Sie sind auf Seite 1von 15

t?

!f * 0, then u 0 .mand ho,m are different. Most calculations involve changes in enthalpy
or internal energy, and if the composition during a process is invariant the values of ho,m or
u 0 .m will cancel.
However, if the composition changes during a process it is necessary to know the
difference between the values of u 0 ,m for the different species at the reference temperature.
This is discussed below.
Obviously u 0 ,m and ho.m are consequences of the ideal gas assumption and the equation
hm = hm(T) + ho,m

contains the assumption that the ideal gas law applies down to T0 If T0 = 0, or a
outside the superheat region for the gas being considered, then the gas ceases to be
often becoming either liquid or solid. To allow for this it is necessary to include
heats. This will not be dealt with here, but the published data do include
corrections.
9.2.3

value
ideal,
latent
these

ENTROPY OF AN IDEAL GAS - THIRD LAW OF THERMODYNAMICS

The change in entropy during a process is defined as


ds = d~ - 9tdp
m
T
p

(9.15)

If the functional relationship hm = hm(T) is known then eqn (9.15) may be evaluated,
giving

sm -s0,m =JT
Tc
o

d~
T

-9tin...
Po

(9.16)

where So.m is the value of Sm at T0 and p 0


It is convenient to take the reference temperature T0 as absolute zero. It was previously
shown that
and hence

d~ = cp.m(T) dT

(9.17)

Consider

If cp.m(T0 ) ~ 0 then the expression is either infinite or indeterminate. However, before


reaching absolute zero the substance will cease to be an ideal gas and will become a solid.
It can be shown by the Debye theory and experiment that the specific heat of a solid is
given by the law cP = aT 3 Hence
d~

lim -

T -> 0 (

aT dT
T

= lim ( - 3-) = lim


T->0

(aT

dT)~O

(9.18)

T-> 0

To integrate from absolute zero to T it is necessary to include the latent heats, but still it is
possible to evaluate f dhm / T.

If sm(T) is defined as
Srr,(T) =

dh,.,

JTo -Td T

(9.19)

Sm= Sm(T) - 9t In !!_ + So.m


Po

(9.20)

then entropy

The term So,m is the constant of integration and this can be compared with u 0.m and ho,m
If the composition is invariant then the value of so,m will cancel out when evaluating
changes in s and it is not necessary to know its value. If the composition varies then it is
necessary to know, at least, the difference between s 0 values for the substances involved. It
is not possible to obtain any information about s0 from classical (macroscopic)
thermodynamics but statistical thermodynamics shows that 'for an isothermal process
involving only phases in internal equilibrium the change in entropy approaches zero at
absolute zero'. This means that for substances that exist in crystalline or liquid form at low
temperatures it is possible to evaluate entropy changes by assuming that differences in so.m
are zero.
It should also be noted that the pressure term, p, in eqn (9.20) is the partial pressure of
the gas ifit is contained in a mixture, and Po is a datum pressure (often chosen as 1 bar or,
in the past, 1 atmosphere).
9.2.4

THE GIBBS ENERGY (FUNCTION) OF AN IDEAL GAS

The Gibbs energy will now be derived for use later when considering the equilibrium
composition of mixtures (dissociation). By definition,
(9.21)
and
hm = hm(T) + ho,m

(9.22)

If the pressure ratio, pf p 0 , is denoted by the symbol p" i.e. Pr= pf p 0 , then
Sm= Sm(T) -

9t In Pr+ So,m

(9.23)

and
g = h(T) + h 0 - T(s(T)- 9t In Pr+ s 0)
= (h 0 - Ts 0 )

+ (h(T)- Ts(T)) + 9tT In Pr

(9.24)

If the terms at T0 are combined to give


go,m = ho,m - Tso,m

(9.25)

and the temperature-dependent terms are combined to give


(9.26)
then
(9.27)

Often gm(T) and g 0 ,m are combined to give g~ which is the pressure-independent portion
of the Gibbs energy, i.e. g~ = gm(T) + g 0 ,m, and then
(9.28)
g~ is the value of the molar Gibbs energy at a temperature T and a pressure of p 0 , and it is
a function of temperature alone. The datum pressure, p 0 , is usually chosen as 1 bar
nowadays (although much data are published based on a datum pressure of 1 atm; the
difference is not usually significant in engineering problems, but it is possible to convert
some of the data, e.g. equilibrium constants).
Gibbs energy presents the same difficulty when dealing with mixtures of varying
composition as u 0 , h 0 and s0 If the composition is invariant, changes in Gibbs energy are
easily calculated because the g 0 terms cancel. For mixtures of varying composition g 0 must
be known. It can be seen from eqn (9.26) that if the reference temperature, T0 , equals zero
then

