You are on page 1of 10

7

Chemical Exergy
In the last chapter the concepts of exergy and physical exergy, in particular,
were introduced. This chapter deals with three other important concepts
namely, exergy of mixing, chemical exergy, and cumulative exergy con-
sumptionand their numerical evaluation.
1 INTRODUCTION
Recall that exergy values reect to which extent a compound or mixture is
out of equilibrium with our environment. Examples are dierences in
pressure and temperature with the environment. Dierences in temperature
lead to heat transfer, while dierences in pressure lead to mass ow. Chapter
6 shows that the physical exergy represents the maximum amount of work
that can be obtained from a system by converting a systems pressure and
temperature to those of our environment.
It appears, however, that when a systems physical exergy is zero and
thus the system prevails in a state of thermomechanical equilibrium with the
environment, it may still be out of equilibrium with that environment in
other respects. The origin is to be found in the dierence in the composition
and nature of the components making up the system and the environment,
respectively. These dierences lead to values for the exergy of mixing and the
chemical exergy. Earlier we pointed out that though the physical exergy of
methane is zero, its chemical exergy is not. Equally, pure nitrogen and
oxygen have nonzero chemical exergies because their mole fraction in the
environment is dierent from 1. Because the process of mixing plays an im-
portant role in the determination of the chemical exergy, the next section
deals with the exergy of mixing.
Copyright 2004 Marcel Dekker, Inc.
2 EXERGY OF MIXING
To clarify the concept of the exergy of mixing, we give the example of pure
oxygen at ambient conditions P
0
and T
0
. Consider a system, for convenience
chosen at P
0
, T
0
, isolated from the environment and consisting of two
separate compartments containing oxygen and air, respectively. The two
compartments, initially separated by an elastic diathermal barrier, and thus
in mechanical and thermal equilibrium, are brought in contact with each
other by removing the barrier. Oxygen and air will diuse into each other,
and eventually an equilibrium will be reached where oxygen and air have
mixed into a homogeneous mixture. The initial condition of oxygen is
apparently not one of complete equilibrium with the environment (i.e., with
air) despite the equality in pressure ( P
0
) and temperature (T
0
). The
thermodynamic potential of pure oxygen is higher than that of oxygen in
air at P
0
and T
0
. On mixing of the components of air in their pure state to a
homogeneous mixture, the thermodynamic potential of each component
decreases. The associated change in exergy is
D
mix
Ex D
mix
HT
0
D
min
S 1
As the mixing process takes place at P
0
, T
0
, we may write
D
mix
Ex
P
0
T
0
D
mix
G
P
0
T
0
2
For the calculation of the exergy value at P, T of a mixture, of a given
composition, with respect to the exergy values of the pure components at P
and T, the exergy dierence is dened as
Ex
mix
uD
mix
Ex 3
with values for D
mix
H and D
mix
S at the conditions P and T. Chapter 10
presents an example of the industrial distillative separation of the mixture of
propane and propene, in which the exergy of mixing is very prominent.
3 CHEMICAL EXERGY
For the determination of a compounds chemical exergy value we need to
dene a reference environment. This reference environment is a reection of
our natural environment, the earth, and consists of components of the at-
mosphere, the oceans, and the earths crust. If, at P
0
and T
0
, the substances
present in the atmosphere, the oceans, and the upper part of the crust of our
earth are allowed to react with each other to the most stable state, the Gibbs
Chapter 7 74
Copyright 2004 Marcel Dekker, Inc.
energy of this whole system will have decreased to a minimum value. We
then can dene the value of the Gibbs energy for a subsystem, the reference
environmentat sea level, at rest, and without other force elds present
than the gravity eldto be zero as well as for each of the phases present
under these conditions. It is a logical extension of these assumptions to
dene the thermodynamic potentials of each of the substances present in the
dierent phases to have a value of 0 J/mol. With respect to this reference
environment, we then determine the thermodynamic potentials of all kinds
of substances in all kinds of phases at P and T. From this reference
environment it is not possible to obtain any work. Therefore, this state is
also meaningful as a reference state for the determination of exergy values at
P
0
and T
0
. This nally leads to the denition Ex
i
( P
0
, T
0
) u A
i
( P
0
, T
0
) for
the subsystem at sea level, at rest, and without the presence of any other
force eld than the gravity eld. The concept of this reference environ-
ment is illustrated for a number of the so-called reference components.
