You are on page 1of 29

Original article

International Journal of Mechanical


Engineering Education
Chemical exergy of ideal 2019, Vol. 47(1) 44–72
! The Author(s) 2017
and non-ideal gas Article reuse guidelines:
sagepub.com/journals-permissions
mixtures and liquid DOI: 10.1177/0306419017749581
journals.sagepub.com/home/ijj

solutions with
applications

Rajinder Pal

Abstract
Exergy or availability, although not a recent concept, is receiving extensive coverage in
scientific publications due to its vast applications in different scientific and engineering
fields. Exergy of a system consists of two parts: thermo-mechanical exergy and chemical
exergy. While thermo-mechanical exergy of systems is covered to a certain extent in
modern undergraduate textbooks on engineering thermodynamics, chemical exergy is
mentioned only briefly. In particular, the theoretical and conceptual developments
related to chemical exergy are not covered in any detail. The focus of this article is
the chemical exergy of materials. Special attention is given to the theoretical treatment
of non-ideal gas mixtures and liquid solutions. The equations necessary to estimate the
chemical exergy of ideal and non-ideal mixtures and solutions are developed from the
fundamental concepts. Where necessary, numerical examples are given to illustrate the
concepts for the benefit of the students. Finally, a practical problem dealing with the
furnace/boiler unit of a practical steam power plant is solved using the concepts of
chemical exergy and exergy analysis. As the material presented in this article involves
advanced level concepts in thermodynamics, it is most suitable for the second, advanced
level, course in engineering thermodynamics in third year, after the students have complet-
ed a full one-term course on introductory thermodynamics in their second year.

Keywords
Exergy, chemical exergy, thermodynamics, chemical potential, non-ideal mixtures

Department of Chemical Engineering, University of Waterloo, Waterloo, Canada


Corresponding author:
Rajinder Pal, University of Waterloo, 200 University Avenue West, Waterloo, Ontario N2L 3G1, Canada.
Email: rpal@uwaterloo.ca
Pal 45

Introduction
Exergy (also referred to as availability) is the maximum amount of useful work
obtainable when the system under consideration is brought from its initial state to
the dead state by processes involving interaction only with the environment.1–4
Alternatively, exergy could be defined as the minimum amount of useful work
required to synthesize a quantity of matter from substances present in the envi-
ronment and to deliver the matter to a specified state involving interaction only
with the environment. The dead state is one where the system attains thermody-
namic equilibrium with the environment. In thermodynamic equilibrium, the con-
ditions of mechanical, thermal, and chemical equilibrium between the system and
the environment are satisfied. The existence of mechanical and thermal equilibrium
between the system and the environment implies equality of pressure and temper-
ature between the system and the environment. The condition of chemical equi-
librium implies that the chemical potential of each chemical species is the same in
the system and the environment. Clearly exergy is a measure of the degree of
departure of the state of the system from its environment. When the system and
the environment are in different states, an opportunity exists to produce useful
work upon interaction between the system and the environment.
In exergy analysis, two types of dead states are often defined for the systems:
restricted dead state and complete dead state.4,5 In the restricted dead state, only the
conditions of mechanical and thermal equilibrium are satisfied whereas in the
complete dead state, the condition of chemical equilibrium is also satisfied. The
restricted dead state is relevant in exergy calculations dealing with constant com-
position systems.
The total exergy of a system can be divided into two parts, namely thermo-
mechanical exergy (wTM ) and chemical exergy (wC )

w ¼ wTM þ wC (1)

The thermo-mechanical exergy wTM of a system is defined as the maximum


amount of useful work obtainable when the system under consideration is brought
from its initial state to the restricted dead state by processes involving interaction
only with the environment. The chemical composition of the system remains the
same in the initial state and in the restricted dead state. The thermo-mechanical
exergy of a closed system at zero velocity and zero elevation relative to the envi-
ronment is given by the expression

wTM ¼ ðu  u0 Þ þ p0 ðv  v0 Þ  T0 ðs  s0 Þ (2)

where u, v, and s are the specific internal energy, specific volume, and specific
entropy, respectively, of the system and T0 is the absolute temperature of the
environment. The subscript “0” refers to the restricted dead state where the
46 International Journal of Mechanical Engineering Education 47(1)

system attains thermo-mechanical equilibrium with the environment. For a fluid


stream, the corresponding expression of thermo-mechanical exergy is given as

V2
wTM ¼ ðh  h0 Þ  T0 ðs  s0 Þ þ þ gz (3)
2
where h, V, g, and z are the specific enthalpy, fluid velocity, acceleration due to
gravity, and elevation, respectively.
The chemical exergy wC of a system is defined as the maximum amount of useful
work that can be obtained when the system under consideration is brought from its
restricted dead state to the complete dead state where the system is in complete
thermodynamic equilibrium with the environment. Note that the temperature and
pressure are the same in restricted and complete dead states. Consider a turbine
where the input stream consists of a pure gaseous species A at environment con-
ditions of temperature T0 and pressure P0 . The gas is expanded isothermally and
reversibly in a turbine to a discharge pressure of yA;e P0 , corresponding to the
partial pressure of species A in the environment, where yA;e is the mole fraction
of species A in the environment. Thus, the discharge stream is in chemical equi-
librium with the environment. The gas is discharged reversibly into the environ-
ment through a semi-permeable membrane (see Figure 1). For this reversible
isothermal expansion of a gas in a turbine, the maximum shaft work produced
is simply a decrease in the Gibbs free energy. Assuming gas to be an ideal gas, the
maximum work obtainable or chemical exergy is thus given by
Z yA;e P0
dP  
wC ¼ DG ¼ RT0 ¼  RT0 ln yA;e (4)
P0 P

where G is the Gibbs free energy and R is the universal gas constant.

Figure 1. Method for evaluating the chemical exergy of a gas.


Pal 47

The importance of different forms of exergy and their applications cannot be


underestimated. In a recent publication6 in Science Europe entitled “In a Resource-
Constrained World: Think Exergy, not Energy”, the authors state that

Exergy analysis can be applied not only to individual processes, but also to industries,
and even to whole national economies. It provides a firm basis from which to judge
the effect of policy measures taken towards energy, resource and climate efficiency. In
the future, consumers could be informed about products and services in terms of their
exergy-destruction footprint in much the same way as they are about their carbon
emissions.

Some of the key recommendations made by the authors are: (a) the concepts
related to exergy must be taught in universities; (b) exergy-destruction footprints
should be introduced in industrial applications to give a useful basis for energy
efficiency improvements; (c) excess exergy-destruction footprints should be taxed
to promote improvements in industrial efficiency and to drive the development of
more efficient technologies; (d) accounts of exergy-destruction footprints should be
created to assist the Intergovernmental Panel on Climate Change in developing
measures to mitigate climate changes; and (e) exergy should be integrated into
policy, law, and everyday practice.
The focus of this article is the chemical exergy of matter. Understanding of
chemical exergy and related concepts is an issue with undergraduate engineering
students. While thermo-mechanical exergy and its applications are covered in sev-
eral modern textbooks on engineering thermodynamics,7–12 the chemical exergy of
materials is covered only briefly, if it is covered at all, in the undergraduate text-
books on thermodynamics. In particular, little information is presented about the
theoretical background related to chemical exergy of non-ideal gas mixtures and
liquid solutions. In processes involving chemical reactions, and hence changes in
chemical composition, chemical exergy plays a very important role in exergy anal-
ysis of the process. Thus, it is important for the students to be well versed in
chemical exergy and its applications.
The key objectives of this article are as follows: (1) to briefly review the funda-
mental concepts of chemical exergy with students in mind; (2) to develop the
relations for chemical exergy of ideal gas mixtures and liquid solutions; (3) to
develop the relations for chemical exergy of non-ideal gas mixtures and liquid
solutions rarely covered in thermodynamics textbooks; and (4) to apply the con-
cepts of chemical exergy in the analysis of practical processes.
The intended student learning outcomes of this article are: (a) to distinguish
chemical exergy from thermo-mechanical exergy; (b) to define the two dead
states and to distinguish the “restricted dead state” from the “complete dead
state”; (c) to understand and appreciate the concept of “reference environment”;
(d) to estimate the standard chemical molar exergies of substances not present in
the reference environment; (e) to estimate the chemical exergies of ideal gas mix-
tures; (f) to estimate the chemical exergies of ideal liquid solutions; (g) to estimate
48 International Journal of Mechanical Engineering Education 47(1)