(9.29)

9.3

Tables of u( T) and h( T) against T

These tables are based on polynomial equations defining the enthalpy of the gas. The
number of terms can vary depending on the required accuracy and the temperature range to
be covered. This section will limit the number of coefficients to six, based on Benson
(1977). The equation used is of the form
~(T)
hm - ho,m
2
3
4
-- =
=a1 +aiT+~T +a4 T +asT

9\T

(9.30)

9\T

The values of the coefficients for various gases are listed in Table 9.3, for the range 500 to
3000 K. If used outside these ranges the accuracy of the calculation will diminish.
Table 9.3 Enthalpy coefficients for selected gases found in combustion processes (based on kJ /
kmol, with the temperature in K)

Substance

a,

a.

a,

a2

a,

a6

ho= ao
kJ/kmol

H,

0.0000
0.0000
0.0000
0.0000
0.0000
0.0000
0.0000
-8.586lle-15
0.0000

- l.44392e-11
-2.19450e-12
-6.57470e-12
-4.90360e-12
8.66002e-l 1
l.53897e-ll
-l.81802e-ll
l.62497e-10
- l.38670e-l l

9.66990e-8
-3.22080e-8
l.95300e-9
-9.58800e-9
-7.88542e-7
- l.49524e-7
4.95240e-8
- l.24402e-6
l.00187e-7

-8.18100e-6
3.76970e-4
2.94260e-4
2.99380e-4
2.73114e-3
6.52350e-4
5.65590e-4
4.96462e-3
-2.51427e-4

3.43328 -3.84470
4.63284
3.317000
3.34435
3.75863
5.11346
3.50174
3.09590
6.58393
5.71243
3.25304
9.65140e-l
3.74292
8.15300
1.93529
3.73309
2.76403

0.0000
-l.13882e5
0.0000
8.99147e4
-3.93405e5
0.0000
-2.39082e5
-6.6930e4
2.46923e5

co
N,
NO

co,
o,
H 20

CH.,
0

Hence the enthalpy, internal energy, entropy and Gibbs energy can be evaluated as
follows.

Enthalpy
hm(T) = 9tT(a 1 + a 2 T + a 3 T 2 + a 4 T 3 + a 5 T 4 )
= 9t(a 1 T

+ a 2 T 2 + a 3T 3 + a 4 T 4 + a 5 T 5 )

(9.31)

=9tLa;T;
i= I

Internal energy
um(T) = 9tT((a 1 - 1) + a 2 T + a 3 T 2 + a 4 T 3 + a 5 T 4 )
=

9t((a 1 - l)T + a 2 T 2 + a 3 T 3 + a 4 T 4 + a 5 T 5 )
5

= 9t L

a;T; - 9tT = h(T) - 9tT

(9.32)

i= I

Entropy

This is defined by eqn (9.15) as


ds

d~ -9t dp
T

Integrating eqn (9 .15) gives


Sm(T)

=Sm - So,m =

J~ d~ - 9t ln(:)

Equation (9.20) provides some problems in solution. The first is that


r d~ =Jr cp,mdT
T
70
T

Jr,,

and this results in ln(O) when T = T0 Fortunately these problems can be overcome by use
of the Van't Hoff equation, which will be derived in Chapter 12, where dissociation is
introduced. This states that

0 )
h = -T 2( - d (':JT)
m
dT

(9.33)

where, for T0 = 0 K
~=chemical potential= gm(T) + g 0 .m = gm(T) + ho,m

(9.34)

and
hm = hm(T) + ho,m

(Note: the term ~. is similar to g~, which was introduced in eqn (9.28). At this stage it is
sufficient to note that the chemical potential has the same numerical value as the specific
Gibbs energy. The chemical potential will be defined in Section 12.2.)