3.1 Reference Components from Air
Apart from dierences in chemical concentration, or better, thermodynamic
potentials, such as for oxygen, there can be other situations for being out of
equilibrium with the environment at P
0
, T
0
. Consider, for instance, the
material graphite. Graphite can spontaneously react with oxygen to from
carbon dioxide, but for kinetic reasons the reaction is very slow and graph-
ite seems to be stable in our environment, although in the presence of oxy-
gen it is metastable with respect to carbon dioxyde. As a result, it has a
signicant amount of chemical exergy available and can be considered as an
important energy carrier because it is highly out of equilibrium with the
environment.
In our environment there are many substances that, like oxygen in our
atmosphere, cannot further diuse and/or react toward more stable cong-
urations and may be considered to be in equilibrium with the environment.
Neither chemical nor nuclear reactions can transform these components into
even more stable compounds. From these components we cannot extract
any useful work, and therefore an exergy value of 0 kJ/mole has been
assigned to them. This has been done for the usual constituents of air: N
2
,
O
2
, CO
2
, H
2
O, D
2
O, Ar, He, Ne, Kr, and Xe at T
0
= 298.15 K and P
0
=
99.31 kPa, the average atmospheric pressure [1]. Their partial pressures P
i
in
air are given in Table 1.
From these data we can calculate the, chemical, exergy values of these
components in the pure state at P
0
and T
0
. Air at these conditions can, to a
good approximation, be considered as an ideal gas, therefore, separation
into its constituents will take place without a heat eect: D
sep
H = 0. And so
Chemical Exergy 75
Copyright 2004 Marcel Dekker, Inc.
the only eect left in the exergy change of separation, D
sep
Ex = Ex
mix
[see
Eq. (3)], is that of the entropy of separation:
D
sep
Ex D
sep
HT
0
D
sep
S
4
T
0
D
mix
S
As we recall from Chapter 2, the change in entropy associated with taking
one mole of an ideal gas isothermally from pressure P
1
to a pressure P
2
is
given by
DS Rln
P
2
P
1

5
From this equation we can show [2] that the standard chemical exergy at P
0
and T
0
of a pure component can be calculated from its partial pressure P
i
in
air with Eq. (6):
Ex
0
ch;i
RT
0
ln
P
0
P
i

6
The standard chemical exergy values for the main constituents of air as
listed in Table 1 are given in Table 2.
Table 1 Partial Pressure of Various Components in Air
Component P
i
(kPa) Component P
i
(kPa)
N
2
75.78 He 0.000485
O
2
20.39 Ne 0.00177
CO
2
0.0335 Ar 0.906
H
2
O 2.2 Kr 0.000097
D
2
O 0.000342 Xe 0.0000087
Table 2 Standard Chemical Exergy Values at P
0
,T
0
of Various
Components Present in Air
Component Ex
ch
0
(kJ/mol) Component Ex
ch
0
(kJ/mol)
N
2
0.72 He 30.37
O
2
3.97 Ne 27.19
CO
2
19.87 Ar 11.69
H
2
O 9.49 Kr 34.36
D
2
O 31.23 Xe 40.33
Chapter 7 76
Copyright 2004 Marcel Dekker, Inc.
Exergy values for the elements in their stable modication at T
0
=
298.15 K and P
0
= 101.325 kPa are called standard chemical exergy values
Ex
ch
0
. For the calculation of the chemical exergy value of all kinds of
substances, the standard chemical exergy values of all elements are required.