the chemical exergies of non-ideal gas mixtures; (h) to estimate the chemical exer-
gies of non-ideal liquid solutions; and (i) to carry out chemical exergy analysis of
processes.

Brief review of exergy concepts covered in modern textbooks


on engineering thermodynamics
The concept of exergy (availability) is covered to a certain extent in several modern
textbooks on engineering thermodynamics.7–12 The standard approach taken in
these textbooks is as follows. The exergy is first defined in a general manner with-
out reference to chemical exergy or to different dead states.7–12 For example, “the
exergy of a system in a given state is the maximum work that can be extracted from
it till it reaches the state of thermodynamic equilibrium with its surroundings”.10
The dead state is then defined without referring to restricted or complete dead
states. The dead state defined is the complete dead state. For example “a system
is said to be in the dead state when it is in thermodynamic equilibrium with the
environment it is in”.7 In the discussion that follows, it is implicitly assumed that
the system is either in chemical equilibrium with the environment or the chemical
composition of the system is invariant. Hence, only thermo-mechanical exergy is
considered to be relevant in the analysis. The expressions for thermo-mechanical
exergy, equations (2) and (3), are derived for closed systems and for fluid streams
based on the definition of the thermo-mechanical exergy. Note that in order to
obtain the maximum work as specified in the definition of exergy, it is necessary
that the system undergoes the change in state from its current state to dead state
reversibly. Following the definition of exergy and the derivation of its mathemat-
ical expressions for closed and flow systems, equations (2) and (3), respectively,
different ways of transferring exergy (thermo-mechanical) into or out of the system
or control volume are considered. For example, thermo-mechanical exergy can be
transferred into or out of the system/control volume by heat, work, and mass
flow.7–10,12 The principle of decrease of exergy is developed next for isolated sys-
tems using the second law of thermodynamics that any spontaneous change in an
isolated system is accompanied by an increase in entropy. This leads to the fun-
damental fact that all real (irreversible) processes are accompanied by exergy
destruction. The exergy balances and their applications are covered next for
closed systems and for control volumes. The chemical exergy of the system is
not considered in exergy balances as the chemical composition of the system is
assumed to be constant. However, some undergraduate textbooks10 on engineering
thermodynamics do present a very brief discussion of chemical exergy of hydro-
carbon fuels.
The concept of chemical exergy and its applications are discussed in specialized
books.4,5,13 The developments related to chemical exergy in these advanced level text-
books are generally limited to ideal gas mixtures and ideal liquid solutions. The
conceptual issues related to chemical exergy of non-ideal gas mixtures and
Pal 49

non-ideal liquid solutions and how to estimate the chemical exergy of such non-ideal
systems are not discussed in any detail even in these advanced level textbooks.

Model for the reference environment


From the definition of exergy, it is clear that exergy of a system is relative to its
environment.4,5,13–16 Thus, exergy is a combined property of the system and the
environment. For the purpose of exergy calculations, the reference environment is
modeled as a very large body in a state of complete thermodynamic equilibrium.
There are no internal gradients of pressure, temperature, and chemical potentials.
The exchanges of matter and energy with the systems do not alter its intensive
properties such as temperature, pressure, and chemical potentials. Thus, there is no
possibility of producing work from interactions (physical or chemical) between
different parts of the environment. Furthermore, all parts of the environment
are considered to be at rest with respect to one another. In this article, the reference
environment is a model atmosphere of infinite extent with uniform temperature,
pressure, and composition. Figure 2 specifies the composition, pressure, and tem-
perature of the reference environment. The reference composition is taken to be
that of air with 60% relative humidity.

Standard molar chemical exergy


The standard molar chemical exergy of a substance is the chemical exergy of 1 mol
of a substance in its standard state. The standard state of a gaseous species is a pure
gas in the ideal gas state at 1 bar. The standard states for liquids and solids are the

Figure 2. Reference environment for the purpose of exergy calculations.


50 International Journal of Mechanical Engineering Education 47(1)

real pure liquid or solid at 1 bar. Note that the relevant temperature in chemical
exergy calculations is the environment temperature T0 fixed at 25 C. The standard
molar chemical exergy of a gaseous species can be calculated using equation (4). For
example, the standard molar chemical exergy of nitrogen gas is 8.314  298.15ln
(0.7651) ¼ 0.664 kJ/mol. Table 1 lists the standard molar chemical exergies of var-
ious species present in the reference environment, calculated using equation (4).

Standard chemical exergy of a reaction


The standard chemical exergy of a reaction worxn can be defined as
X
worxn ¼  i woC;i (5)

where  i is the stoichiometric coefficient and woC;i is the standard molar chemical
exergy of chemical species i. The stoichiometric coefficient is positive for the prod-
ucts and negative for the reactants.
The standard molar chemical exergy of any species can be expressed in terms of
the chemical potentials as
 
woC;i ¼ loi  b
l i;e (6)

where loi is the standard chemical potential of species i and l b i;e is the chemical
potential of species i in the reference environment. Note that the presence of “hat”
on the top of chemical potential signifies mixture (solution) property. From equa-
tions (5) and (6), it follows that
X   X o X
worxn ¼  i loi  b
l i;e ¼  i li  ib
l i;e (7)

By definition the standard Gibbs free energy of a reaction is


X
DGorxn ¼  i loi (8)

Table 1. Standard molar chemical exergies of reference species.

Standard molar chemical


Species Mol. fraction (yi;e ) exergy (woC;i ) (kJ/mol)

N2(g) 0.7651 0.664


O2(g) 0.2062 3.914
H2O(g) 0.0190 9.824
Ar(g) 0.0094 11.57
CO2(g) 0.0003 20.11
Pal 51

As the reference environment is in a state of complete thermodynamic equilibrium


X
ib
l i;e ¼ 0 (9)
Combining equations (7) to (9), the following important result is obtained

worxn ¼ DGorxn (10)


Thus, the standard chemical exergy of a reaction is equal to the standard Gibbs
free energy of a reaction.

