) may be obtained by plotting a graph of ln KP against 1/T.
The values of KP have been calculated using eqn (12.95), and are listed in Table 12.3, at
the end of the chapter, for four reactions. (The values have also been depicted as a graph in
Fig 12.6, and it can be seen that over a small range of temperature, ln KP,., A - b/T, where
Tis the temperature in K: this is to be expected from eqn (12.95)).

12.9
12.9.1

The effect of pressure and temperature on degree of dissociation


The effect of pressure

The effect of pressure on the degree of dissociation is defined by eqn (12.62), namely,

It can be seen that the ratio of the amounts of substance is given by


(vc +vJ-va-vh)

p,

(12.104)

This can be interpreted in the following way. If v c + v d - v a - vb = 0 then the mole fractions of the products will not be a function of pressure. However, if v c + v d - v a - vb> 0
then the species on the left-hand side of the chemical equation (i.e. the reactants) increase,
whereas if v, + v d - v 0 - vb< 0 then the species on the right-hand side of the equation (i.e.
the products) increase. The basic rule is that the effect of increasing the pressure is to shift
the equilibrium to reduce the total amount of substance.
Considering the two reactions introduced previously. The equation for the carbon
monoxide reaction (12.27) was
1
a
CO+ - C)z <=::> (1 - a)C0 2 + aCO + - 0 2

and the equilibrium equation (12.72) was


=

K
P,

1-a l+a/2
1 + a/ 2
a

~-1-=

~ ----;;/2

JPr

1-a
a

~-1-

~ ----;;/2

JPr

This means that v c + v d - v a - vb< 0, and this would result in the constituents on the
'products' side of the equation increasing. This is in agreement with the previous statement
because the total amounts of reactants in eqn (12.27) is 1.5, whilst the total amounts of
products is 1 + a/2, where a is less than 1.0.
A similar calculation for the combustion and dissociation of a stoichiometric methane
(CH.i) and air mixture, performed using a computer program entitled EQUIL2 gave the
results in Table 12.1. The chemical equation for this reaction is
CH.i + 2(0 2 + 3.76N2 )
:=:::} (1 - a 1 )C0 2 + a 1CO + 2(1 - a 2 )Hi0 + 2a 2 H2 + (ai/2 + a 2 )0 2 + 7.52N 2
(12.105)
where a 1 is the degree of dissociation of the C0 2 reaction and a 2 is the degree of
dissociation of the H20 reaction. It can be seen that dissociation tends to increase the
Table 12.1 Amount of products for constant pressure combustion of methane
in air. Initial temperature = 1000 K; equivalence ratio = 1.00

Pressure (bar)

No
dissociation

Amount of C0 2
Amount of CO
Amount of H20
Amount ofH2
Amount of0 2
Amount of N 2
Total amount of substance

1
0
2
0
0
7.52
10.52

10
0.6823
0.3170
1.8654
0.1343
0.2258
7.52
10.7448

0.7829
0.2170
1.920
0.0801
0.1485
7.52
10.6685

100
0.8665
0.1335
1.9558
0.0444
0.0889
7.52
10.6091

amount of substance of products, and hence the effect of an increase in pressure should be
to reduce the degree of dissociation. This effect can be seen quite clearly in Table 12.1,
where the amount of substance of products is compared under four sets of conditions: no
dissociation, dissociation at p = 1 bar, dissociation at p = 10 bar, and dissociation at
p = 100 bar. The minimum total amount of substance occurs when there is no dissociation,
while the maximum amount of substance occurs at the lowest pressure. Figure 12.4 shows
how the degrees of dissociation for the carbon dioxide and water reactions vary with
pressure.
0.35

O.J

---o-- Carhon d10\1de


- - a - - Vv'ater

=
025
0

:;:::
~

c;
0

"1

0.2

:.;;"'

....
0

<IJ

0.15

<IJ
....

<IJ

01

0 05

10

100

Pressure I (bar)

Fig. 12.4 Effect of pressure on degree of dissociation

12.9.2

THE EFFECT OF TEMPERATURE

The effect of temperature can be considered in a similar way to the effect of pressure.
Basically it should be remembered that the changes in composition that take place during
dissociation do so to achieve the minimum value of Gibbs energy for the mixture. The Gibbs
energy of each constituent is made up of three components, the Gibbs energy at absolute zero
(g 0 ), the Gibbs energy as a function of temperature (g(T)), and that related to partial
pressure (see Fig 12.2). The equilibrium point is achieved when the sum of these values for
all the constituents is a minimum. This means that, as the temperature rises, the constituents
with the most positive heats of formation are favoured. These constituents include 0 2
(g 0 = 0), H2 (g 0 = 0), and CO (g 0 = -113 Ml /kmol). Both water and carbon dioxide have
larger (negative) values of g 0 This effect can be seen from the results in Table 12.2, which
have been calculated for the combustion of methane in air. Another way of considering this
effect is simply to study eqn (12.55), and to realise that for gases with negative heats of
formation an increase in temperature leads to a decrease in the value of KP. This means that
the numerator of eqn (12.55) must get smaller relative to the denominator, which pushes the
reaction backwards towards the reactants. This is borne out in Table 12.2: the degree of
dissociation for this reaction increases with temperature, and this is shown in Fig 12.5.

