Sie sind auf Seite 1von 12

Energy Policy 35 (2007) 39673978

Understanding energy and exergy efciencies for improved energy


management in power plants
Mehmet Kanoglu
a,1
, Ibrahim Dincer
b,
, Marc A. Rosen
b
a
Department of Mechanical Engineering, University of Gaziantep, 27310 Gaziantep, Turkey
b
Faculty of Engineering and Applied Science, University of Ontario Institute of Technology, 2000 Simcoe Street North, Oshawa, Ont., Canada, L1H 74K
Received 19 December 2006; accepted 21 January 2007
Available online 26 March 2007
Abstract
An extensive overview is provided of various energy- and exergy-based efciencies used in the analysis of power cycles. Vapor and gas
power cycles, cogeneration cycles and geothermal power cycles are examined, and consideration is given to different cycle designs. The
many approaches that can be used to dene efciencies are provided and their implications discussed. Improvements of the management
of energy in power plants that stem from understanding the efciencies better are described. Examples are given to illustrate the
efciencies and their differences, with the results presented using combined energy and exergy diagrams. It is anticipated that the results
will provide a convenient and practical tool for engineers and researchers dealing with the analysis, design, optimization and
improvement of power cycles.
r 2007 Elsevier Ltd. All rights reserved.
Keywords: Efciency; Energy; Exergy
1. Introduction
Efciency is one of the most frequently used terms in
thermodynamics, and it indicates how well an energy
conversion or process is accomplished. Efciency is also
one of the most frequently misused terms in thermody-
namics and is often a source of misunderstanding. This is
because efciency is often used without being properly
dened rst (Cengel and Boles, 2006). For an engineering
system, efciency, in general, can be dened as the ratio of
desired output to input. Although this denition provides a
simple general understanding of efciency, a variety of
specic efciency relations for different engineering sys-
tems and operations have been developed. Some research-
ers have recognized some of the difculties related to
denitions of efciencies for energy systems (e.g., Struchtr-
up and Rosen, 2002; Rosen et al., 2005; Kanoglu, 2001;
Bisio and Rubatto, 2001; Zaleta et al., 2001; Brookes, 2004;
Cornelissen et al., 1995).
Efciency is sometimes used as a synonym for perfor-
mance. This is another source of confusion since perfor-
mance and efciency are different and sometimes pose
contradictory design objectives. For example, designing a
more powerful engine for an automobile usually requires a
sacrice in fuel efciency, dened as the rate of fuel
consumed per unit power produced. Note that in the
preceding sentence, we are careful to specify the denition
of fuel efciency used; alternative denitions include the
amount of fuel consumed per kilometer traveled (in
Europe, Canada and many other countries) or miles
traveled per gallon of fuel (in the US).
Efciency traditionally has been primarily dened based
on the rst law (i.e., energy). In recent decades, exergy
analysis has found increasingly widespread acceptance as a
useful tool in the design, assessment, optimization and
improvement of energy systems (e.g., Bejan, 2006; Kotas,
1995; Szargut et al., 1988). Determining exergy efciencies
for an overall system and/or the individual components
making up the system constitutes a major part of exergy
analysis. A comprehensive analysis of a thermodynamic
ARTICLE IN PRESS
www.elsevier.com/locate/enpol
0301-4215/$ - see front matter r 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.enpol.2007.01.015