Determination of standard molar chemical exergy of


substances not present in the reference environment
If the substance is present in the reference environment, then the calculation of
standard molar chemical exergy of a substance can be done readily. For example,
equation (4) can be used to calculate woC;i for any gaseous species. In general, woC;i is
given as a difference in the chemical potential of the species in the standard and
reference environment states, as expressed by equation (6). This raises the question
as to what happens if the species of interest is not present in the reference envi-
ronment. In this case, it so happens that one can still calculate the chemical exergy
of the species provided that all the atoms present in the species are also present in
the reference environment. However, we need to set up a chemical reaction to
produce the species from the existing species in the reference environment. For
example, consider liquid methanol as the species whose standard molar chemical
exergy needs to be determined. The chemical reaction required to produce meth-
anol from the existing species in the reference environment is as follows

CO2 ðgÞ þ 2H2 OðgÞ ! CH3 OHðlÞ þ ð3=2ÞO2 ðgÞ (11)

The process for synthesizing methanol from the reference environment species is
illustrated in Figure 3. Using equations (5) and (10), one can determine the stan-
dard molar chemical exergy of liquid methanol as
X
woC;methanol ¼ DGorxn   i woC;i (12)
ðother than
methanolÞ

DGorxn for the formation reaction of methanol (equation (11)) can be determined
from the standard Gibbs free energy of formation DGof of various species17 as

DGorxn ¼ DGof ðCH3 OH; lÞ þ ð3=2ÞDGof ðO2 ; gÞ  DGof ðCO2 ; gÞ  2DGof ðH2 O; gÞ
(13)
¼ 166:4 þ ð3=2Þð0Þ  ð394:36Þ  2ð228:589Þ ¼ 685:138 kJ=mol
52 International Journal of Mechanical Engineering Education 47(1)

Figure 3. Process to synthesize liquid methanol from the reference environment species.

The standard molar chemical exergies of the species present in the reference
atmosphere and involved in the reaction are as follows: woC ðO2 ; gÞ ¼ 3:914;
woC ðCO2 ; gÞ ¼ 20:11; and woC ðH2 O; gÞ ¼ 9:824 kJ/mol. Upon substitution of DGorxn
and woC;i values in equation (12), the standard molar chemical exergy of liquid
methanol is found to be 719:025 kJ=mol.

Chemical exergy of ideal gas mixtures and liquid solutions


Ideal gas mixtures
The chemical potential of any species i in an ideal gas mixture at temperature T0
l ig
and pressure P0 , denoted as b i ðT0 ; P0 Þ, is given by

l ig
b i ¼ li þ RT0 lnðyi Þ
o
(14)
Pal 53

where loi is the standard chemical potential of gas species i at T0 (standard state
being pure ideal gas at pressure P0 ). and xi is the mole fraction of species i The
chemical exergy of an ideal gas mixture is
X  ig 
b ig ¼
W li  b
ni b l i;e (15)
C

where ni is the number of moles of species i and b l i;e is the chemical potential of
species i in the reference environment. From equations (14) and (15), it follows that
X   X  o 
b ig ¼
W ni loi þ RT0 lnðyi Þ  ni li þ RT0 lnðyi;e Þ
C
X   (16)
¼ RT0 ni ln yi =yi;e

Thus, the molar chemical exergy of an ideal gas mixture is given as


X  
b ig ¼ W
w b ig =N ¼ RT0 yi ln yi =yi;e (17)
C C

where N is the total number of moles of the mixture including all the species. This
result could be expressed in a more general form as
X X
b ig ¼
w yi woC;i þ RT0 yi lnyi (18)
C

According to equation (18), the molar chemical exergy of a mixture of ideal


gases can be calculated from the knowledge of the standard molar chemical exer-
gies of the species involved and the composition of the mixture.
Example: Determine the molar chemical exergy of a binary ideal-gas mixture of
nitrogen and methane at 298.15 K and 1 bar. The mixture contains 40 mol%
nitrogen.
Solution: In order to solve this problem, we first need to determine the standard
molar chemical exergy of methane by setting up the formation reaction of methane
using the reference species shown in Figure 2. The formation reaction for methane
from the reference species is

CO2 ðgÞ þ 2H2 OðgÞ ! CH4 ðgÞ þ 2O2 ðgÞ (19)

From equations (5) and (10)


X
woC;methane ¼ DGorxn   i woC;i (20)
ðother than
methaneÞ
54 International Journal of Mechanical Engineering Education 47(1)

DGorxn for the formation reaction of methane (equation (19)) can be determined
from the standard Gibbs free energy of formation DGof of various species as

DGorxn ¼ DGof ðCH4 ; gÞ þ 2DGof ðO2 ; gÞ  DGof ðCO2 ; gÞ  2DGof ðH2 O; gÞ


(21)
¼ 50:8 þ 2ð0Þ  ð394:36Þ  2ð228:589Þ ¼ 800:74 kJ=mol

The standard molar chemical exergies of the species present in the reference
atmosphere and involved in the reaction are as follows: woC ðO2 ; gÞ ¼ 3:914;
woC ðCO2 ; gÞ ¼ 20:11; and woC ðH2 O; gÞ ¼ 9:824 kJ=mol. Upon substitution of
DGorxn and woC;i values in equation (20), the standard molar chemical exergy of
methane gas is found to be 832:67 kJ=mol. The molar chemical exergy of the CH4
ðgÞ þ N2 ðgÞ mixture can now be calculated using equation (18)

b ig
C;mixture ¼ yCH4 wC;CH4 þ yN2 wC;N2 þ RTo ðyCH4 lnyCH4 þ yN2 lnyN2 Þ
o o
w
¼ 0:60ð832:67Þ þ 0:40ð0:664Þ
(22)
þ ð8:314=1000Þð298:15Þ½0:60lnð0:60Þ þ 0:40lnð0:40Þ
¼ 498:20 kJ=mol

Ideal liquid solutions


The chemical potential of any species i in an ideal liquid solution at temperature T0
l id
and pressure P0 , denoted as b i ðT0 ; P0 Þ, is given by

l id
b i ¼ li þ RT0 lnðxi Þ
o
(23)

where loi is the standard chemical potential of liquid species i at T0 (standard state
being pure liquid at pressure P0 ). and xi is the mole fraction of species i. The
chemical exergy of an ideal liquid solution is
X  id 
b id ¼
W li  b
ni b l i;e (24)
C

From equations (23) and (24), it follows that


X  
b id ¼
W ni loi þ RT0 lnðxi Þ
C
X X X
 ni b
l i;e ¼ ni ðloi  b
l i;e Þ þ RT0 ni lnðxi Þ (25)
X X
¼ ni woC;i þ RT0 ni lnðxi Þ
Pal 55

Thus, the molar chemical exergy of an ideal liquid solution is given as


X X
b id ¼ W
w b id =N ¼ xi woC;i þ RT0 xi lnxi (26)
C C

Example: Determine the molar chemical exergy of a binary liquid solution of


methanol and ethanol at 298.15 K and 1 bar. The mole fraction of methanol is
0.67. The solution can be assumed to be ideal.
Solution: As ethanol is not present in the reference atmosphere, we need to set
up the formation reaction of ethanol using the reference species so as to determine
the standard molar chemical exergy of ethanol. The formation reaction for ethanol
from the reference species is

2CO2 ðgÞ þ 3H2 OðgÞ ! CH3 CH2 OHðlÞ þ 3O2 ðgÞ (27)

From equations (5) and (10)


X
woC;ethanol ¼ DGorxn   i woC;i (28)
ðother than
ethanolÞ

DGorxn for the formation reaction of ethanol (equation (27)) can be determined
from the standard Gibbs free energy of formation DGof of various species as

DGorxn ¼ DGof ðCH3 CH2 OH; lÞ þ 3DGof ðO2 ; gÞ  2DGof ðCO2 ; gÞ  3DGof ðH2 O; gÞ
(29)
¼ 174:9 þ 3ð0Þ  2ð394:36Þ  3ð228:589Þ ¼ 1299:587 kJ=mol