3.2 Exergy Values of the Elements
The following example for graphite illustrates how the chemical exergy
value for all other elements can now be calculated (Table 3). For the
calculation of Ex
ch
0
of graphite, we make use of the reaction in which CO
2
is formed from the elements in their stable modication at P
0
, T
0
:
Cgraphite; s O
2
g !CO
2
g 7
Table 3 Standard Chemical Exergy Values of the Elements [1]
Element Ex
0
ch
(kJ/mol) Element Ex
0
ch
(kJ/mol)
Ag (s) 70.2 Kr (g) 34.36
Al (s) 888.4 Li (s) 393.0
Ar (g) 11.69 Mg (s) 633.8
As (s) 494.6 Mn (s
a
) 482.3
Au (s) 15.4 Mo (s) 730.3
B (s) 628.5 N2 (g) 0.72
Ba (s) 747.7 Na (s) 336.6
Bi (s) 274.5 Ne (g) 27.19
Br
2
(1) 101.2 Ni (s) 232.7
C (s, graphite) 410.26 O
2
(g) 3.97
Ca (s) 712.4 P (s, red) 863.6
Cd (s
a
) 293.2 Pb (s) 232.8
Cl
2
(g) 123.6 Rb (s) 388.6
Co (s
a
) 265.0 S (s, rhombic) 609.6
Cr (s) 544.3 Sb (s) 435.8
Cs (s) 404.4 Se (s, black) 346.5
Cu (s) 134.2 Si (s) 854.6
D
2
(g) 263.8 Sn (s, white) 544.8
F
2
(g) 466.3 Sn (s) 730.2
Fe (s
a
) 376.4 Ti (s) 906.9
H
2
(g) 236.1 U (s) 1190.7
He (g) 30.37 V (s) 721.1
Hg (1) 115.9 W (s) 827.5
I
2
(s) 174.7 Xe (g) 40.33
K (s) 366.6 Zn (s) 339.2
Chemical Exergy 77
Copyright 2004 Marcel Dekker, Inc.
The corresponding change in standard Gibbs energy is called the
standard Gibbs energy of formation of CO
2
, D
f
G
0
298.15
, and is dened as
D
f
G
0
298:15
u
X
r
i
A
0
i;298:15
8
in which r
i
is the so-called stoichiometric coecient, dened as positive for
products and negative for reactants, and A
i
0
= G
i
0
is the standard thermo-
dynamic potential or Gibbs energy for substance i. Equation (8) is based on
the formation of 1 mole of the compound considered, in this instance 1 mole
of CO
2
. If we dene the change in exergy in the same way:
D
f
Ex
0
298:15
D
f
G
0
298:15

X
i
r
i
A
0
i;298:15
9
u
X
r
i
Ex
0
ch;i
then the exergy of graphite can be calculated from
Ex
0
ch;Cs
D
f
G
0
298:15
1 Ex
0
ch;CO
2
g
1 Ex
0
ch;O
2
g
10
The values of D
f
G
298.15
0
for many compounds are listed in standard tables
[2], and the value for CO
2
reads 394.359 kJ/mole. With the help of Table 2,
which gives the standard chemical exergy values for CO
2
and O
2
, Equation
(10) allows the calculation of Ex
ch,C(s)
0
= 394.359 + 19.87 3.97 = 410.26
kJ/mole.
For the remaining elements, reference compounds have been chosen,
as they occur in seawater or in the lithosphere, the earths crust. An
important aspect of this choice has been that the calculated exergy values
of most compounds should be positive. Table 3 lists the standard chemical
exergy values of the elements as presented in Szarguts well-known standard
work [1]. Chapter 8 gives an example, the adiabatic combustion of H
2
, to
illustrate the use of these exergy values in an interesting application.
3.3 Chemical Exergy Values of Compounds
Table 3 is useful for the calculation of the standard chemical exergy values
of compounds. We illustrate this for methane and start from its hypothetical
formation reaction at standard conditions:
Cs 2H
2
g !CH
4
g 11
Applying Eq. (9) results in
Ex
0
ch;CH
4
g
D
f
G
0
298:15
Ex
0
ch;Cs
2Ex
0
ch;H
2
g
12
The rst term on the right-hand side of this equation is the standard Gibbs
energy of formation of methane, which is listed [2] as 50.460 kJ/mole and
Chapter 7 78
Copyright 2004 Marcel Dekker, Inc.
thus Ex
0
ch,CH
4
(g)
can be calculated to be 831.6 kJ/mole. Chapter 9 illus-
trates the use of this exergy value in the analysis of a natural gas-driven
powerstation.