Corresponding author.
E-mail addresses: mehmet.kanoglu@uoit.ca (M. Kanoglu),
ibrahim.dincer@uoit.ca (I. Dincer), marc.rosen@uoit.ca (M.A. Rosen).
1
On leave at University of Ontario Institute of Technology (UOIT) as a
visiting professor.
system includes both energy and exergy analyses in order to
obtain a more complete picture of system behavior.
To assist in improving the efciencies of power plants,
their thermodynamic characteristics and performances are
usually investigated. Power plants are normally examined
using energy analysis but, as pointed out previously, a
better understanding is attained when a more complete
thermodynamic view is taken, which uses the second law of
thermodynamics in conjunction with energy analysis via
exergy methods.
Although exergy analysis can be generally applied to
energy and other systems, it appears to be a more powerful
tool than energy analysis for power cycles because of the
fact that it helps determine the true magnitudes of losses
and their causes and locations, and improve the overall
system and its components. In this article, we provide an
overview of various energy- and exergy-based efciencies
used in the analysis of power cycles, including vapor and
gas power, cogeneration and geothermal power plants.
Differences in design aspects are considered. The various
approaches that can be used in dening efciencies are
identied and their implications discussed. In addition,
improvements in the management of energy in power
plants that stem from improved understanding of efcien-
cies are described. Numerical examples are provided to
illustrate the use of the different efciencies, and the results
include combined energy and exergy diagrams.
Note that the emphasis in this article is to describe
various energy- and exergy-based efciencies used in
power plants and discuss the implications associated
with each denition. Therefore, simple cycles are selected
to keep the complexity of the plants at a minimum level
for gas and vapor cycles to better facilitate under-
standing of the efciencies, which can be very useful for
improved energy management in power plants. One can
easily adapt the efciencies discussed here to more complex
power systems. Some efciency denitions for gas cycles
found in many thermodynamics textbooks are repeated so
that the coverage in this article is comprehensive and can
serve as a convenient and practical tool for engineers and
researchers.
2. Efciency and energy management and policy
In the analysis of an energy conversion system, it is
important to understand the difference between energy and
exergy efciencies. By considering both of these efciencies,
the quality and quantity of the energy used to achieve a
given objective is considered and the degree to which
efcient and effective use of energy resources is achieved
can be understood (Dincer et al., 2004). Improving
efciencies of energy systems is an important challenge
for meeting energy policy objectives. Reductions in energy
use can assist in attaining energy security objectives. Also,
efcient energy utilization and the introduction of renew-
able energy technologies can signicantly help solve
environmental issues. Increased energy efciency benets
the environment by avoiding energy use and the corre-
sponding resource consumption and pollution generation.
From an economic as well as an environmental perspective,
improved energy efciency has great potential.
Accelerated gains in efciency in energy production and
use, particularly in the power generation and utility sectors,
can help reduce environmental impact and promote energy
security. While there is a large technical potential for
increased efciency, there exist signicant social and
economic barriers to its achievement. Priority should be
given to energy policies and strategies that will yield
efciency gains. However, reliance on such policies alone is
unlikely to overcome these barriers. For this reason,
innovative and bold approaches are required by govern-
ment, in cooperation with decision makers in the power
generation industry, to realize the opportunities for
efciency improvements, and to accelerate the deployment
of new and more efcient technologies.
An engineer designing a system is often expected to aim
for the highest reasonable technical efciency at the lowest
cost under the prevailing technical, economic and legal
conditions, and with regard to ethical, ecological and social
consequences. Exergy methods can assist in such activities,
and offer unique insights into possible improvements.
Exergy analysis is a useful tool for addressing the
environmental impact of energy resource utilization, and
ARTICLE IN PRESS
Nomenclature
c specic heat, kJ/kg K
ex specic ow exergy, kJ/kg
_
E energy rate, kW
_
Ex exergy rate, kW
h specic enthalpy, kJ/kg
k specic heat ratio
_ m mass ow rate, kg/s
q specic heat transfer, kJ/kg
_
Q heat transfer rate, kW
r compression ratio
r
c
cutoff ratio
r
p
pressure ratio
T temperature, K
T
0
ambient temperature, K
T
s
source temperature, K
DT
pp
pinch-point temperature difference, K
_
W power, kW
Greek letters
Z
th
thermal efciency
Z
ex
exergy efciency
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3968
for furthering the goal of more efcient energy-resource
use, for it enables the locations, types and true magnitudes
of losses to be determined. Also, exergy analysis reveals
whether or not and by how much it is possible to design
more efcient energy systems by reducing inefciencies.
Commodities with high-exergy drive tasks, like the
manufacture of products and provision of services. It is,
in fact, this very characteristic of exergy that we value.
Intuitively, therefore, exergy should be correlated to
economics. For a process or system, one can argue that
costs are better distributed among outputs when cost
accounting is based on exergy because exergy often is a
consistent measure of economic value (Tsatsaronis, 2007;
Tsatsaronis et al., 1993). That is, a large quantity of exergy
is often associated with a valuable commodity. Energy, on
the other hand, is only sometimes a measure of economic
value.
Exergy is also strongly related to sustainability and
environmental impact. Sustainability increases and envir-
onmental impact decreases as the exergy efciency of a
process increases. As exergy efciency approaches 100%,
the environmental impact associated with process opera-
tion approaches zero, since exergy is only converted from
one form to another without loss (either through internal
consumption or waste emissions). Also the process
approaches sustainability since it approaches reversibility.
As exergy efciency approaches 0%, the process deviates as
much as possible from sustainability because exergy-
containing resources (fuel, ores, steam, etc.) are used but
nothing is accomplished. Also, environmental impact
increases markedly because, to provide a xed service, an
ever increasing quantity of resources must be used and a
correspondingly increasing amount of exergy-containing
wastes are emitted (Dincer and Rosen, 2005; Cornelissen,
1997).
The exergy analysis of a complete life cycle is known as
exergetic life cycle assessment and is a useful tool for
quantifying the environmental problems associated with
the depletion of natural resources (Cornelissen and Hirs,
2002; Lombardi, 2003; Consoli, 1993).
Energy and exergy efciencies are considered by many to
be useful for the assessment of energy conversion and other
systems and for efciency improvement. However, the use
of ambiguous efciencies that are not clearly dened does
not serve this purpose well. A clear, correct and effective
use of energy and exergy efciencies is crucial in efciency
improvement efforts, which are often a key objective in
energy management and policy making.
For governments seeking to improve energy and
resource security, by increasing the efciency with which
a society or country uses such resources, exergy provides a
critical perspective. It establishes the limits on what can be
done and identies target areas for efciency improvement
(i.e., those areas with high-exergy losses). Some work has
been done on tracking the exergy ows through regions
and economies (e.g., countries, states, provinces). These
efforts mainly focus on understanding the true efciency of
energy and resource use in these regions and countries,
thereby providing information that is useful to govern-
ments and policy makers.
3. Vapor and gas power cycles
3.1. Efciencies
The thermal efciency, also referred to as the energy
efciency or the rst-law efciency, of a power cycle is
dened as
Z
th1