The standard molar chemical exergies of the species present in the reference
atmosphere and involved in the reaction are as follows: woC ðO2 ; gÞ ¼ 3:914;
woC ðCO2 ; gÞ ¼ 20:11; and woC ðH2 O; gÞ ¼ 9:824 kJ=mol. Upon substitution of
DGorxn and woC;i values in equation (28), the standard molar chemical exergy of
ethanol is found to be 1416:48 kJ=mol. The molar chemical exergy of metha-
nol–ethanol solution can now be calculated using equation (26)

b id
C;solution ¼ xmethanol wC;methanol þ xethanol wC;ethanol
o o
w
þRT0 ðxmethanol lnxmethanol þ xethanol lnxethanol Þ
¼ 0:67ð719:025Þ þ 0:33ð1416:48Þ (30)

þ ð8:314=1000Þð298:15Þ½0:67lnð0:67Þ þ 0:33lnð0:33Þ
¼ 947:613 kJ=mol
56 International Journal of Mechanical Engineering Education 47(1)

Chemical exergy of non-ideal gas mixtures and liquid solutions


The chemical exergy of a non-ideal system (gas mixture or liquid solution) can be
expressed as
X  
bC ¼
W li  b
ni b l i;e (31)

The change in the chemical potential of any species present in the mixture/
solution is given in terms of the fugacity as

l i ¼ RT0 dlnðfbi Þ
db (32)

where fbi is the fugacity of species i in the mixture/solution. Upon integration,


equation (32) gives
!
fb
b
li ¼ loi þ RT0 ln oi (33)
fi

where foi is the standard fugacity of species i. From equations (31) and (33)
!
bC ¼
X   X fb
W ni loi  b
l i;e þ RT0 ni ln oi (34)
fi

The chemical potential difference in the first summation term on the right-hand
side of equation (34) is the standard molar chemical exergy of species i and there-
fore, equation (34) could be re-written as
!
bC ¼
X X fb
W ni woC;i þ RT0 ni ln oi (35)
fi

This is a very general expression for the total chemical exergy of a system,
applicable to both non-ideal gas mixtures and non-ideal liquid solutions. The
expression is, of course, applicable to ideal systems as well.

Non-ideal gas mixtures


b of any species is defined as18
For non-ideal gas mixtures, the fugacity coefficient / i

b
b ¼ fi
/ (36)
i
yi P
Pal 57

b into equation (35) and recognizing that P ¼ P0 and


Upon substitution of / i
foi ¼ P0 , we obtain
X X  
bC ¼
W ni woC;i þ RT0 b
ni ln yi / (37)
i

Thus, the molar chemical exergy of a non-ideal gas mixture is


X X  
b ¼
w yi woC;i þ RT0 b
yi ln yi / (38)
C i

Fugacity coefficient of pure non-ideal gas. For a single-component (pure) non-ideal gas,
equation (38) reduces to

wC ¼ woC þ RT0 lnð/Þ (39)

where / is the fugacity coefficient of the pure gas. The first term on the right-hand
side of equation (39), that is, the standard molar chemical exergy of gas, treats gas
to be an ideal gas and the second term accounts for the non-ideal behavior. As the
fugacity coefficient / is a measure of non-ideality of gas, it is related to the com-
pressibility factor Z defined as

Z ¼ PV=RT (40)

The fugacity coefficient is related to compressibility factor as


Z P
dP
lnð/Þ ¼ ðZ  1Þ (41)
0 P
For ideal gas, Z ¼ 1 and / ¼ 1. In order to evaluate /, we need an expression
for the compressibility factor and evaluate the integral on the right-hand side of
equation (41). The two-term virial expansion of compressibility factor in P is

BP
Z¼1þ (42)
RT

where B=RT is the second virial coefficient of Z expansion. This is the simplest
form of virial expansion with only first-order correction in Z. As pressure involved
in chemical exergy calculations is 1 bar, this simple truncated two-term virial
expansion is good enough for the purpose of exergy calculation of gases. Upon
substitution of the expression for Z in equation (41) and integrating, we obtain the
following expression for the fugacity coefficient of pure gas

BP
lnð/Þ ¼ (43)
RT
58 International Journal of Mechanical Engineering Education 47(1)

According to Smith et al.,18 the virial constant B can be expressed as follows



RTc  o 
B¼ B þ xB1 (44)
Pc

where Tc and Pc are critical temperature and pressure, respectively, and x is the
acentric factor of the material. Bo and B1 are functions of reduced temperature Tr
ð¼ T=Tc Þ only and are given as

0:422
Bo ¼ 0:083  (45)
T1:6
r

0:172
B1 ¼ 0:139  (46)
T4:2
r

Thus, the fugacity coefficient of pure non-ideal gas, and hence its molar chem-
ical exergy (see equation (39)), can be calculated from equation (43) in conjunction
with equations (44) to (46).
Example: Determine the fugacity coefficient of propane gas at T0 ¼ 298:15 K
and P0 ¼ 1 bar. The critical properties of propane are: Tc ¼ 369:8 K,
Pc ¼ 42:48 bar, and x ¼ 0:152.
Solution: Tr ¼ T0 =Tc ¼ 298:15=369:8 ¼ 0:8062; Pr ¼ P0 =Pc ¼ 1=42:48 ¼ 0:0235;
from equation (45), Bo ¼ 0:5126; from equation (46), B1 ¼ 0:286; from equa-
tion (44), B ¼ 402:457 cm3 =mol; from equation (43), / ¼ 0:9839.

Fugacity coefficients of non-ideal gas mixture. For non-ideal multicomponent gas mix-
tures, the fugacity coefficient expression given in the form of equation (43) can be
generalized to18
" #
P 1 XX  
b ¼
ln/ Bkk þ yi yj 2dik  dij (47)
k
RT 2 i j

where i and j are dummy indices which run over all species and

dik ¼ 2Bik  Bii  Bkk (48)

The parameter d is zero when its indices are the same, that is, d11 ¼ d22 ¼ 0; etc:
Furthermore, d is symmetric meaning d12 ¼ d21 . When the indices of the virial
coefficient B are the same, the virial coefficient is a pure species coefficient. For
example, B11 is a virial coefficient of pure species “1”. The values of pure species
Pal 59

virial coefficients B11 ; B22 ; etc: can be calculated from equation (44). For cross
coefficients Bik ; Bij ; etc:, the following expression is to be used

RTcij  o 
Bij ¼ B þ xij B1 (49)
Pcij

where

xi þ xj
xij ¼ (50)
2
 1=2
Tcij ¼ Tci Tcj (51)

Zci þ Zcj
Zcij ¼ (52)
2

1=3 1=3
!3
Vci þ Vcj
Vcij ¼ (53)
2

Zcij RTcij
Pcij ¼ (54)
Vcij

For binary non-ideal gas mixtures, equation (47) gives the following expressions
for the fugacity coefficients

b ¼ P  
ln/ 1 B11 þ y22 d12 (55)
RT

b ¼ P  
ln/ 2 B22 þ y21 d12 (56)
RT

where

d12 ¼ 2B12  B11  B22 (57)

Thus, the fugacity coefficients of different components of a non-ideal multi-


component gas mixture, and hence the molar chemical exergy of the mixture (see
equation (38)), can be calculated from equation (47) in conjunction with equations
(48) to (54). Note that the calculations require the critical properties
(Tc ; Pc ; Vc ; and Zc ) and acentric factors (x) of the pure components which are
readily available in thermodynamics textbooks.18
60 International Journal of Mechanical Engineering Education 47(1)