In general, we can calculate the standard chemical exergy of a com-
ponent j from the standard chemical exergy of its elements with the equation
Ex
0
ch; j
D
f
G
0
j;298:15

X
v
i
Ex
0
ch;i
13
We recall that the exergy of methane will be dierent for other values
of P and T than P
0
, T
0
and refer to Table 1 of Chapter 6 to demonstrate the
inuence of pressure and temperature on this exergy value. It is clear that
the chemical contribution to the total exergy, Ex = Ex
phys
+ Ex
0
ch
, in this
case is dominant. At the same time we should be aware that in a simple
compression step this contribution is irrelevant and should not be included
in an exergy eciency calculation. On the level of, let us say, 10 kJ/mole of
physical exergy, the loss of 2.5 kJ/mole of exergy due to ineciencies of the
compressor results in a thermodynamic or exergetic eciency of 75%. Had
we included the 832 kJ of chemical exergy of methane, the thermodynamic
eciency would have been as high as 99.7%, which gives a completely
blurred picture of the compressors performance.
Finally, Table 4 gives the standard exergy values of a selected number
of compounds that are relevant for the examples and topics presented in this
book.
Table 4 Standard Chemical Exergy Values of
Selected Compounds
Substance kJ/mole
CH
4
(g) natural gas 832
CH
3
OH (g) 722
CH
3
OH (l) 718
UCH
2
U
a
oil 652
(CH
2
O)
b
biomass 480
CO
2
(g) 20
SiO
2
(s, a quartz) 1.9
TiO
2
(s, rutile) 21.4
Al
2
O
3
.H
2
O (s) bauxite 200.8
Fe
2
O
3
(s) haematite 16.5
NH
3
(g) 337.9
CO(NH
2
)
2
(s) urea 689.0
a
Crude oil on a per-carbon basis.
b
Biomass (glucose) on a per-carbon basis.
Chemical Exergy 79
Copyright 2004 Marcel Dekker, Inc.
3.4 The Convenience of the Chemical Exergy Concept
In chemical thermodynamics the reference components have been selected
as the elements in their most common state at standard conditions, the
standard state. These elements have been dened as having a zero standard
Gibbs energy of formation. The standard Gibbs energy of formation of a
compound is related to that of the elements from which it has been
composed. Let us take liquid methanol, CH
3
OH [1]. Its standard Gibbs
energy of formation is 166.270 kJ/mole, a number that does not say very
much other than that in the reaction
Cs 2H
2
g
1
2
O
2
g !CH
3
OH1 14
the standard Gibbs energies at the left-hand side are zero and the standard
Gibbs energy of reaction is also 166.270 kJ/mole. However, following the
procedures as outlined in the above sections, we can calculate with Eq. (13)
the standard chemical exergy of liquid methanol to be 166.270 + 410.26 +
2 236.10 + 1/2 3.97 = 718.2 kJ/mole. This number is very meaningful,
as it expresses the maximum amount of work available to us embodied in one
mole of liquid methanol. We can then compare this with the value for
methane and notice that the partial oxidation of methane to methanol has
lowered the exergy value somewhat, from 832 kJ/mole to 718 kJ/mole. But
methanol is in the liquid state, and this is an attractive feature for a
transportation fuel. On the other hand, methanol has double the mass of
methane, and so per unit of mass its available work or exergy is less than half.
And last but not least, the eciency of converting methane into methanol
may be about 5060% (see Chapter 14) and much of the advantage of using
methanol seems to have gone. Nevertheless, although, strictly speaking, the
concept of exergy does not add anything in the fundamental sense, it
certainly adds conveniencefor example, for the discussion on the pros
and cons of energy conversion such as in the above comparison of methane
and methanol. This is one of the attractive features of the exergy concept that
has made it so popular with many practitioners.
4 CUMULATIVE EXERGY CONSUMPTION
Suppose we deal with a process in which iron, Fe, has to be used as a
reactant, for example in a reduction reaction. The standard chemical exergy
of Fe is 376.4 kJ/mole. If we wish to carry out a thermodynamic or exergy
analysis of this process, this value is not appropriate. After all, to put the
exergy cost of the product, for which Fe was needed as a reactant, in proper
Chapter 7 80
Copyright 2004 Marcel Dekker, Inc.
perspective, we need to consider all the exergetic costs incurred in order to
produce this product all the way from the original natural resourcesiron
ore and fossil fuel in this example. The production of iron from, for ex-
ample, the iron ore haematite and coal has a thermodynamic eciency of
about 30% [1], and therefore it is not 376.4 kJ/mole Fe that we need to
consider but 376.4/0.3 = 1250 kJ/mole Fe. This value is called the cumu-
lative exergy consumption (CExC) of Fe. It may well be that for proper
analysis of the eciency of the step consuming Fe to produce the product,
we want to take the standard chemical exergy of Fe, but for the calculation
of the CExC of the nal product, we need to include the CExC of Fe.