w
net;out
q
in
1
q
out
q
in
, (1)
where w
net,out
is the specic net work output, q
out
the
specic heat rejected from the cycle and q
in
is the specic
heat input to the cycle, which is usually taken to be the
specic heat input to the steam in the boiler of a steam
power plant. That is,
q
in
h
3
h
2
, (2)
where h denotes specic enthalpy and the subscripts
refer to state points in Fig. 1. This simple approach
neglects the losses occurring in the furnaceboiler system
due to the energy lost with hot exhaust gases, incomplete
combustion, etc. To incorporate these losses, one can
express the thermal efciency of the cycle by a second
approach as
Z
th2

_
W
net;out
_ m
fuel
q
HV
, (3)
ARTICLE IN PRESS
Q
in
Condenser
Turbine
Pump
Boiler
3 2
4 1
Furnace
Fuel
Fig. 1. Simple steam power plant.
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3969
where _ m
fuel
is the mass ow rate of fuel and q
HV
is the
heating value of the fuel, which can be chosen as the higher
or lower heating value. For furnace-boiler systems where
the water in the exhaust gases is not expected to condense,
like in internal combustion engines, it is customary to use
the lower heating value (Pulkrabek, 2004).
Some tend to use lower heating values to make a device
appear more efcient. This is frequently done in manu-
facturer descriptions of commercial boilers. Often a
claimed efciency exceeds 100%. This is due to recovering
some of the heat of condensation of steam in the exhaust
gases while still dening boiler efciency based on lower
heating value. This is misleading and a thermodynamically
improper use of efciency. If there is any possibility of
recovering some of the energy of condensing steam in
exhaust gases, the efciency should be based on the higher
heating value.
The second-law efciency, also referred to as exergy
efciency, of a power producing cycle is dened as
Z
ex

w
net;out
ex
in
1
ex
dest
ex
in
, (4)
where ex
in
is the specic exergy input to the cycle and ex
dest
is the specic total exergy destruction in the cycle. One can
express the exergy input to the cycle as the exergy increase
of the working uid in the boiler of a steam power plant
(Fig. 1) as
ex
in
h
3
h
2
T
0
s
3
s
2
, (5)
where T
0
is the dead-state or environment temperature and
s is the specic entropy. Substituting Eq. (5) into (4)
Z
ex1

w
net;out
h
3
h
2
T
0
s
3
s
2

. (6)
In this denition, the irreversibilities during energy
transfer from the furnace to the steam in the boiler are
not accounted for. Alternatively, the exergy input to the
cycle may be dened as the exergy input to the boiler
accompanying the heat transfer. The exergy efciency in
this case becomes
Z
ex2

w
net;out
q
in
1 T
0
=T
s

, (7)
where T
s
is the source temperature, which is the
temperature of the heat source (i.e., furnace), and q
in
is
given by Eq. (2). This efciency denition incorporates the
irreversibility during heat transfer to the steam in the
boiler. We may also incorporate in the efciency denition
the exergy destruction associated with fuel combustion and
the exergy lost with exhaust gases from the furnace. In this
third approach, the exergy efciency can be expressed as
Z
ex3