Example: Estimate the fugacity coefficients / b and / b for an equimolar binary


1 2
mixture of species 1 and 2 at T0 ¼ 298:15 K and P0 ¼ 1 bar. The critical properties
of the pure species are given in Table 2.
Solution: The critical properties of the binary mixture calculated from equations
(50) to (54) are summarized in Table 3.
Using the data given in Table 3, the values of Trij , Bo , B1 , and Bij are calculated
at T0 ¼ 298:15 K and are summarized in Table 4.
d12 ¼ 2B12  B11  B22 ¼ 2ð2006:654Þ  ð1715:395Þ  ð2330:588Þ ¼ 32:675
cm3 =mol: Thus, the fugacity coefficients are as follows

b ¼ P   1  
ln/ 1 B11 þ y22 d12 ¼ 1715:395 þ 0:52 ð32:675Þ ¼ 0:06887
RT 83:14ð298:15Þ
(58)

b ¼ P   1  
ln/ 2 B22 þ y21 d12 ¼ 2330:588 þ 0:52 ð32:675Þ ¼ 0:09369
RT 83:14ð298:15Þ
(59)

Table 2. Critical properties of pure species 1 and 2.

Species Tc ðKÞ Pc ðbarÞ Vc ðcm3 =molÞ Zc x

1 535.5 41.5 267 0.249 0.323


2 591.8 41.1 316 0.264 0.262

Table 3. Critical properties of binary mixture of species 1 and 2.

ij Tcij ðKÞ Pcij ðbarÞ Vcij ðcm3 =molÞ Zcij xij

11 535.5 41.5 267 0.249 0.323


22 591.8 41.1 316 0.264 0.262
12 563.0 41.3 291 0.256 0.293

Table 4. The values of Trij , Bo , B1 , and Bij at T0 ¼ 298:15 K.

ij Trij Bo B1 Bij ðcm3 =molÞ

11 0.5568 0.994 1.873 1715.395


22 0.5038 1.1809 2.9233 2330.588
12 0.5296 1.0838 2.3438 2006.654
Pal 61

From equations (58) and (59), /b ¼ 0:9334 and / b ¼ 0:9106. Clearly, the gas
1 2
mixture under consideration is non-ideal as the fugacity coefficients are significant-
ly off from unity.

Non-ideal liquid solutions


The fugacity of any species in an ideal liquid solution is given by the Lewis–
Randall rule18 as
id
fbi ¼ xi fi l (60)
id
where fbi is the fugacity of species i in an ideal solution, xi is the mole fraction of
species i in the solution, and fil is the fugacity of pure liquid species i at the solution
temperature and pressure. The activity coefficient ci of any species i in a liquid
solution is defined as

fbi
ci ¼ (61)
xi fli
For an ideal liquid solution, the Lewis–Randall leads to cid i ¼ 1. From equations
(35) and (61), it follows that
X X  l
b xi fi ci
WC ¼ ni wC;i þ RT0
o
ni ln (62)
foi

Since the pressure and temperature in chemical exergy calculations are the ref-
erence atmosphere pressure P0 ¼ 1 bar and temperature T0 ¼ 298:15 K

fio ¼ fil (63)

Thus, equation (62) reduces to the following expression for molar chemical
exergy of non-ideal liquid solutions
X X
b ¼
w xi woC;i þ RT0 xi lnðxi ci Þ (64)
C

According to equation (64), we need the knowledge of activity coefficients in


order to calculate the molar chemical exergy of non-ideal liquid solutions.
Some of the commonly used empirical models for the activity coefficients of
non-ideal binary liquid solutions18 are: Margules model, Van Laar model, and
Wilson model. They are given as follows:
Margules model

lnc1 ¼ x22 ½A12 þ 2ðA21  A12 Þx1  (65)


62 International Journal of Mechanical Engineering Education 47(1)

lnc2 ¼ x21 ½A21 þ 2ðA12  A21 Þx2  (66)

where A12 and A21 are the constants for the binary system; they depend on tem-
perature for a given binary system.
Van Laar model

2
A0 12 x1
lnc1 ¼ A0 12 1 þ (67)
A0 21 x2

2
A0 21 x2
lnc2 ¼ A0 21 1 þ (68)
A0 12 x1

where A0 12 and A0 21 are the constants for the binary system; they depend on tem-
perature for a given binary system.
Wilson model

X12 X21
lnc1 ¼ lnðx1 þ x2 X12 Þ þ x2  (69)
x1 þ x2 X12 x2 þ x1 X21

X12 X21
lnc2 ¼ lnðx2 þ x1 X21 Þ  x1  (70)
x1 þ x2 X12 x2 þ x1 X21

where X12 and X21 are the constants for the binary system; they depend on tem-
perature for a given binary system. The temperature-dependence of X12 and X21 is
further expressed as

V2  a 
12
X12 ¼ exp  (71)
V1 RT

V1  a 
21
X21 ¼ exp  (72)
V2 RT

where a12 and a21 are the constants independent of temperature. They depend only
on the chemical nature of the binary system. V1 and V2 are the molar volumes of
species 1 and 2, respectively. The Wilson model has the additional advantage that
it can easily be generalized to multicomponent systems.18
Example: Estimate and plot the activity coefficients c1 and c2 for binary solutions of
methanol (1) and water (2) for different values of x1 . Also determine the molar chem-
ical exergy of methanol (1)–water (2) solution with x1 ¼ 0:10 at 298.15 K and 1 bar.
Use the Wilson model to estimate the activity coefficients. Given that a12 ¼ 107.38 cal/
mol, a21 ¼ 469.55 cal/mol, V1 ¼ 40.73 cm3/mol and V2=18.07 cm3/mol.
Pal 63

Solution: The values of X12 and X21 are calculated from equations (71) and (72)
as X12 ¼ 0:3701 and X21 ¼ 1:0203. Using these values of X12 and X21 , the activity
coefficients are calculated from equations (69) and (70) and are plotted in Figure 4.
Clearly, methanol–water solutions are non-ideal as the activity coefficients deviate
from a value of unity.
The molar chemical exergy of methanol–water solution can be estimated from
equation (64), that is
b ¼ x1 wo þ x2 wo þ RTo ½x1 lnðx1 c Þ þ x2 lnðx2 c Þ
w (73)
C C;1 C;2 1 2

The values of various quantities present in equation (73) are as follows:


x1 ¼ 0:10, x2 ¼ 0:90, c1 ¼ 1:9927, c2 ¼ 1:0144, woC;1 ¼ 719:025 kJ=mol, T0 ¼ 298:15 K,
R ¼ 8:314 J=mol K. The only unknown quantity on the right-hand side of equation
(73) is the standard molar exergy of liquid water woC;2 . To estimate woC;2 (standard
molar chemical exergy of liquid water), we need to set up the formation reaction
for liquid water using the reference species as
H2 OðgÞ ! H2 OðlÞ (74)
For this reaction, DGorxn ¼ 8:589 kJ=mol. The standard molar chemical exergy
of liquid water can now be calculated as shown below

woC ðH2 O; lÞ ¼ DGorxn þ woC ðH2 O; gÞ ¼ 8:589 þ 9:824 ¼ 1:235 kJ=mol (75)

Figure 4. Activity coefficients of methanol and water in methanol–water solutions.