Table 12.2 Amount of products for constant pressure combustion of methane in air.
Pressure = 1 bar; equivalence ratio = 1

Temperature (K)

No
dissociation

T=lOOOK

T=l500K

T=2000 K

Amount of C0 2
Amount of CO
Amount of HP
AmountofH2
Amountof0 2
AmountofN 2
Total amount of substance

1
0
2
0
0
7.52
10.52

0.6823
0.3170
1.8654
0.1343
0.2258
7.52
10.7448

0.4814
0.5186
1.7320
0.2679
0.3933
7.52
10.9132

0.3031
0.6969
1.5258
0.4741
0.5855
7.52
11.1055

07

0.6

c: 0.5

.::
";

o;
0

"'

04

---o-- Carbon dioxide

"'
:0

....

--<r-- Water

~ 0.3
~

OJ)
Q;

0.2

01

1000

1100

1200

1300

1400

1500

1600

1700

1800

1900

2000

Temperature I (K)

Fig. 12.5 Effect of temperature on degree of dissociation

Finally, it should be noted that in all the cases shown the degree of dissociation in the
hydrogen reaction is much less than that for the carbon reaction. This supports the assumption
made in previous work that the hydrogen will be favoured in the oxidation process.

12.10

Dissociation calculations for the evaluation of nitric oxide

If it is necessary to evaluate the formation of nitric oxide (NO) in a combustion chamber then

the equations have to be extended to include many more species. While it is possible to add
the calculation of NO to a simple dissociation problem, as is done in Example 5 below, this
does not result in an accurate estimate of the quantity of NO formed. The reason for this is
that NO is formed by a chain of reactions that are more complex than simply
N 2 +0 2 ~2NO

This chain of reactions has to include the formation of atomic oxygen and nitrogen, and

also the OH radical. To obtain an accurate prediction of the NO concentration a total of 12


species has be considered:
[1] H 20,

[2] H2 ,

[3] OH,

[4] H,

[5] N 2 ,

[6] NO,

[7] N,

[8] C02 ,

[9] CO,

[10] 0 2 ,

[11] 0,

[12] Ar.

The numbers in [ ] will be used to identify the species in later equations.


The formation and breakdown of these species are defined by the following set of
equations:

where b =

x1/x 2

Fig. 12.6 Variation of equilibrium constant, KP<, at a standard pressure, p 0 = 1 bar, with temperature
for the reactions CO+~ 0 2 ~ C0 2 , H2 + ~ 0 2 ~ H20, C0 2 + H2 ~CO+ H20, i N 2 + ~ 0 2 ~NO.

A numerical method for solving these equations is given in Horlock and Winterbone
(1986), based on the original paper by Lavoie et al. (1970).
Figure 12.7 shows the results of performing such calculations using a simple computer
program. The coefficients in the program were evaluated using the data presented in
Table 9.3, but with the addition of data for OH and N. The three diagrams are based on
combustion of octane in air at a constant pressure of 30 bar, and show the effect of
varying equivalence ratio, , at different temperatures. The conditions are the same as
quoted in Heywood (1988). Figure 12.7 (a) shows the results for the lowest temperature
of 1750 K, and it can be seen that the graph contains no atomic nitrogen (N) or atomic
oxygen (0) because these are at very low concentrations (<10- 4 ). When the mixture is
weak some OH is produced. It is apparent from this diagram that there is not much
dissociation of the carbon dioxide or water because the concentration of oxygen drops to
very low values above = 1. Some nitric oxide is formed in the weak region
(xNo = 3 x 10- 3 at the weakest mixture), but this rapidly reduces to a very low value at
stoichiometric simply because there is no oxygen available to combine with the nitrogen.
Figure 12. 7 (b) shows similar graphs for a temperature of 2250 K, and it can be seen that
the NO level has increased by almost a factor of 10. There is also some NO formed by
combustion with rich mixtures; this is because, by this temperature, the carbon dioxide
and water are dissociating, as is indicated by the increase in the CO and the OH radical in
the weak mixture region. By the time 2750 K is reached, shown in Fig 12.7(c), there is a
further tripling of the NO production, and CO is prevalent throughout the weak mixture
zone. There are also significant amounts of OH and oxygen over the whole range of
equivalence ratio.
The diagrams shown in Fig 12.7 are relatively contrived because they do not depict real
combustion situations. However, they do allow the parameters which control the
production of the products of combustion to be decoupled, to show the effects of changing
the parameters independently of each other. The equilibrium concentrations depicted in
Fig 12.7 are the values which drive the formation of the exhaust constituents through the
chemical kinetics equations, which will be discussed in Chapter 14.

12.11

Dissociation problems with two, or more, degrees of dissociation

The previous example, which considered the dissociation of carbon monoxide, shows the
fundamental techniques involved in calculating dissociation but is an unrealistic example
because rarely are single component fuels burned. Even if a single component fuel such as
hydrogen was burned in an engine, it would be necessary to consider other 'dissociation'
reactions because it is likely that nitric oxide (NO) will be formed from the combination
of the oxygen and nitrogen in the combustion chamber. These more complex examples
will be considered here.

12.11.1

EXAMPLE 4: COMBUSTION OF A TYPICAL HYDROCARBON FUEL

A weak mixture of octane (C 8H18 ) and air, with an equivalence ratio of 0.9, is ignited at
10 bar and 500 K and bums at constant volume. Assuming the combustion is adiabatic,
calculate the conditions at the end of combustion allowing for dissociation of the carbon
dioxide and water, but neglecting any formation of NO.

Vous aimerez peut-être aussi