_
W
net;out
_ m
fuel
ex
fuel
, (8)
where ex
fuel
is the specic exergy of the fuel. The exergy of
a fuel may be obtained by writing the complete combustion
reaction of the fuel and calculating the reversible work by
assuming all products are at the state of the surroundings.
Then the exergy of fuel is equivalent to the calculated
reversible work. For fuels whose combustion reaction
involves water in the products, the exergy of the fuel is
different depending on the phase of water (vapor or liquid).
The exergies of various fuels listed in Szargut et al. (1988)
are based on the vapor phase of water in combustion gases.
Different efciency denitions are possible if one selects
different system boundaries. Clearly dening the system
boundary allows the efciency to be dened unambigu-
ously. For example, the exergy efciencies in Eqs. (7) and
(8) correspond to systems whose boundaries are given by
the inner and outer dashed lines, respectively, in Fig. 1.
Simplied thermal efciency relations of idealized cycles
for internal combustion engines and gas-turbine cycles are
available. When the Otto cycle is used to represent the
operation of an internal combustion engine, the thermal
efciency under air-standard assumptions is
Z
th;Otto
1
1
r
k1
, (9)
where r is the compression ratio and k is the specic heat
ratio. Similarly, the thermal efciency of the Diesel cycle,
which is the idealized model for compression ignition
engines, is
Z
th;Diesel
1
1
r
k1
r
k
c
1
kr
c
1

, (10)
where r
c
is the cutoff ratio, dened as the ratio of cylinder
volumes after and before the combustion process. The
efciency relation for the dual cycle is
Z
th;Dual
1
1
r
k1
r
p
r
k
c
1
kr
p
r
c
1 r
p
1

, (11)
where r
p
is the ratio of pressures after and before the
constant-volume heat addition process. The thermal
efciency of the simple Brayton cycle, which is the idealized
model for gas-turbine engines, is expressed using the air-
standard assumption as
Z
th;Brayton
1
1
r
k1=k
p
, (12)
where r
p
is the ratio of maximum and minimum pressures
in the cycle. For the idealized regenerative Brayton cycle,
the efciency relation is
Z
th;Brayton;regen
1
T
1
T
3