64 International Journal of Mechanical Engineering Education 47(1)

Upon substitution of the values of various quantities on the right-hand side of


equation (73), the molar chemical exergy of methanol–water solution turns out to
be 72:411 kJ=mol:

Practical application of chemical exergy and exergy analysis


Consider a practical steam power plant where steam is generated in a furnace/
boiler unit by burning methane completely to CO2 ðgÞ and H2 OðgÞ with 25% excess
air. Methane and air enter the furnace at 25 C and 1 bar and the flue gases leave
the furnace at 186.85 C and 1 bar. The molar flow rate of methane is 457.85 mol/s
and the mass flow rate of steam in the steam cycle is 105.932 kg/s. Figure 5 shows
the furnace/boiler unit of the steam power plant. Calculate the following: (a)
exergy supplied by the furnace; (b) exergy recovered by the boiler; (c) exergy
destroyed in the furnace/boiler unit; and (d) the exergetic efficiency of the fur-
nace/boiler unit. The reference environment is as shown in Figure 2.

Compositions of the furnace streams


The reaction occurring in the furnace is

CH4 ðgÞ þ 2O2 ðgÞ ! CO2 ðgÞ þ 2H2 OðgÞ (76)

On the basis of 1 mol of methane burned with 25% excess air, the compositions
of the various furnace streams can be calculated as follows: (a) air entering the
furnace: O2 ðgÞ ¼ 2ð1:25Þ ¼ 2:5 mol; N2 ðgÞ ¼ 2:5ð79=21Þ ¼ 9:405 mol; total air
entering per mole of methane burnt ¼ 11.905 mol. Thus, the composition of air
is yO2 ¼ 0:21 and yN2 ¼ 0:79 ; (b) flue gases leaving the furnace: O2 ðgÞ ¼ 0:5 mol;

Figure 5. Furnace/boiler unit of a steam power plant.


Pal 65

N2 ðgÞ ¼ 9:405 mol; CO2 ðgÞ ¼ 1:0 mol; H2 OðgÞ ¼ 2:0 mol; total moles of flue gases
per mole of methane burnt ¼ 12.905 mol. Thus, the composition of flue gases is
yO2 ¼ 0:0387; yN2 ¼ 0:7288 ; yCO2 ¼ 0:0775 ;.yH2 O ¼ 0:1550.

Chemical exergies of the furnace streams


The chemical exergies of the furnace streams are determined on the basis of 1 mol
of methane burnt in the furnace.

Methane entering the furnace. The chemical exergy of methane is calculated as


follows

wC;CH4 ¼ woC;CH4 þ RT0 ln/CH4


¼ 832:67 þ ð8:314=1000Þð298:15Þlnð0:9983Þ ¼ 832:66 kJ=mol
(77)

The fugacity coefficient of methane gas is calculated using equation (43).

Air stream entering the furnace. The chemical exergy of air is calculated as follows
n    o
b
w ¼ y N wo
þ yO w o
þ RT0 yN ln y N
b
/ þ y O ln yO
b
/
C;Air 2 C;N2 2 C;O2 2 2 N2 2 2 O2

(78)

b ¼ 0:9998,
yN2 ¼ 0:79, yO2 ¼ 0:21, woC;N2 ¼ 0:664 kJ=mol, woC;O2 ¼ 3:914 kJ=mol, / N2
b ¼ 0:9994, T0 ¼ 298:15 K, R ¼ 8:314 J=mol K. Thus, w b
/ O2 C;Air ¼ 0:07179 kJ=mol.
The fugacity coefficients of a binary gas mixture were calculated using equations
(55) and (56). As the number of moles of air entering per mole of methane burnt in
b
the furnace is 11.905, w C;Air ¼ 11:905ð0:07179Þ ¼ 0:85466 kJ=mol: Note that hat is
now removed from the symbol w b
C;Air for the sake of convenience.

Flue gases leaving the furnace. The chemical exergy of the flue gas consisting of four
components is calculated as follows

b
C;flue gas ¼ y1 wC;1 þ y2 wC;2 þ y3 wC;3 þ y4 wC;4
o o o o
w
h        i (79)
þ RT0 y1 ln y1 / b þ y2 ln y2 / b þ y3 ln y3 /
b þ y4 ln y4 /
b
1 2 3 4

The fugacity coefficients of a quaternary gas mixture were calculated using


equation (47). They are nearly unity indicating that the flue gas is nearly ideal.
66 International Journal of Mechanical Engineering Education 47(1)

Table 5. Values of various quantities present in equation (79).

Standard molar
chemical exergy Fugacity coefficient
Species Composition, yi (woC;i ) (kJ/mol) b)
(/ i

1¼ O2(g) 0.0387 3.914 1.0002


2¼ N2(g) 0.7288 0.664 1.0004
3¼ CO2(g) 0.0775 20.11 0.9992
4¼ H2O(g) 0.1550 9.824 0.9971

Upon substitution of the quantities given in Table 5 into equation (79), we obtain
the following value of the chemical exergy of flue gas (per mole of flue gas):
b
w C;flue gas ¼ 1:625 kJ=mol: As the number of moles of flue gas leaving the furnace
b
per mole of methane burnt is 12.905, w C;flue gas ¼ 12:905ð1:625Þ ¼ 20:9706 kJ=mol:

Thermo-mechanical exergy of the flue gases


The thermo-mechanical exergy associated with the air and methane streams enter-
ing the furnace is zero. However, the flue gases leaving the furnace are at a tem-
perature of 186.85 C. Thus, they possess thermo-mechanical energy in addition to
chemical exergy by virtue of having a temperature much higher than the reference
environment temperature of 25 C. The thermo-mechanical energy can be
expressed as

wTM;fluegas ¼ ðh  ho Þ  To ðs  so Þ (80)

As the flue gas is nearly ideal, equation (80) gives



T
wTM;flue gas ¼ Cpm ðT  T0 Þ  T0 Cpm ln
T0

460 (81)
¼ 31:1ð460  298:15Þ  298:15ð31:1Þln
298:15
¼ 1012:72 J=mol

where Cpm, the mean heat capacity of the flue gas, is 31.1 J/mol.K. As the
number of moles of flue gases is 12.905 per mole of methane burnt, the total
thermo-mechanical exergy of flue gases is wTM;flue gas ¼
12:905ð1:0127Þ ¼ 13:07 kJ=mol:
Pal 67

Solution to the problem


a. The exergy supplied by the furnace per mole of methane burnt is as follows
 
wfurnace ¼ wC;methane þ wC;air  wC;flue gas þ wTM;flue gas
(82)
¼ 832:66 þ 0:85466  ð20:9706 þ 13:07Þ ¼ 799:474 kJ=mol

This is the exergy supplied by the furnace per mole of methane burnt. The molar
flow rate of methane is 457.85 mol/s. Thus, the total rate of exergy supplied in the
furnace is W_ furnace ¼ 457:85ð799:474Þ ¼ 366:039:2 kW.

b. The exergy recovered in the boiler is equal to the increase in exergy of the boiler
fluid stream (see Figure 5) given as

_ boiler ¼ m_ ½ðh5  h4 Þ  T0 ðs5  s4 Þ


W
¼ ð105:932Þ½ð3391:6  203:4Þ  ð298:15Þð6:6858  0:658Þ (83)
¼ 147; 352:62 kW

c. The exergy destroyed in the furnace/boiler unit is as follows

_ destroyed ¼ 366039:2  147352:62 ¼ 218; 686:58 kW


W (84)

d. The exergetic efficiency of the furnace/boiler unit is the ratio of exergy recovered
in the boiler to exergy supplied in the furnace

