You are on page 1of 7

EXERGY FLOWS ANALYSIS IN CHEMICAL REACTORS

M. SORIN*, J. LAMBERT* and J. PARIS*


*Energy Diversication Research Laboratory, CANMET, Varennes, Quebec, Canada
Chemical Engineering Department, Ecole Polytechnique, Montreal, Quebec, Canada
A
new thermodynamic criterion of chemical reactor performance, the utilizable exergy
coefcient is proposed. This criterion can be used for simultaneous assessment of
internal and external exergy losses and of the conversion performance of chemical
reactors. An exergy analysis of the steam methane reforming process has been performed using
alternatively the conventional exergy efciency and the utilizable exergy coefcient. It is
shown that the exergy efciency does not account for the conversion effect resulting from by
the process improvements, while the utilizable exergy coefcient does.
Keywords: exergy analysis; reactor design; steam methane reforming
INTRODUCTION
A new method for exergy analysis of multi-step processes
known as exergy load distribution analysis was rst
proposed by Sorin and Brodyansky
1
. It is based on an
analytical expression of the relationship between the exergy
efciencies of the individual operations constituting a
process and its overall efciency. The graphical representa-
tion of this expression is a convenient tool to identify design
and operating changes which enhance the overall process
efciency. The implementationof the method to monitor the
stepwise improvement of a typical hydrogen production unit
has been recently demonstrated by Sorin and Paris
2
. It has
also been shown that the results of application of the method
depend on the expression of the exergy efciency of the
individual operations. For example the application of the
traditional exergy efciencies as dened by Fratzcher
3
for
chemical reactors promotes solutions which improve energy
efciency of chemical processes. However, energy ef-
ciency performance is not necessarily the most important
factor in chemical processes. Process yield is of primary
concern (Smith
4
). One of the important factors which
inuences the yield is conversion performance in chemical
reactors. Therefore, it is important to formulate a criteria
that characterizes simultaneously the thermodynamic and
conversion performance of chemical reactors.
Second law efciency analysis has been already success-
fully applied for the thermodynamic performance evaluation
of chemical reactors
5
,
6
. The relationship between the special
efciency coefcient, intrinsic exergy efciency, and the
conversion of a single exothermic reaction system has been
recently established
7
for the case of constant inlet exergy.
This paper presents the graphical representation of exergy
ows enteringand transformed within an endothermic system
with multiple reactions by use of the Grassman diagram
8
. The
values of exergy inlet and outlet change simultaneously. The
exergy ows which traverse the reactor without transforma-
tion are readily visualized on the diagram and their relation
with the conversionfactor is explained. A newcoefcient, the
utilizable exergy coefcient, is proposed for the simultaneous
assessment of internal and external exergy losses and the
characterization of the conversion performance of chemical
reactors. The application of the conventional exergy
efciency coefcient and of the utilizable exergy coefcient
to the analysis and improvement of a steam methane reformer
for hydrogen productionis also presented as an example. The
difference between the results of the analysis according to
both coefcients is explained.
BACKGROUND
To any material, heat and work stream can be associated
an exergy content which is completely dened by
temperature, pressure and composition of the stream itself
and of a reference state which normally is the environment
in which the system operates
9
. It has been shown by Szargut
et al.
10
,
11
that the exergy content of a material stream is the
amount of work which would be produced by bringing this
stream in thermal, mechanical and chemical equilibrium
with the reference state by a sequence of reversible
operations. This process can be achieved in two steps. In
the rst step, thermal and mechanical equilibrium is
accomplished (i.e. the end temperature and pressure of the
stream are the same as those of the reference state) through
reversible thermal exchanges only; the corresponding part
of the exergy content of the stream is its thermo-mechanical
exergy. It is composed of two interdependent components,
the mechanical and thermal exergies. The thermal exergy is
due to the temperature difference between the stream and
the reference state. The mechanical exergy is due to the
pressure difference. In the second step, the stream is brought
into chemical equilibrium with the reference state by
reversible heat and mass exchange operations effected at
the reference temperature and pressure; the corresponding
part of the exergy content of the stream is its chemical
exergy. Both thermo-mechanical and chemical exergies are
uniquely dened. Brodyansky
12
has shown that the exergy
content of a heat stream can be dened as a function of
temperature conditions which can be above or below the
reference state. The exergy content of a work stream is equal
to its energy and a reference state is not required. The
389
02638762/98/$10.00+0.00
q Institution of Chemical Engineers
Trans IChemE, Vol 76, Part A, March 1998
classication of all exergy forms taken from Sorin et al.
13
is reproduced in Table 1. It is important to notice that to
any single stream may correspond one or more forms of
exergy.
It is therefore possible to compute the exergy contents of
all in-coming and out-going streams to and from a system
and to establish an overall exergy balance over any system
(Figure 1). The total exergy input, E
9
, of a real system is
always higher than its exergy output, E
99
, because a certain
amount of exergy is irreversibly destroyed within the
system. This exergy generally referred to as the internal
exergy losses, D
i n t
, is directly linked to the thermodynamic
irreversibilities in the system. Therefore the total exergy
balance satises the relationship:
E99
=
E9
-
D
int
(1)
The second law efciency of the process as was rst
proposed by Grassman
8
is,