r
k1=k
p
, (13)
where T
1
and T
3
are the temperatures at the inlets of the
compressor and the turbine, respectively.
The operational description of these idealized cycles may
be found in most thermodynamics textbooks (e.g., Cengel
and Boles, 2006; Sonntag et al., 2002). Eqs. (9)(13) are
only applicable to the idealized cycles considered, and they
should not be used to determine the thermal efciencies of
actual internal combustion engines or gas-turbine cycles.
Eqs. (9)(13) are useful in that they illustrate the effects of
ARTICLE IN PRESS
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3970
some key design parameters such as compression ratio,
cutoff ratio and pressure ratio on cycle efciency.
The thermal efciency of a gas power cycle can be
conveniently calculated using Eq. (3). Eq. (1) can also be
used provided that the actual combustion process is
represented by an equivalent heat addition process and
the exhaust is replaced by an equivalent heat rejection
process. For example, in a closed-cycle gas-turbine engine,
the states in Eq. (2) would correspond to the states of the
working uid before and after the heat addition process. In
the Otto cycle model of a spark-ignition engine, where the
actual combustion is represented by a constant-volume
heat addition process, the heat input is equal to the internal
energy change of the working uid during the heat addition
process.
If we follow a similar approach for calculating the exergy
efciency of a gas power cycle, the resulting relations yield
different exergy efciencies, as is the case for the example
considered earlier of a steam power plant. When calculat-
ing the exergy efciency of an Otto cycle using Eq. (6), the
term in the denominator must be replaced by the exergy
increase of the working uid during a constant-volume
heat addition process in a closed system. Eq. (6) can be
applied directly to the Diesel cycle, while this relation
should be modied to account for constant-volume and
constant-pressure heat addition processes in the Dual cycle.
Eqs. (7) and (8) can be conveniently used to determine the
exergy efciency of a gas power cycle. Eq. (8) incorporates
the exergy destruction associated with combustion of the
fuel and uses the chemical exergy of the fuel as the input
while the exergy of the heat released after combustion is the
input in Eq. (7).
3.2. Illustrative example
A numerical example is used to illustrate and contrast
the various efciencies dened in this section. We consider
a simple steam power plant with a net power output of
10 MW and boiler and condenser pressures of 10,000 and
10 kPa, respectively (Fig. 1). We assume a turbine inlet
temperature of 500 1C and isentropic efciencies of 85%
for both the turbine and the pump. In addition, we assume
that the furnaceboiler system has an efciency of 75%.
That is, 75% of the lower heating value of the fuel is
transferred to the steam owing through the boiler while
the remaining 25% is lost, mostly with the hot exhaust
gases passing through the chimney. The source and sink
temperatures in Eq. (7) are taken as 1300 and 298 K,
respectively. We consider methane as the fuel with a lower
heating value of 50,050 kJ/kg and a chemical exergy of
51,840 kJ/kg (Szargut et al., 1988).
For the given values and assumptions, an analysis of this
cycle yields
w
net;out
1081 kJ=kg; q
in
3172 kJ=kg,
ex
in1
1400 kJ=kg; ex
in2
2444 kJ=kg;
as well as the following efciency values:
Z
th1
34:1%; Z
th2
25:6%; Z
ex1
77:2%,
Z
ex2
44:2%; Z
ex3
24:7%.
When the energy and exergy losses in the furnace-boiler
system are not considered, the thermal efciency is 34.1%
while the corresponding exergy efciency is much higher
(77.2%). However, when the losses in furnace-boiler are
considered, the exergy efciency (24.7%) is lower than the
thermal efciency (25.6%). When teaching undergraduate
thermodynamics, it is normally stated that the exergy
efciency is greater than the thermal efciency for heat
engines, referring to the rst approach here. This point is
made by emphasizing that thermal efciency is the fraction
of heat input that is converted to work while exergy
efciency is the fraction of the work potential of the heat
(this work potential, i.e., exergy, is smaller than heat) that
is converted to work. However, when one considers the
effect of furnace-boiler losses, and uses the chemical exergy
of the fuel in the exergy efciency and the heating value of
the fuel in the thermal efciency, the exergy efciency
becomes smaller than the thermal efciency. In thermo-
dynamics, it is often misleading to make generalized
statements as they may not always apply. For example,
can we state that the exergy efciency, based on the third
approach in Eq. (8) (Z
ex3
), is always lower than the
thermal efciency as dened by the second approach in
Eq. (3) (Z
th2
)? The answer will be yes only if the chemical
exergy of the fuel is always greater than its heating value.
According to data in Szargut et al. (1988), this is the case
for methane but not for hydrogen (q
LHV
119,950 kJ/kg,
ex
fuel
117,120 kJ/kg).
For a reversible heat engine cycle operating between
a source at T
s
and a sink at T
0
, the thermal efciency is
given by
Z
th;rev
1
T
0
T
s
. (14)
The ratio of the actual thermal efciency to the thermal
efciency of a reversible heat engine operating between the
same temperature limits gives a type of exergy efciency of
the heat engine. For a furnace temperature of T
s
1300 K
and an environment temperature of T
0
298 K, the
reversible thermal efciency found with Eq. (9) is 77.1%.
Dividing the actual thermal efciency of 34.1% by this
efciency (0.341/0.771) gives 44.2%. Note that this is the
same as the exergy efciency obtained using Eq. (7).
The results of the numerical example considered in this
section are shown in a combined energy and exergy
diagram in Fig. 2. In many articles with energy and exergy
analyses of power cycles, energy and exergy ow diagrams
are given separately. The combined ow diagram approach
used here appears to be useful in conveying energy and
exergy results of the cycle in a scaled, compact and
comprehensive manner. The heating value of the fuel is
normalized to 100 units of energy and other values
are normalized accordingly. The thermal and exergy
ARTICLE IN PRESS
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3971
efciencies discussed in this section can easily be obtained
using the values in this diagram by taking the ratios of
various terms. The total exergy destruction in this power
plant is 78 kJ for a total exergy input of 103.6 kJ. The
exergy destruction in the cycle based on an exergy input of
33.2 kJ is only 7.6 kJ (33.225.6), which is only 9.7% of
total exergy destruction. That is, the exergy destructions in
the furnaceboiler system account for the remaining 90.3%
of the total exergy destruction. This signicant exergy
destruction is not considered in an exergy efciency
denition neglecting the destructions in the furnaceboiler
system (see Eq. (4)).
One may question the value of exergy analysis as a tool
for assessing a power plant because the thermal efciency
based on the heating value of the fuel (Eq. (3)) and the
exergy efciency based on the exergy of the fuel (Eq. (8))
are very close. Although the exergy efciency in this case
adds little new information for addressing cycle efciency,
we have to remember that a major use of exergy analysis is
to analyze the system components separately and to
identify and quantify the sites of exergy destruction. This
information can then be used to improve the performance
of the system by trying to minimize the exergy destructions
in a prioritized manner. Note that the exergy efciency
dened in Eq. (6) addresses the fact that only a fraction of
the heat from combustion that is transferred to the steam
in the boiler is available for work, and the exergy efciency
compares the actual work output to this available work
(i.e., exergy). The exergy efciencies in these cases become
greater than the corresponding thermal efciencies, provid-
ing more realistic measures of system performance
compared to the corresponding thermal efciencies. For a
more comprehensive thermodynamic analysis of a power
cycle, the various energy- and exergy-based efciencies are
best considered.
4. Cogeneration plants
Cogeneration refers to the simultaneous generation of
more than one form of energy product.
4.1. Efciencies
For a cogeneration plant producing electric power and
process heating, a rst-law-based efciency is dened as the
ratio of useful energy output to energy input:
Z
cogen