W_ boiler 147352:62
e¼ ¼ ¼ 0:4026 (85)
_ furnace
W 366039:2

Implementation of the proposed concepts and analysis


The concepts and analysis described here are suitable for third-year engineering
students who have completed an introductory course on engineering thermodynam-
ics in their second year where they are taught the laws of thermodynamics and
related fundamental concepts such as work, heat, internal energy, enthalpy, entropy,
closed and open systems, cyclic and non-cyclic processes. The material of this article
should be introduced after the students have covered phase and chemical equilibria
of ideal and non-ideal systems (vapor mixtures and liquid solutions) in their
advanced level course in thermodynamics. This material could be covered in a
one-week period (three lectures of 50 min each and one tutorial of 50 min) provided
that the students are familiar with the concepts of phase and chemical equilibria.
68 International Journal of Mechanical Engineering Education 47(1)

Table 6. True/false type questions for the assessment of intended learning outcomes.

(i) Exergy is also referred to as “Availability”. True/False


(ii) Exergy is a thermodynamic property of the system alone. True/False
(iii) Exergy is the maximum possible useful work that can be extracted from a system till
it reaches the state of thermodynamic equilibrium with its environment.
True/False
(iv) For a system to be in thermodynamic equilibrium with its environment, the only
requirements are that the temperatures and pressures of the system and its envi-
ronment are the same. Tue/False
(v) The “restricted dead state” of a system is one where the system is in mechanical and
thermal equilibrium with its environment. True/False
(vi) A system is said to be in “complete dead state” when it is in thermodynamic equi-
librium with its environment. True/False
(vii) The chemical compositions of a system in its current state and in the restricted dead
state are the same. True/False
(viii) The thermo-mechanical exergy of a system is defined as the maximum possible useful
work that can be extracted from the system till it reaches the restricted dead state.
True/False
(ix) The chemical exergy of a system is defined as the maximum possible useful work that
can be extracted from the system when it goes from the “restricted” dead state to
the “complete” dead state. True/False
(x) As exergy is a property of the system relative to its reference environment, it is
essential to define the “reference environment” in order to do any exergy calcu-
lations. True/False
The fugacity of any species in an ideal liquid solution (b
(xi) id
f ) is given by the following
i

Lewis–Randall rule: bf i ¼ xi fi l where xi is the mole fraction of species i in the


id

solution and fi l is the fugacity of pure liquid species i . True/False


(xii) b ig
The fugacity of any species i in an ideal gas mixture ðf Þ is equal to the partial pressure
i
of species i : b
ig
f i ¼ yi P where yi is the mole fraction of species i in the gas mixture and
P is the total pressure. True/False
(xiii) The activity coefficient of any species in a non-ideal liquid solution is defined as: ci ¼ bf i
=bf i ¼ bf i =xi fi l where bf i is the fugacity of species i in the actual solution and b
id id
f i is the
fugacity of species i in the solution if the solution was ideal. True/False
(xiv) The fugacity coefficient of any species in a non-ideal gas mixture is defined as: u b i ¼ bf i
=bf i ¼ b
f i =yi P where bf i is the fugacity of species i in the actual gas mixture and b
ig ig
f i is
the fugacity of species i in the gas mixture if the gas mixture was ideal.
True/False
(xv) The liquid solutions are generally ideal in nature. True/False
Pal 69

Table 7. Calculation type problems for the assessment of intended learning outcomes.

1. Problem statement:
Determine the standard molar chemical exergy of liquid octane C8H18(l) at T0 ¼ 298:15
K with respect to the reference environment specified in Figure 2.
Solution:
As C8H18(l) is not a part of the reference atmosphere, we need to set up a chemical
reaction to produce liquid octane from the existing species in the reference environ-
ment. The formation reaction of C8H18(l) is as follows

8CO2 ðgÞ þ 9H2 OðgÞ ! C8 H18 ðlÞ þ ð25=2ÞO2 ðgÞ

The standard Gibbs free energies of formation of various compounds in units of kJ/mol
are as follows: DGfo ðCO2 ; gÞ ¼ 394:36; DGfo ðH2 O; gÞ ¼ 228:589; DGfo ðO2 ; gÞ ¼ 0;
DGfo ðC8 H18 ; lÞ ¼ 6:4. The standard Gibbs free energy of reaction, DGrxn
o
, for the for-
mation reaction of liquid octane can be determined from the standard Gibbs free
energy of formation DGfo of various species as
DGrxn
o
¼ DGfo ðC8 H18 ; lÞ þ ð25=2ÞDGfo ðO2 ; gÞ  8DGfo ðCO2 ; gÞ  9DGfo ðH2 O; gÞ
¼ 6:4 þ ð25=2Þð0Þ  8ð394:36Þ  9ð228:589Þ ¼ 5218:581 kJ=mol
The standard molar chemical exergies of the species present in the environment are:
woC ðO2 ; gÞ ¼ 3:914; woC ðCO2 ; gÞ ¼ 20:11; and woC ðH2 O; gÞ ¼ 9:824 kJ=mol.
Upon substitution of DGrxn o
and woC;i values in the following equation
X
woC ðC8 H18 ; lÞ ¼ DGrxn
o
  i woC;i
ðother than
C8 H18 Þ
the standard molar chemical exergy of liquid octane is found to be 5418:952 kJ=mol.
2. Problem statement:
(a) Calculate the molar chemical exergy of an equimolar non-ideal gas mixture of species
1 and 2 at 298.15 K and 1 bar. The virial coefficients are given as follows:
B11 ¼ 1500 cm3 =mol; B22 ¼ 2500 cm3 =mol; B12 ¼ 1900 cm3 =mol and the
parameter d12 ¼ 2B12  B11  B22 ¼ 200 cm3 =mol: The standard molar chemical
exergies of species 1 and 2 are given as: woC;1 ¼ 0:5 kJ=mol; woC;2 ¼ 6:5 kJ=mol.
(b) What would be the molar chemical exergy of the gas mixture, if the mixture is ideal?
Solution:
(a) We first need to calculate
 the fugacity coefficients. They are as follows
b ¼ P B11 þ y2 d12 ¼
ð83:14Þð298:15Þ ð1500 þ 0:5  200Þ ¼ 0:0585
1 2
ln / 1 RT  2 
b
ln/ 2 ¼ RT B22 þ y1 d12 ¼ ð83:14Þð298:15Þ ð2500 þ 0:5  200Þ ¼ 0:09884
P 2 1 2

Thus, / b ¼ 0:9432 and / b ¼ 0:9059, and the molar chemical exergy of the gas mixture
1 2
is calculated as shown below  
X X
b ¼
w yi wo þ RT0 b ¼ 0:5ð0:5Þ þ 0:5ð6:5Þ þ ð8:314  298:15=1000Þ
yi ln yi /
C C;i i

 ½0:5lnð0:5  0:9432Þ þ 0:5lnð0:5  0:9059Þ ¼ 1:587 kJ=mol


(continued)
70 International Journal of Mechanical Engineering Education 47(1)

Table 7. Continued
(b) If the gas mixture is ideal, the fugacity coefficients are unity and the molar chemical
exergy becomes
X X
b ¼
w yi wo þ RT0 yi lnðyi Þ ¼ 0:5ð0:5Þ þ 0:5ð6:5Þ þ ð8:314  298:15=1000Þ
C C;i