By definition, in eqn (9.31),

hu,(T)

hu, - ho.m

9tT

9tT

-- =

= al + ai.T + ll]T

+ ~T + asT

(9.35)

giving
hu,(T)

a1

- - =-+ai+llJT+~T +asT
9tT
T

(9.36)

Substituting eqns (9.34), (9.13) and (9.36) into eqn (9.33) and rearranging gives
(9.37)
which can be integrated to
2

a T + _4_
a T + _s_
aT - ~
z.
z.
- a ln T + a2T + _3_
+A = g,m(T) + "0,m
I
2
3
4
9tT
9tT

(9.38)

It is conventional to let
(9.39)

(9.40)

The value of entropy can then be obtained from

s. (T) = hu,(T) - 8m(T)


m

(9.41)

(9.42)

Gibbs energy

The Gibbs energy can be evaluated from eqn (9.40) or eqn (9.26):
8m(T) = hm(T) - Tsm(T)

The values obtained from this approach, using eqns (9.31) to (9.42) are given in Tables
9.4 for oxygen, nitrogen, hydrogen, water, carbon monoxide, carbon dioxide, nitric oxide
and methane. The tables have been evaluated up to 3500 k, slightly beyond the range
stated in relation to Table 9.3, but the error is not significant. This enables most
combustion problems to be solved using this data. Other commonly used tables are those
of JANAF (1971).

9.3.1

TABLES OF MEAN SPECIFIC HEAT

Sometimes data are given in terms of mean specific heat rather than the actual specific heat
at a particular temperature. This approach will now be described.
Consider the change of enthalpy, h, between temperatures T 1 and T2 This is
h2 - h, =

1i

7j

(9.43)

Cp(T) dT = h(1;) - h(T..)

Now, this can be written


(9.44)

where cP is the mean specific heat between T1 and T2 Normally cP is given at a particular
temperature T, and it is then defined as the mean between the temperature T and a
reference temperature Trer i.e.
_
h(T) - h(Trer)
(cph=----T-Trer

(9.45)

It is also possible to write the mean specific heat as a polynomial function of temperature,
in which case
(9.46)
where a, band care tabulated coefficients.
If it is required to calculate the enthalpy difference between two temperatures T1 and T2 ,
then
(9.47)
Having calculated the change in enthalpy, h 2 - h 1 , then the change in internal energy,
u2 - u 1 , may be evaluated as
(9.48a)

>.-::. .E.~tha_l_PY at T,
.. .. ... .
T0

T,

Temperature, T

Fig. 9.1 Mean specific heat at constant pressure

or, in molar terms, by


um,2 - um.1 = hm.2 - ~T2 - (hm.1 - ~Ti)= hm,2 - hm.1 - ~(T2 - Ti)

(9.48b)

If the change of internal energy, u 2 - u 1 , is known, then the change of enthalpy, h 2 - h 1 ,


may be calculated in a similar way. The use of mean specific heat values enables more
accurate evaluation of temperature changes because the mean specific heats enable the
energy equation to be solved with an allowance for the variation of specific heats. The
mean specific heats approach is depicted in Fig 9.1.

9.4

Mixtures of ideal gases

Many problems encountered in engineering involve mixtures of gases - air itself is a


mixture of many gases although it can be considered to be oxygen and atmospheric
nitrogen (which can be assumed to contain the nitrogen in the air and the other 'inert'
gases such as argon and carbon dioxide). If the gases in a mixture are at a temperature well
above their critical temperature and a pressure below the critical pressure, they act as ideal
gases. Expressions relating the properties of mixtures of ideal gases will now be
formulated. These expressions are a direct consequence of the Gibbs- Dalton laws.
9.4.1

DALTON PRINCIPLE

Dalton stated that


any gas is as a vacuum to any gas mixed with it.
This statement was expanded and clarified by Gibbs to give the Gibbs-Dalton laws.
9.4.2
(a)

GIBBS-DALTON LAW

A gas mixture as a whole obeys the equation of state (eqn (9.2))


pV=n~T

(b)

(c)

9.4.3

where n is the total amount of substance in the mixture.


The total pressure exerted by a mixture is the sum of the pressures exerted by the
individual components as each occupies the whole volume of the mixture at the same
temperature.
The internal energy, enthalpy and entropy of the mixture are, respectively, equal to
the sums of the internal energy, enthalpy and entropy of the various components as
each occupies the whole volume at the temperature of the mixture.
MIXTURE RELATIONSHIPS