_
W
net;out

_
Q
process
_
Q
in
1
_
Q
loss
_
Q
in
, (15)
where
_
Q
process
is the output rate of process heat and
_
Q
loss
is
the heat lost in the condenser. This relation is referred to as
the utilization efciency to differentiate it from the thermal
efciency which is used for a power plant where the single
output is power. Students are consistently taught not to
compare apples and oranges, which usually refers to two
commodities that are different. Work and heat have the
same units but are fundamentally difcult to add because
they are different, with work being a more valuable
commodity than heat.
We can overcome this situation by dening the efciency
of a cogeneration plant based on exergy, as the ratio of
total exergy output to exergy input:
Z
ex;cogen

_
Ex
out
_
Ex
in

_
W
net;out

_
Ex
process
_
Ex
in
1
_
Ex
dest
_
Ex
in
, (16)
where
_
Ex
process
is the exergy transfer rate associated with
the transfer of process heat, expressible as
_
Ex
process

Z
d
_
Q
process
1
T
0
T

, (17)
where T is the instantaneous source temperature from
which the process heat is transferred. This relation is of
little practical value unless the functional relationship
between the process heat rate
_
Q
process
and temperature T is
known. In many cases, the process heat is utilized by the
transfer of heat from a working uid exiting the heat
producing device (e.g., a turbine or an internal combustion
ARTICLE IN PRESS
q
LHV
=100
ex
fuel
=103.6
q
in
=75.0
q
lost,boiler
=25.0
w
net,out
=25.6
ex
dest
=78.0
q
lost,cond
=49.4
ex
steam
=33.2
ex
heat
=57.9
Fig. 2. Combined energy and exergy diagram for the steam power plant considered.
Diesel
engine
Heat
exchanger
Fuel
Air
Exhaust
Power
1
2
4
3
Water
Fig. 3. Cogeneration plant with a diesel engine and a heat exchanger for
steam production.
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3972
engine) to a secondary uid in a heat exchanger (Fig. 3).
One can express the exergy rate of process heat as the
exergy decrease of the hot uid in the heat exchanger as
_
Ex
process1
D
_
Ex
hot
_ m
hot
h
1
h
2
T
0
s
1
s
2

hot
,
(18)
or by the increase of the exergy of the cold uid in the heat
exchanger
_
Ex
process2
D
_
Ex
cold
_ m
cold
h
4
h
3
T
0
s
4
s
3

cold
,
(19)
where the subscripts refer to state points in Fig. 3. The
difference between these two exergies is the exergy
destruction in the heat exchanger. Then, from Eq. (16),
the exergy efciencies based on these two approaches
become
Z
ex;cogen1