=
E
99
E
9
(2)
As illustrated in Figure 1, part of the exergy output from the
system may be dissipated into the environment as, for
example, heat losses, sewered wastes or smokestack
efuents. This wasted exergy, no longer utilizable by
subsequent processes constitutes the external losses, D
e x t
. It
is more appropriate, from the standpoint of downstream
operations, to consider the exergy that remains utilizable,
E
u
, rather than the total output, E
99
, and, accordingly,
equation (1) can be rewritten as:
E
u
=
E9
-
D
int
-
D
ext
(3)
The exergy efciency as dened by Fratzcher
3
is:
g
e =
E
u
E
9
=
1
-
D
int
+
D
ext
E
9
(4)
In the past, exergy analysis has essentially been based on
those concepts, i.e. the estimation of the internal and
external exergy loses and the expressions of the second law
and exergy efciencies.
It has been observed that can be deceptively inated,
and that , in particular, it may assume a value close to one,
for operations which, from an engineering point of view,
have a poor performance. For example, a chemical reactor
with very low conversion rate or a heat exchanger with very
small heat duty would produce such an effect
13
. The reason
is the fact that only part of the utilizable exergy is produced
by the system in the accomplishment of all the physico-
chemical phenomena which take place within its bound-
aries. The rest of the exergy that leaves the system with the
utilizable exergy stream is a part of the exergy input which
has simply traversed the system without undergoing any
transformation (Figure 1). This fundamental fact was rst
recognized by Kostenko
14
who gave the name transiting
exergy, E
t r
, to this fraction of the exergy supplied to a
system. Typically, in a chemical reactor part (but not all,
because of temperature and pressure changes) of the exergy
associated with unreacted feed or inerts would constitute
transiting exergy. Transiting exergy was further character-
ized by Sorin and Brodyansky
15
and Brodyansky et al.
16
who have developed algorithms for its direct computation.
The algorithms are given in the Appendix. Therefore only
part of the exergy input is consumed by the system in order
to produce new forms of utilizable exergy. On the basis of
these observations, Sorin and Brodyansky
15
have dened a
390 SORIN et al.
Trans IChemE, Vol 76, Part A, March 1998
Table 1. Forms and components of exergy.
Associated Stream Material Heat Work
Forms of exergy Chemical Thermo-Mech. (T-M) Conduct. Radiative Work
Components of T-M Exergy Mechanical (E
p
) Thermal (E
T
)
Figure 1. Graphical presentation of the exergy balance.
new coefcient of thermodynamic efciency, later named
by Sorin et al.
17
the intrinsic exergy efciency:
g
i =
E
p
E
c =
E
99
-
E
tr
E
9
-
E
tr
(5)
The terms E
c
, E
p
and E
t r
are the consumed, produced and
total transiting exergies respectively (Figure 1). Intrinsic
exergy efciency is the measure of the true ability of the
system to produce new exergy from a given amount of
consumed exergy. Hence its name. However, g
i
does not