Chapter 14 discusses many examples where the exergy of the nal product is
compared with the CExC of the same product. Together these two values
allow the calculation of the thermodynamic, exergetic, eciency of a process
yielding the product from natural resources. This is part of the subject of
Chapter 14.
We recall that, without mentioning it, we touched upon the topic of
cumulative exergy consumption before. In Chapter 6 we illustrate the
application of the concept of physical exergy with the simple example of
mixing liquid water of 100jC with that of 0jC. In that example we rst take
the exergy value of hot water as 34 kJ/kg. But when this water has been
produced from natural gas, its accumulated exergy consumption is calcu-
lated according to Table 2 of Chapter 6 to be 1/0.12 34 = 283 kJ/kg.
Finally, we consider another important contributor to the CExC of a
product. We refer to the equipment being used in the process. This equip-
ment also has to be manufactured from resources originally taken from the
environment. This cumulative exergy consumption has to be discounted
over the lifetime of this equipment and then added as a contribution to the
cumulative exergy consumption of the product. Our experience is that this
contribution is negligible for equipment that works continuously. For equip-
ment performing with an irregular operation such as a laundry machine at
home, this contribution may be substantial and makes up a large part of the
total exergy cost of the product.
5 CONCLUSIONS
The concept of chemical exergy has a distinct advantage over the standard
Gibbs energy of formation. Whereas the latter is zero for the elements at
standard conditions, the chemical exergy has a zero value for compounds or
elements in equilibrium with and as they occur in our natural environment.
Thus the standard chemical exergy of a compound clearly represents the
amount of work available with respect to the environment in which we live
Chemical Exergy 81
Copyright 2004 Marcel Dekker, Inc.
and work. The chemical exergy can be simply calculated from the Gibbs
energy of formation. The only dierence between the two concepts is that
their zero values are dened for dierent reference substances.
The chemical exergy of a molecule in a mixture is smaller than in its
pure state, as it will require work to separate the mixture in its pure
constituents, the exergy of separation. This exergy will be lost as the exergy
of mixing when the pure constituents spontaneously form the mixture. The
total exergy of a pure compound or element is therefore composed of three
contributions: the chemical exergy; the exergy of mixing, and the physical
exergy. The last element accounts for the fact that the molecule may be at
dierent conditions of pressure and temperature than those of the environ-
ment, P
0
and T
0
.
The concept of cumulative chemical exergy consumption is very useful
and accounts for the fact that when a compound (e.g., ammonia) is
introduced into a process, its chemical exergy has to be corrected for the
exergy consumption accumulated since this compound was manufactured
from its natural constituents (air and natural gas in the case of ammonia).
If the thermodynamic eciency of a process step is calculated, the
chemical exergies should be excluded from the calculation if the process step
does not include chemical conversions. If it does, it may be appropriate to
distinguish between the physical and the chemical eciency, D
phys
and D
chem
,
of the process step.
Finally, although the exergy concept is not strictly necessary for the
calculation of the available work lost in the process, it is an extremely handy
tool to calculate losses and eciencies and for making a quick assessment of
process options. Chapter 8 gives some simple illustrations, whereas Part III,
Case Studies, presents the results of integrated studies in the world of energy
and chemical technology.
REFERENCES
1. Szargut, J.; Morris, D.R.; Steward, F.R. Exergy Analysis of Thermal, Chemical,
and Metallurgical Process; Hemisphere Publishing Corp.: New York, 1988.
2. Smith, J.M.; Van Ness, H.C.; Abbott, M.M. Introduction to Chemical Engi-
neering Thermodynamics, 5th ed.; McGraw-Hill: New York, 1996.
Chapter 7 82
Copyright 2004 Marcel Dekker, Inc.

You might also like