 ½0:5lnð0:5Þ þ 0:5lnð0:5Þ ¼ 1:782 kJ=mol


3. Problem statement:
Determine the molar chemical exergy of binary non-ideal liquid solutions of 2-propanol
(species 1) and water (species 2) at T0 ¼ 298:15 K and 1 bar with the following
compositions: (a) x1 ¼ 0:05; (b) x1 ¼ 0:5; and (c) x1 ¼ 0:95. Use Wilson model to
estimate the activity coefficients. Given that a12 ¼437.98 cal/mol, a21=1238 cal/mol,
V1 ¼ 76.92 cm3/mol, and V2=18.07 cm3/mol. The reference environment is specified in
Figure 2.
Solution:
The molar chemical exergy of a non-ideal liquid solution is given as
X X
b ¼
w xi woC;i þ RT0 xi lnðxi ci Þ
C

We need the standard molar chemical exergies of individual species woC;i and the activity
coefficients ci .
The standard molar chemical exergy of liquid water was calculated earlier (see equation
(75)) to be 1.235 kJ/mol. As 2-propanol is not a part of the reference atmosphere, we
need to set up a chemical reaction to produce liquid 2-propanol from the existing
species in the reference environment. The formation reaction of 2-propanol, that is,
CH3CHOHCH3(l), is as follows
3CO2 ðgÞ þ 4H2 OðgÞ ! CH3 CHOHCH3 ðlÞ þ ð9=2ÞO2 ðgÞ
The standard Gibbs free energies of formation of various compounds in units of kJ/mol
are as follows: DGfo ðCO2 ; gÞ ¼ 394:36; DGfo ðH2 O; gÞ ¼ 228:589; DGfo ðO2 ; gÞ ¼ 0;
DGfo ðC3 H8 O; lÞ ¼ 180:3. The standard Gibbs free energy of reaction, DGrxn
o
, for the
formation reaction of liquid 2-propanol can be determined from the standard Gibbs
free energy of formation DGfo of various species as
DGrxn
o
¼ DGfo ðC3 H8 O; lÞ þ ð9=2ÞDGfo ðO2 ; gÞ  3DGfo ðCO2 ; gÞ  4DGfo ðH2 O; gÞ
¼ 180:3 þ ð9=2Þð0Þ  3ð394:36Þ  4ð228:589Þ ¼ 1917:136 kJ=mol
The standard molar chemical exergies of the species present in the environment are:
woC ðO2 ; gÞ ¼ 3:914; woC ðCO2 ; gÞ ¼ 20:11; and woC ðH2 O; gÞ ¼ 9:824 kJ=mol.
Upon substitution of DGrxn o
and woC;i values in the following equation
X
woC ðC3 H8 O; lÞ ¼ DGrxn
o
  i woC;i
ðother than
C3 H8 OÞ
the standard molar chemical exergy of liquid 2-propanol is found to be
1999:149 kJ=mol.
Now that the standard molar chemical exergies of the species are known, we need to
calculate the activity coefficients. The values of X12 and X21 are calculated from
(continued)
Pal 71

Table 7. Continued
equations (71) and (72) as: X12 ¼ 0:1122 and X21 ¼ 0:5267. Using these values of X12
and X21 , the activity coefficients are calculated from Wilson equations (69) and (70) for
the given compositions. Once the standard molar chemical exergies and the activity
coefficients of different species are known, the molar chemical exergy of a non-ideal
liquid solution is calculated from the equation given at the beginning of this solution.
The results are summarized below:
Composition γ 1 (2-propanol) γ 2 (water) ψˆC (molar chemical exergy) in
kJ/mol.
x1=0.05 7.557 1.015 100.925
x1=0.50 1.409 1.672 999.536
x1=0.95 1.003 4.035 1898.94

Assessment of student learning outcomes


The assessment of the intended student learning outcomes stated in the
Introduction section can be achieved using two types of tests: true/false questions
based test and calculations based test. Table 6 lists 15 true/false type questions with
correct answers in bold face, and Table 7 gives three calculations based problems
with complete solutions.

Conclusions
A comprehensive treatment of the chemical exergy of materials is presented.
Starting from the fundamental concepts, the equations describing the chemical
exergy of materials are presented. Special attention is given to non-ideal gas mix-
tures and liquid solutions. Numerical examples are included where necessary to
illustrate the concepts for the benefit of the students. The concepts and the equa-
tions developed in the article are applied for the solution of a practical problem
dealing with furnace–boiler unit of a practical steam power plant. Assessment
problems involving true/false type and calculations type questions and their sol-
utions are included in order to assess the intended student learning outcomes of
this article.

Declaration of conflicting interests


The author(s) declared no potential conflicts of interest with respect to the research, author-
ship, and/or publication of this article.

Funding
The author(s) received no financial support for the research, authorship, and/or publication
of this article.
72 International Journal of Mechanical Engineering Education 47(1)

References
1. Pal R. Demystification of the Gouy–Stodola theorem of thermodynamics for closed
systems. Int J Mech Eng Educ 2017; 45: 142–153.
2. Pal R. On the Gouy-Stodola theorem of thermodynamics for open systems. Int J Mech
Eng Educ 2017; 45: 194–206.
3. Pal R. Quantification of irreversibilities in practical cyclic processes using exergy anal-
ysis and Gouy-Stodola theorem. Int J Mech Eng Educ 2017; published online, DOI:
10.1177/0306419017720427.
4. Kotas TJ. The exergy method of thermal plant analysis. London: Exergon Publishing
Company, 2012.
5. Bejan A, Tsatsaronis G and Moran M. Thermal design & optimization. New York: John
Wiley & Sons, 1996.
6. Brockway P, Dewulf J, Kjelstrup S, et al. In a resource-constrained world: think exergy,
not energy. Brussels: Science Europe, 2016.
7. Cengel YA and Boles MA. Thermodynamics – an engineering approach. 8th ed. New
York: McGraw-Hill, 2015.
8. Moran MJ, Shapiro HN, Boettner DD, et al. Fundamentals of engineering thermody-
namics. 9th ed. New York: John Wiley & Sons, 2018.
9. Borgnakke C and Sonntag RE. Fundamentals of thermodynamics. 8th ed. New York:
John Wiley & Sons, 2013.
10. Nag PK. Engineering thermodynamics. 5th ed. New Delhi: McGraw-Hill, 2013.
11. Rajput RK. Engineering thermodynamics. 3rd ed. Sudbury: Jones and Bartlett
Publishers, 2010.
12. Potter MC and Somerton CW. Thermodynamics for engineers. 3rd ed. New York:
McGraw-Hill, 2014.
13. Dincer I and Rosen MA. Exergy. Amsterdam: Elsevier, 2012.
14. Martinez I. Chemical exergy, http://webserver.dmt.upm.es/isidoro/dat1/Chemical%
20exergy.pdf (accessed 20 August 2017).
15. Szargut J. Chemical exergies of the elements. Appl Energy 1989; 32: 269–286.
16. Rivero R and Garfias M. Standard chemical exergy of elements updates. Energy 2006;
31: 3310–3326.
17. Castellan GW. Physical chemistry. Menlo Park: The Benjamin/Cummings Publishing
Company, 1983.
18. Smith JM, Van Ness HC and Abbott MM. Introduction to chemical engineering ther-
modynamics. New York: McGraw-Hill, 2004.

You might also like