_
W
net;out
_ m
hot
h
1
h
2
T
0
s
1
s
2

hot
_
Ex
in
(20)
and
Z
ex;cogen2

_
W
net;out
_ m
cold
h
4
h
3
T
0
s
4
s
3

cold
_
Ex
in
.
(21)
The exergy input in these relations can be expressed
differently using various inputs as in Eqs. (6)(8), yielding
different exergy efciencies.
4.2. Illustrative example
To illustrate the use of these efciencies, we consider a
diesel engine-based cogeneration plant. The outputs are
electrical power and process heat, which is transferred from
the hot exhaust gases to water to produce steam in a heat
exchanger (Fig. 3). Some of the data used in this example
are from an actual diesel engine power plant (Kanoglu
et al., 2005). The net power output from the plant is
18,900 kW when the fuel consumption rate is 1.03 kg/s and
airfuel ratio is 40.4. This corresponds to an exhaust ow
rate of 41.6 kg/s. The plant uses heavy diesel fuel with a
lower heating value of 39,300 kJ/kg. The exhaust gases
enter the process heating unit (i.e., heat exchanger) at
383 1C and experience a temperature drop of 175 1C while
compressed liquid water enters at 15 1C and exits as
saturated vapor at 200 1C. Applications of Eqs. (15)(21)
produce the following results:
_
Q
process
7784 kW;
_
Ex
in
43; 110 kW;
_
Ex
process1
3678 kW;
_
Ex
process2
2509 kW;
Z
cogen
65:9%; Z
ex;cogen1
52:4%,
Z
ex;cogen2
49:7%.
The exergy of heavy diesel fuel with an unknown
composition is taken as 1.065 times the lower heating
value of the fuel following the approach by Brzustowski
and Brena (1986). Properties of air with variable specic
heats are used for exhaust gases.
The difference between the energy and exergy efciencies
in this cogeneration plant appears to be much greater than
the difference for a power plant, when the energy and
exergy efciencies are, respectively, dened based on the
energy and exergy of the fuel, as discussed in the previous
section. The difference is attributable to one of the product
outputs being process heat. The different approaches used
in Eqs. (18) and (19) to dene the exergy of the process heat
result in a small exergy efciency difference of only
52.449.7 2.7%. The greater the average temperature
difference between the hot and cold uids in the process
heater, the greater is the exergy destruction and the greater
is the difference between the two denitions of exergy
efciencies in Eqs. (20) and (21), respectively.
The results are presented in a combined energy and
exergy diagram in Fig. 4. The heating value of the fuel
(i.e., heat input) is normalized to 100 units of energy and
other values are modied accordingly. The energy and
exergy efciencies discussed in this section can be found
using this diagram. The total exergy destruction in this
cogeneration plant is 50.7 kJ based on the rst approach
and 53.6 kJ based on the second approach, for a total
exergy input of 106.5 kJ. The difference between these
exergy destructions is the exergy destruction in the process
heater, which is 5.7% of total exergy destruction or 2.7%
of exergy input.
5. Geothermal power plants
The technology for producing power from geothermal
resources is well-established and there are many geother-
mal power plants operating worldwide (Barbier, 1997).
Depending on the state of the geothermal uid in the
reservoir, different power producing cycles may be used
including direct steam, ash steam (single and double
ash), binary and combined ash-binary cycles.
5.1. Efciencies
In general, the thermal efciency of a geothermal power
plant may be expressed as
Z
th1

_
W
net;out
_
E
in
, (22)
where
_
E
in
is the energy input rate to the power plant, which
may be expressed as the specic enthalpy of the geothermal
water with respect to environment state multiplied by the
mass ow rate of geothermal water _ m
geo
. That is
Z
th1