The total mass, m, of a mixture can be related to the mass of constituents by


n

m=m0 +mb+mc+ =

I:lnj
; ~

From the ideal gas law


m

pV= mRT= mwRT= n~T


mw

(9.49)

and hence for the individual constituents


pVa = maRaT = n/Jl.T

(9.50)

pVb = m~bT = nb9\T


pVc = mficT = n/J\.T, etc

In eqn (9.50), Va= volume of constituent a at the pressure and temperature of the mixture,
and Vb and Ve are similar volumes for constituents b and c. But the total volume, V, is

given by
n

V=V+V.b+v+
.. = '\"'v
a
c
L...ii

(9.51)

i= 1

and therefore
n

n=na+nb+nc+ = Ini

(9.52)

i= 1

Let

= xa = mole fraction of a in the mixture

(9.53)

Similarly, nb/ n = xb, and nJ n = x 0 etc, with nJ n = xi. Therefore


n

Xa+Xb+Xc+=IA;=l

(9.54)

i= I

It is possible to develop the term for the partial pressure of each constituent from
statement (b) of the Gibbs-Dalton laws. Then, for constituent a occupying the total
volume of the mixture at the pressure and temperature of the mixture,
PaV= n/Jl.T

But, for the mixture as a whole (eqn (9.2))


pV=n9\T

Hence, by dividing eqn (9.50) by eqn (9.2)


(9.55)
giving
Pa= xap, and in general Pi= X;P

(9.56)

and, also
n

LP;= Ix;p=p Ix;=p


i= 1

i= I

(9.57)

i= I

Often mixtures are analysed on a volumetric basis, and from the volumetric results it is
possible to obtain the partial pressures and mole fractions of the components. Normally

volumetric analyses are performed at constant pressure and temperature. Then, considering
the ith component,

n/Jl.T

V;=--

(9.58)

and the total volume of the mixture, V, is


n~T

V=-p

(9.59)

Hence, from eqns (9.58) and (9.59)

V;

'1j

= -

(9.60)

=X;

If the gas composition is given in volume percentage of the mixture, the mole fraction is
V;(%)

(9.61)

X;=lQO

and the partial pressure is


V;(%)
p;=--p
100

(9.62)

The terms relating to the energy of a mixture can be evaluated from statement (c) of the
Gibbs-Dalton laws. If em= molar internal energy, and hm =molar enthalpy, then for the
mixture
n

=nem =I n;em,i

(9.63)

i= I

and
n

H=n~= In;~,;

(9.64)

i= I

If the molar internal energy and molar enthalpy of the mixture are required then

(9.65)
and
i:,

"m

1 n
n
= - "'\'"'
. ="'\'"'xi:, .
L.,. n.l.
11 'm,1
L.,. i''m,1

i=I

(9.66)

i=l

Neglecting motion, gravity, electricity, magnetism and capillary effects, then em = um, and
hence
(9.67)

The definition of enthalpy for an ideal gas is


(9.68)

hm= um+ !RT


Thus, for the mixture

hm = L(Xi(um + ~T);)
= L X;Um,i + L X;~T = L X;Um,i +~TL X;
= L X;Um,i + ~T

9.4.4

(9.69)

SPECIFIC HEATS OF MIXTURES

Statement (c) of the Gibbs-Dalton law and the above expressions show that
and

hm = L X;hm,i
By definition, the specific heats are
c

du
dT

=-

and

dh
dT

=-.

Thus
Cv m

= (

dum)
dT

L X;( dum) = L X;(Cv m}i


dT ;
.

(9.70)

and, similarly
Cp,m

9.5

= L X;(Cp,m);

(9.71)

Entropy of mixtures

Consider a mixture of two ideal gases, a and b. The entropy of the mixture is equal to the
sum of the entropies which each component of the mixture would have if it alone occupied
the whole volume of the mixture at the same temperature. This concept can present some
difficulty because it means that when gases are separate but all at the same temperature
and pressure they have a lower entropy than when they are mixed together in a volume
equal to the sums of their previous volumes; the situation can be envisaged from Fig 9.2.
This can be analysed by considering the mixing process in the following way:
(i) Gases a and b, at the same pressure, are contained in a control volume but prevented
from mixing by an impermeable membrane.
(ii) The membrane is then removed, and the components mix due to diffusion, i.e. due to
the concentration gradient there will be a net migration of molecules from their
original volumes. This means that the probability of finding a particular molecule at
any particular point in the volume is decreased, and the system is in a less ordered
state. This change in probability can be related to an increase in entropy by statistical
thermodynamics. Hence, qualitatively, it can be expected that mixing will give rise to
an increase in entropy.