account for the fact that, because of the external exergy
losses, D
e x t
, which are determined by factors exterior to the
system itself, all of the exergy produced, E
p
, is no longer
utilizable.
In this work, an alternative exergy coefcient which is
more pertinent to the evaluation of practical systems
performance, the utilizable exergy coefcient, g
u
, is
introduced. It is dened as:
g
u =
E
pu
E
c =
E
99
-
D
ext
-
E
tr
,
u
E
9
-
E
tr
,
u
(6)
E
p u
is the produced utilizable exergy; it constitutes part of
E
p
(Figure 1). It is important to emphasize that the value
E
t r , u
is only the part of the total transiting exergy which is
included in the utilizable exergy stream. As illustrated in
Figure 1 there may also be transiting exergy, E
t r , D
, in the
external exergy losses stream, D
e x t
. For example, exergy of
the part of the initial feed traversing the system without
transformation and lost into the environment. To compute
g
u
according to equation (6), there is no need to evaluate the
term E
t r , D
. It should also be mentioned that the utilizable
exergy coefcient is not equivalent to the rational efciency
proposed earlier by Kotas
18
.
Application of the utilizable exergy coefcient, g
u
, and of
the exergy efciency, g
e
, to the analysis and improvement of
a steam methane reforming process will be illustrated. The
results of the analysis on the basis of both coefcients will
be compared.
EXERGY ANALYSIS OF A STEAM METHANE
REFORMING PROCESS
The Process
The process is essentially based on two main reactions
(Cromatry, 1993
19
), the steam reforming of the methane:
CH
4 +
H
2
O
P
CO
+
3H
2
(R1)
and the water-gas shift reaction:
CO
+
H
2
O
P
CO
2 +
H
2
(R2)
Reaction (R1) is strongly endothermic and thermodynami-
cally favoured by high temperature, low pressure and excess
steam. Reaction (R2) is slightly exothermic and thermo-
dynamically favoured by low temperature and excess steam.
In the reforming furnace (Figure 2) conversion of the
methane according to equations (R1) and (R2) takes place in
catalyst lled tubes disposed vertically in a chamber where
the heat needed for the reaction is produced by combustion
of fuel (natural gas in the case treated here). The outlet
synthesis gas is composed of H
2
, CO, CO
2
, CH
4
and steam.
Its temperature is 800 to 9008C and pressure 20,000 to
30,000kPa. Usually the combustion air is mixed with fuel
before entering the reforming furnace.
The application of oxygen-enriched combustion to steam
methane reforming is a potential improvement of the
process which has been examined in a previous study by
Lambert et al.
20
. The enrichment of air by oxygen can be
achieved through membrane separation. This step is
included in the process scheme presented in Figure 2. The
391 EXERGY FLOWS ANALYSIS IN CHEMICAL REACTORS
Trans IChemE, Vol 76, Part A, March 1998
Figure 2. Scheme and nomenclature of the steam methane reforming furnace.
separation consumes electrical power as indicated in the
gure. Oxygen enrichment of air has positive effects on the
heat production step:

It reduces the fraction of parasitic nitrogen in the


combustion gases, thus decreasing the amount of sensible
heat escaping in the ue gases and increasing the heat
available to the process.

It improves heat transfers in the reformer because of


higher ame temperature and higher emissivity of
combustion gases due to the increase in CO
2
and H
2
O
concentrations. The improvement of heat transfer through
the tubes of a steam reforming furnace increases the
amount of heat received by the methane and steam
mixture, and hence its temperature, so that the reform-
ing equilibrium is displaced towards higher hydrogen
production.
These effects may be utilized to enhance the performance
of a reforming unit according to specic objectives. For
example, hydrogen production may be increased by
increasing the level of air enrichment by oxygen at constant
feedstock and fuel consumption. The results of the steam
methane reforming process computations for different
levels of air enrichment taken from Lambert et al.
20
are
presented in Table 2. This shows that application of the
oxygen-enriched combustion leads on the one hand to
hydrogen and carbon monoxide production increase and
simultaneous rise of the efuent temperature, T
99
, but , on the
other hand, to an increase in electrical power consumption
by the membrane separation unit.
Exergy Forms Consumed and Produced by the Process
In this section, the expressions for variations of the
exergy forms and components presented in Table 1 are
developed. Each of these expressions presents exergy which
is consumed or produced for further utilization by the
reforming furnace and constitute the net thermodynamic
effect.
An exergy ow diagram for the process presented in
Figure 2 is given in Figure 3. The symbols E
1
, E
2
E
6
in
Figure 3 correspond to the exergy ows at the points
1,2

6 in Figure 2. Figure 3 illustrates qualitatively the


fact that the chemical exergies of the product,

E
99
H2
, and
byproducts,

E
99
CO
and

E
99
CO2
, are produced from a fraction of
the reactant exergy, E
9
nCH4
. It is assumed that the exergy
of the ue gas, E
5
, is rejected and constitutes the external
losses D
e x t
. Following Szargut
11
, the exergy of air is taken
as zero and the chemical exergy of water is considered as
negligible
16
.
The overall exergy balance of the process is:
E
1 +
E
2 +
E
3 +
E
4 =
E
6 +
D
ext
+
D
int
(7a)
or
(E9
x
)
1 +
(E9
p
,
T
)
1 +
(E9
p
,
T
)
2 +
E
3 +
E
4
=
(E99
x
)
6 +
(E99
p
,
T
)
6 +
D
ext
+
D
int
(7b)
Then according to equation (4) the exergy efciency of the
reactor is:
g
e =
E
6
E
1 +
E
2 +
E
3 +
E
4
(8)
According to equation (A2), the transiting chemical exergy
in the process streams 1 and 6 (Figure 3a) can be written:
E
tr
x =
(n
CH4
)
6
( e
CH4
)
6
(9)
According to equation (A3), the transiting thermomechani-
cal exergy is the sum of the transiting exergies in streams
1,2 and 6, i.e.,
E
tr
p
,
T =
(E
tr
p
,
T
)
1
-
6 +
(E
tr
p
,
T
)
2
-
6,
(10)
where
(E
tr
p
,
T
)
1
-
6 =
m
1
e
6
(p
6,
T
6
)
,
(11)
and
(E
tr
p
,
T
)
2
-
6 =
m
2
e
2
(p
6,
T
2
)
.
(12)
Figure 3 presents also the components of the transiting
exergy (gray bands). By subtraction of E
tr
x
(equation(9))
from the term ( E
9
x
)
1
at the left side of equation (7b) the
chemical exergies of the methane consumed by the process
can be calculated as:
E
nCH
4 =
(n
9
CH4 -
n99
CH4
)e9
CH4 +
n99
CH4
(e9
CH4 -
e99
CH4
) (13)
The term E
n C H 4
is made up of two parts, the chemical
exergy of the reacted methane, (n
9
i -
n
99 )e
9
i
(i
=
CH
4
), and
the reduction of chemical exergy of the non reacted methane
due to dilution by the products of reaction, n
99
i
(e
9
i -
e
99
i
).
Similarly, the substraction of E
tr
x
from the right side of
equation (7b) removes from this term the exergy of the non
reacted methane and enables computation of the chemical
exergies of H
2
, CO and CO
2
produced as:

E99
H2 =
n99
H2
e99
H2
(14)