,' Gas 'a' ,


,
Pa

,'

l;_...,,_ _ _ _..ir.._

_ ._ _ _._..._..........__Wll

Fig. 9.2 Two gases at the same pressure, p, contained in an insulated container and separated by a
membrane; p =Pb= p
0

Considering the mixing process from a macroscopic viewpoint, when the membrane is
broken the pressure is unaffected but the partial pressures of the individual components
are decreased.
The expression for the entropy of a gas is (eqn (9.20))

where sm(T) is a function of T alone. Hence, considering the pressure term, which is
actually the partial pressure of a component, a decrease in the partial pressure will cause
an increase in the entropy of the gas. Equation (9.20) can be simplified by writing the
pressure ratio as pr = p I Po' giving
Sm= sm(T) - 9\ In Pr+ So,m

Now consider the change from an analytical viewpoint. Consider the two gases a and b.
The total amount of substance in the gases is
(9.72)

The entropies of the gases before mixing are


Sm. = nasma = naCsm.(T) - 9\ ln Pra + So,mJ

(9.73)

sm b = nbsmb = nb[smb(T) - 9\ In p,b+ So,mbl

(9.74)

and

giving the sum of the entropies before mixing (i.e. separated) as


(SmT)scp = (nTsmT)SCp = nT[xasm,(T) + xbsm/T) + XaSO,ma + xbso,mb - 9\(xa ln Pra + xb ln p,b)]

(9.75)
Since the gases are separated and both are at the pressure, p,, then
(Sm T)sep = nT[xasm.(T) + XbSm b(T) + XaSo,ma + xbso,mb - 9\ ln p)

(9.76)

Hence, the molar entropy, when the gases are separate, is


(smr)sep = X0 Sm 0 (T) + XbSm 0 (T) + X0 So,ma + XbSo,mb -9\ ln p
=

L X;Sm; + L X;So,m ; - 9\ In p

(9.77)

The entropy of the gases after mixing is


(9.78)
where sma and smb are the molar entropies if each component occupied the whole volume.
Thus, substituting for sm from eqn (9.20a) gives
(nTsmT)mix = nasma(T) + nbsmb(T) - 9\(na ln Pra + nb ln Prb) + n.so,ma + nbsO,mb

(9.79)

(smT)mix = XaSma(T) + xbsmb(T) - 9\(xa ln Pra + xb ln Prb) + xasO,ma + xbsO,mb

(9.80)

where Pr;= pj p 0 , and p; =partial pressure of constituent i.


Equation (9.80) may be written
(smT)mix =

L X;Sm/T) + L X;So,m; - 9\ L X; ln Pr;

(9.81)

Now, consider the final term in eqn (9.81), based on two constituents to simplify the
mathematics
Xa ln Pra + xb ln Prb = Xa ln(XaPr) + xb ln(xbpr)

= Xa ln Xa + Xa ln Pr+ xb ln xb + xb ln Pr
= I X; lo X; + In Pr I X;
= I X; ln X; + In Pr = ln Pr - I X; ln -

(9.82)

X;

Hence
(sm)mix =I X;sm,(T) +I X;So,m, + 9\(I X; In

~; - In Pr)

(9.83)

Considering the terms on the right-hand side of eqn (9.83):


1st and 2nd terms
3rd term
4th term

summation of entropies before mixing, (SmT)sep


change of entropy due to mixing; this is due to a change of partial
pressures when mixed
pressure term

The change of entropy due to mixing is given by


J}.S = (SmT)mix - (SmT)sep

(9.84)

which is the difference between eqns (9.83) and (9.77). This gives
J}.S

9\ I

X;

ln -

(9.85)

X;

The numerical value of J}.S must be positive because xi is always less than unity.
Equation (9.85) shows that there is an entropy increase due to mixing, and this is caused
by the reduction in the order of the molecules. Before mixing it is possible to go into one
side of the container and guarantee taking a particular molecule, because only molecules of
a are in the left-hand container, and only molecules of b are in the right-hand container.
After mixing it is not possible to know whether the molecule obtained will be of a or b:
the order of the system has been reduced.

What happens if both a and bin Fig 9.2 are the same gas?
Superficially it might be imagined that there will still be an increase of entropy on mixing.
However, since a and b are now the same there is no change in partial pressure due to
mixing. This means that the molar fraction is unaltered and
LiS = 0

9.6

(9.86)

Concluding remarks

A detailed study of ideal gases and ideal gas mixtures has been undertaken in preparation
for later chapters. Equations have been developed for all properties, and enthalpy
coefficients have been introduced for nine commonly encountered gases. Tables of gas
properties have been presented for most gases occurring in combustion calculations.
Equations for gas mixtures have been developed and the effects of mixing on entropy
and Gibbs energy have been shown.