E99
CO =
n99
CO
e99
CO
(15)

E99
CO2 =
n99
CO2
e99
CO2
16)
The substraction of E
tr
p
,
T
(equation (10)) from the terms
(E
9
p
,
T
)
1
, (E
9
p
,
T
)
2
and (E
99
p
,
T
)
6
of equation(7b) gives respectively:
(=E
p
,
T
)
1 =
m
1
[e
1
(p
1,
T
1
)
-
e
6
(p
6,
T
6
)
]
(17)
(=E
p
)
2 =
m
2
[e
2
(p
2,
T
2
)
-
e
2
(p
6,
T
2
)
]
(18)
(D
E
T
)
6 =
m
2
[e
6
(p
6,
T
6
)
-
e
6
(p
6,
T
2
)
]
(19)
The term (=E
p , T
)
1
is the reduction in thermo-mechanical
exergy of the process stream 1 caused by its pressure and
392 SORIN et al.
Trans IChemE, Vol 76, Part A, March 1998
Table 2. Operating data as function of the oxygen enrichment level.
Enrichment level (%O
2
) 21 29 30 31 32 33 34 35
Synthesis gas temperature at tubes outlet (8C) 815 862 867 871 875 879 882 884
H
2
ow rate in synthesis gas (kmol h
-
1
) 192.1 204.4 205.5 206.3 207.0 207.7 208.4 208.7
CO ow rate in synthesis gas (kmol h
-
1
) 30.2 38.0 38.7 39.2 39.8 40.3 40.8 41.0
Electrical power (MJ h
-
1
) 0.0 1.13 1.23 1.34 1.45 1.56 1.69 1.83
temperature drops as well as by changes in thermodynamic
properties such as, heat capacity, specic volume etc

,
resulting from the variations in composition between feed
and efuent conditions. Similarly, (=E
p
)
2
is the isothermal
decrease in mechanical exergy of stream 2 computed at the
inlet temperature T
2
(the lowest temperature in the system)
and caused by pressure drop and the variations in the stream
thermodynamic properties. Finally, (
D
E
T
)
6
is the isobaric
increase in thermal exergy caused by the temperature rise of
the process stream 6 and changes in its thermodynamic
properties.
The terms computed according to equations (14), (15),
(16) and (17) are the forms of exergy produced by the
reforming unit. The terms computed according to equations
(13), (17) and (18) as well as the exergies of the fuel, E
4
, and
the electrical power, E
3
, are the forms of exergy consumed
by the process. The terms are visualized in Figure 3. The
utilizable exergy coefcient g
u
is then:
g
u =

E
99
H2 +

E
99
CO +

E
99
CO2 +
(D
E
T
)
6
E
nCH4 +
E
3 +
E
4 +
(=E
p
)
2 +
(=E
p
,
T
)
1
(20)
Table 3 presents a summary of the exergy forms
consumed and produced by the process as well as their
numerical values for the base case (i.e. without oxygen-
enriched combustion).
It shouldbe notedthat the values

E
99
CO2
, (=E
p ,T
)
1
, (=E
p
)
2
are
small and can be neglected, which simplies expression (20):
g
u =

E99
H2 +

E99
CO +
(D
E
T
)
6
E
nCH4 +
E
3 +
E
4
(21)
Table 3 presents also the difference between g
u
and g
e
. It is
explained by the fact that the transiting exergy E
t r
due
mostly to the methane traversing the reactor, constitutes
14% of the overall exergy input.
Comparison of h
u
and h
e
as a Function of Oxygen
Enrichment
To compare the exergy efciency g
e
and intended
efciency g
u
, the effect of oxygen enrichment level is
now considered. The pertinent operating data are given in
Table 2. Figure 4 presents the evolution of the exergy
coefcients as a function of the oxygen enrichment level.
The exergy efciency g
e
decreases with enrichment, an
effect linked to the fact that the ratio (D
i n t
+
D
e x t
)/E
9
,
393 EXERGY FLOWS ANALYSIS IN CHEMICAL REACTORS
Trans IChemE, Vol 76, Part A, March 1998
Figure 3. Exergy ow diagram.
Table 3. Exergy forms consumed and produced for the further utilization by
the process.
Consumed Exergy Produced and Utilizable Exergy
Symbol Value, MJ h
-
1
Symbol Value, MJ h
-
1
E
n CH4
46256

E99 H2
44838
(=E
p,T
)
1
103

E99
CO
8127
(=E
p
)
2
5

E
99 CO2
346
E
3
0 (DE
T
)
6
3317
E
4
27243
E
c
85975 E
p
68676
E
tr
11989 E
tr
11989
g
e
79.88%
g
u
76.62%
increases with the conversion resulting from oxygen-
enriched combustion (Table 4). The increase in this ratio,
causes, g
e
, to decrease (equation (4)). On the basis of the
exergy efciency criteria it would seem preferable to let the
whole feedstock ow through the reactor. There would be
no exergy losses because there is neither chemical
transformation nor fuel combustion. An opposite situation
is observed when applying criteria g
u
. As illustrated in
Table 4, the increase in conversion results in a decrease of
the transiting exergy, E
t r
, and a substantial rise in exergy
production, E
p
, due to the increase in the terms

E
99
H2
,

E
99
CO
and (
D
E
T
)
6
in the numerator of equation (21). As a result,
the intended efciency g
u
increases with the enrichment
level attaining a maximum value at 3334% O
2
in the air
(Figure 4). The following decrease in g
u
with the enrichment
level is explained by the rise in exergy consumption, E
c
(Table 4) due to higher demand in electrical power for air
separation. As illustrated in Table 4, a reason for the
divergence between the evolution trends of g
e
and g
u
is that
the exergy efciency, g
e
, is primarily linked to internal
exergy losses per unit of outlet exergy (D
i n t
+
D
e x t
)/E
99
,
while the intended efciency, g
u
, depends more upon the
exergy losses per unit of produced exergy, (D
i n t
+
D
e x t
)/E
p
.
The ratio (D
i n t
+
D
e x t
)/E
99
increases monotonically with
conversion. The ratio (D
i n t
+
D
e x t
)/E
p
rst decreases then
reaches a minimum at the 90.0% conversion level and
nally increases. It is apparent that there is no quantitative
correspondence between the exergy efciency, g
e
, and the
impact of an oxygen-enriched combustion. On the contrary,
the intended efciency, g
u
, shows that there is an optimum
operation, from the thermodynamic standpoint and that it
represents a balance between power consumption for air
enrichment and chemical conversion. This coefcient
appears to be a more appropriate index to assess the process
performance.
CONCLUSION
A new coefcient, the utilizable exergy coefcient, has
been proposed for the thermodynamic assessment of
chemical reactors. The procedure for the computation of
this coefcient is based on the nonambiguousexpressions of
the unaltered part of exergy (transiting exergy) and external
exergy losses which are excluded from the analysis. The list
of the forms of exergy produced and consumed by a
chemical reactor is established. These forms are linked to
changes in conversion rate, temperature, pressure and
species concentration variations in feedstock and product
ows. Using the example of the steam methane reforming
process, it has been shown that the conventional exergy
efciency coefcient decreases with the conversion pro-
voked by the process improvement. Meanwhile, the
utilizable exergy coefcient has a smooth maximum