Combustion is an oxidation process and is usually exothermic (i.e. releases the chemical
(or bond) energy contained in a fuel as thermal energy). The most common combustion
processes encountered in engineering are those which convert a hydrocarbon fuel (which
might range from pure hydrogen to almost pure carbon, e.g. coal) into carbon dioxide and
water. This combustion is usually performed using air because it is freely available,
although other oxidants can be used in special circumstances, e.g. rocket motors. The
theory that will be developed here will be applicable to any mixture of fuel and oxidant
and any ratio of components in the products; however, it will be described in terms of
commonly available hydrocarbon fuels of the type used in combustion engines or boilers.
The simplest description of combustion is of a process that converts the reactants
available at the beginning of combustion into products at the end of the process. This model
presupposes that combustion is a process that can take place in only one direction and it
ignores the true statistical nature of chemical change. Combustion is the combination of
various atoms and molecules, and takes place when they are close enough to interact, but
there is also the possibility of atoms which have previously joined together to make a
product molecule separating to form reactants again. The whole mixture is really taking part
in a molecular 'barn dance' and the tempo of the dance is controlled by the temperature of
the mixture. The process of molecular breakdown is referred to as dissociation; this will be
introduced in Chapter 12. In reality a true combustion process is even more complex than
this because the actual rate at which the reactions can occur is finite (even if extremely
fast). This rate is the basic cause of some of the pollutants produced by engines,
particularly NOx. In fact, in most combustion processes the situation is even more complex
because there is an additional factor affecting combustion, which is related to the rate at
which the fuel and air can mix. These ideas will be introduced in Chapter 15. Hence, the
approach to combustion in this chapter is a simplified one but, in reality, it gives a
reasonable assessment of what would be expected under good combustion conditions. It
cannot really be used to assess emissions levels but it can be extended to this simply by the
introduction of additional equations: the basic approach is still valid.
The manner in which combustion takes place is governed by the detailed design of the
combustion system. The various different types of combustion process are listed in
Table 10.1, and some examples are given of where the processes might be found. There is an
interdependence between thermodynamics and fluid mechanics in combustion, and this
interaction is the subject of current research. This book will concentrate on the thermo-

dynamics of combustion, both in equilibrium and non-equilibrium states. The first part of the
treatment of combustion will be based on equilibrium thermodynamics, and will cover
combustion processes both with and without dissociation . It will be found that equilibrium
thermodynamics enables a large number of calculations to be performed but, even with
dissociation included, it does not allow the calculation of pollutants, the production of which
are controlled both by mixing rates (fluid mechanics) and reaction rates (thermodynamics).
Table 10.1 Factors affecting combustion processes
Conditions of combustion

Classification

Examples

Time dependence

steady
unsteady

gas turbine combustion chamber, boilers


petrol engine, diesel engine

Spatial dependence

zero-dimensional

only used for modelling purposes, welJstirred reactors


approximated in pipe flows, flat flame
burners
axisymmetric flames, e.g. Bunsen burner
general combustion

one-dimensional
two-dimensional
three-dimensional
Mixing of initial reactants

premixed
non-premixed

petrol , or spark ignition, engine


diesel engine, gas turbine combustion
chamber

Flow

laminar
turbulent

special cases for measuring flame speed


most real engine cases, boilers

Phase of reactants

single

spark-ignited gas engines, petrol engines


with fuel completely evaporated; gas-fired
boilers
diesel engines, gas turbines, coal- and oilfired boilers

multiphase
Reaction sites

homogeneous
heterogeneous

spark-ignition engines
diesel engines, gas turbines, coal fired
boilers

Reaction rate

equilibrium chemistry
(infinite rate)

approached by some processes in which the


combustion period is long compared with
the reaction rate
all real processes: causes many pollutant
emissions

finite rate
Convection conditions

natural
forced

Bunsen flame, gas cooker, central heating


boiler
gas turbine combustion chamber, large
boilers

Compressibility

incompressible
compressible

free flames
engine flames

Speed of combustion

deflagration
detonation

most nonnal combustion processes


'knock' in spark ignition engines,
explosions

Das könnte Ihnen auch gefallen