which establishes the thermodynamic compromise between
energy consumption and conversion rate. It is concluded
that the utilizable exergy coefcient is a more suitable
criterion of the thermodynamic performance of such a
chemical system. The main reason for the divergence
between the two criteria is that the exergy efciency
promotes solutions which minimize the exergy losses per
unit of outlet exergy, while the intended efciency promotes
solutions which minimize the exergy losses per unit of
produced exergy.
APPENDIX
Computation of Transiting Exergy Associated with
Material Streams
A1.Transiting Chemical Exergy
It has been shown by Brodyansky et al.
16
that the
chemical exergy of a stream containing different compo-
nents, is given, at the temperature and pressure of its
environment, T
o
and p
o
, by:
E
x =
i
n
i
(e
x
,
i +
RT
0
ln c
i
x
i
)
=
i
n
i
e
x
,
i
(A1)
394 SORIN et al.
Trans IChemE, Vol 76, Part A, March 1998
Figure 4. Inuence of the oxygen enrichment level on the coefcients g
e
and g
u
Table 4. Exergy terms as functions of the oxygen enrichment level or methane conversion.
Enrichment level (%O
2
) 21 29 30 31 32 33 34 35
CH
4
conversion 0.805 0.878 0.884 0.889 0.894 0.899 0.903 0.904
E, MJ h
-
1
85975 87074 87172 87275 87382 87496 87623 87766
E99 , MJ h
-
1
68674 69433 69497 69553 69603 69650 69695 69707
E
tr
, MJ h
-
1
11987 7815 7452 7168 6901 6653 6420 6313
E
c
, MJ h
-
1
73988 78128 78488 78770 79035 79281 79513 79619
E
p
, MJ h
-
1
56687 61618 62045 62385 62702 62997 63275 63394
(D
int
+
D
ext
), MJ h
-
1
17301 17641 17675 17722 17779 17846 17928 18059
(D
int
+
D
ext
)/E9, % 20.12 20.26 20.27 20.31 20.35 20.40 20.46 20.58
(D
int
+
D
ext
)/E
99
, % 25.19 25.41 25.43 25.48 25.54 25.62 25.72 25.91
(D
int
+
D
ext
)/E
p
, % 30.52 28.63 28.49 28.41 28.35 28.33 28.33 28.49
with:
n
i
molar ow rate of component i,
x
i
molar fraction of component i in the mixture,
c
i
activity coefcient of component i in the mixture,
e
x , i
molar chemical exergy of pure component i,
e
x
,
i
molar chemical exergy of component i in the mixture.
The transiting chemical exergy ow rate through a system
entered and exited by any number of streams is:
E
tr
x =
i
n
tr
i .
min e
9
x
,
i ,
e99
x
,
i
(A2)
The molar ow rate n
tr
i
is the rate of through molar ow of
component i, i.e. the minimum value of the net in and out
molar ow rates of component i (i.e. ow rates summed
over all in and out streams respectively).
A2.Transiting Thermo-Mechanical Exergy
For the input and output conditions (T
9$ T
o
; T
99$ T
o
) the
transiting thermo-mechanical exergy can be computed as:
E
tr
p
,
T =
j
m
tr
j
min e
9
p
,
T
(p9
,
T
min
); e9
p
,
T
(p99
,
T
min
);
e99
p
,
T
(p9
,
T
min
); e99
p
,
T
(p99
,
T
min
)) (A3)
The mass ow rate m
tr
i
is the rate of mass ow of stream j
traversing a system.
NOMENCLATURE
D exergy loss due to irreversibilities
e molar chemical exergy of a pure component
e molar chemical exergy of a component in a mixture
E exergy ow rate of a pure component

E exergy ow rate of a component in a mixture


n molar ow rate
m mass ow rate
p pressure
T temperature
x molar fraction
Greek Characters
second law efciency
g
i
intrinsic efciency
g
e
exergy efciency
g
u
utilizable exergy coefcient
c activity coefcient
D produced exergy
l chemical potential
Symbols
= expenditure of exergy
Subscripts
e exergy
i chemical species in a mixture
p mechanical exergy
p,T thermomechanical exergy
T thermal exergy
x chemical exergy
0 environment
Superscripts
99 process output
9
process input
c consumed exergy
ext external exergy losses
int internal exergy losses
min minimal value
p produced exergy
pu produced utilizable exergy
tr transiting exergy
u utilizable exergy
REFERENCES
1. Sorin, M. V. and Brodyansky, V. M., 1992, Thermodynamic
optimisation and integration of processes using a general formula for
the efciency, I. The theory and application to an ammonia synthesis
plant. Energy: The Int J, 11: 10191031.
2. Sorin, M. and Paris, J., 1997, Combined exergy and pinch approach to
process analysis, Comput Chem Eng, 21(suppl): S23S28.
3. Fratzcher, W., 1961, Exergetical efciency, B.W.K., 13(11): p. 483.
4. Smith, R., 1995, Chemical Process Design (McGraw-Hill, Inc, NY).
5. Denbigh, K. G., 1956, The second-law efciency of chemical
processes, Chem Eng Sci, 6: 19.
6. Riekert, L., 1974, The efciency of energy-utilization in chemical
processes, Chem Eng Sci, 29: 16131620.
7. Sorin, M., Bonhivers, J-C. and Paris, J., 1997, Exergy efciency and
conversion of chemical reactions, Proc Florence World Energy
Research Symp, Flowers 94, pp. 941949.
8. Grassmann, P., 1950, Zur allgemeinen denition des wirkungsgrades,
Chem Ing Technik, 4: 7780.
9. Rant, Z., 1956, Exergie, ein neues wort fur `technische arbeitsfahig-
keit, Forsch Gebiete Ingenieurwes.
10. Szargut, J. and Petela, R., 1965, Egzergia (in Polish) (Wydawnictwa
Naukowo-Techniczne, Warsaw).
11. Szargut, J., Morris, D. and Stewart, F., 1988, Exergy Analysis of
Thermal, Chemical and Metallurgical Processes (Hemisphere Publ
Co, NY).
12. Brodyansky, V. M., 1973, Exergy Method of Thermodynamic Analysis
(in Russian) (Energiya, Moscow).
13. Sorin, M., Brodyansky, V. M. and Paris, J., 1994, Observations on
exergy efciency coefcients, Proc Florence World Energy Research
Symp, Flowers 94, pp. 941949.
14. Kostenko, G., 1983, Thermodynamic assessment of heat processes
efciency, Promishlenaya Teplotechnika (in Russian), (4): 7073.
15. Sorin, M. V and Brodyansky, V. M., 1985, A method of efciency
denition for power and chemical systems, (in Russian), University
News of the USSR-Energetics, N3: pp. 7888.
16. Brodyansky, V. M., Sorin, M. and LeGoff, P., 1994, The Efciency of
Industrial Processes: Exergy Analysis and Optimization (Elsevier
Science B.V., Amsterdam).
17. Sorin, M., Sapundzhiev, C. and Paris, J., 1994, Application of a new
exergy analysis method to a ow reversal SO
2
convertor, Chem Eng
Sci, 49(24B): 56035613.
18. Kotas, T. J., 1985, The Exergy Method of Thermal Plant Analysis
(Butterworths, London).
19. Cromatry, B., 1993, Modern aspects of steam reforming for hydrogen
plants, Hydrogen Energy Progress, IX(1): 1322.
20. Lambert, J., Sorin, M. and Paris, J., 1997, Analysis of oxygen-enriched
combustionfor steammethane reforming, Energy: Int J, 22(8): 817825.
ADDRESS
Correspondence concerning this paper should be addressed to Dr M.
Sorin, Energy Diversication Research Laboratory, CANMET, 1615
Lionel-Boulet, PO Box 4800, Varennes, Quebec, J3X 1S6, Canada.
(E-mail: mikhael.sorin@nrcan.gc.ca).
The manuscript was received 17 July 1997 and accepted for publication
after revision 4 December 1997.
395 EXERGY FLOWS ANALYSIS IN CHEMICAL REACTORS
Trans IChemE, Vol 76, Part A, March 1998

You might also like