_
W
net;out
_ m
geo
h
geo
h
0

. (23)
The state of geothermal water may be taken as that in
the reservoir or at the well head. Those who use the
reservoir state argue that a realistic and more meaningful
ARTICLE IN PRESS
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3973
comparison between geothermal power plants needs to
account for methods of harvesting the geothermal uid.
However, those who use the well head state argue that
taking the reservoir as the input is not appropriate for
geothermal power plants since conventional power plants
are evaluated on the basis of the energy of the fuel burned
at the plant site (Kestin, 1980; DiPippo, 1994; Kanoglu,
2004). In Eq. (23), the energy input to the power plant
represents the maximum heat the geothermal water can
deliver, which occurs when the geothermal water is cooled
to the temperature of the environment.
The simplest geothermal cycle is the direct steam cycle.
Steam from the geothermal well is passed through a turbine
and exhausted to the atmosphere or to a condenser. Flash
steam plants are used to generate power from liquid-
dominated resources that are hot enough to ash a
signicant proportion of the water to steam in surface
equipment, either at one or two pressure stages (single-ash
or double-ash plants) as shown in Figs. 5 and 6,
respectively. The steam ows through a steam turbine to
produce power while the brine is reinjected back to the
ground. Steam exiting the turbine is condensed with
cooling water obtained in a cooling tower or a spray pond
before being reinjected. Binary cycle plants use the
geothermal brine from liquid-dominated resources
(Fig. 7). These plants operate on a Rankine cycle with a
binary working uid (isobutane, isopentane, R-114, etc.)
that has a low-boiling temperature. The working uid is
completely vaporized and usually superheated by the
geothermal heat in the vaporizer. The vapor expands in
the turbine, and then condenses in a water-cooled
condenser or dry cooling tower before being pumped
back to the vaporizer to complete the cycle. Combined
ARTICLE IN PRESS
q
LHV
=100
ex
fuel
=106.5
w
net,out
=46.7
q
out
=34.1
ex
process-1
=9.1
ex
dest-1
=50.7
q
process
=19.2
ex
dest-1
=53.6
ex
process-2
=6.2
Fig. 4. Combined energy and exergy diagram for the cogeneration plant considered.
from
production
well
to
reinjection
well
separator
steam
turbine
1
2
4
3
5
6
condenser
expansion
valve
7
8
Fig. 5. Single-ash geothermal power plant.
from
production
well
to
reinjection
well
steam
turbine
condenser
separator
separator
expansion
valve
1
2
3
5
4
8
9
11
10
6
7
Fig. 6. Double-ash geothermal power plant.
Condenser
Heat exchanger
Geothermal
water
2
1
4 3
5 6
Q
Pump
Turbine
Fig. 7. Binary geothermal power plant.
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3974
ash/binary plants (Fig. 8) incorporate both a binary unit
and a ashing unit to exploit the advantages associated
with both systems. The liquid portion of the geothermal
mixture serves as the input heat for the binary cycle while
the steam portion drives a steam turbine to produce power.
The actual heat input to a geothermal power cycle is less
than the term in the denominator of Eq. (23) since part of
geothermal water is reinjected back to the ground at a
temperature much greater than the temperature of the
environment. In an approach that accounts for the actual
reinjection temperature, the thermal efciency is expressed as
Z
th2

_
W
net;out
_
Q
in
. (24)
For a single-ash cycle, the thermal efciency may be
expressed as
Z
th;singleflash

_
W
net;out
_
Q
in

_
W
net;out
_ m
2
h
2
_ m
3
h
3
, (25)
where the subscripts refer to state points in Fig. 5. For a
double-ash cycle, the efciency becomes
Z
th;doubleflash

_
W
net;out
_ m
2
h
2
_ m
3
h
3
_ m
5
h
5
_ m
6
h
6

, (26)
where the state points are shown in Fig. 6. Referring to
Fig. 7, for a binary cycle we obtain
Z
th;binary

_
W
net;out
_ m
geo
h
1
h
2

(27)
or
Z
th;binary

_
W
net;out
_ m
binary
h
4
h
3

, (28)
where _ m
binary
is the mass ow rate of binary working
uid. For a combined ash-binary cycle, the thermal
efciency is
Z
th;flashbinary

_
W
net;out
_ m
2
h
2
_ m
3
h
3
_ m
3
h
3
_ m
7
h
7

, (29)
where the state points are shown in Fig. 8.
Using the exergy of geothermal water (in the reservoir or
at the well head) as the exergy input to the plant, the exergy
efciency of a geothermal power plant can be expressed as
Z
ex

_
W
net;out
_
Ex
in

_
W
net;out
_ m
geo
h
geo
h
0
T
0
s
geo
s
0

. (30)
Using the exergy change of geothermal water in the cycle
as the exergy input to the cycle, the exergy efciencies may
be expressed for single ash, double ash and combined
ash-binary cycles as
Z
ex;singleflash

_
W
net;out
_ m
2
ex
2
_ m
3
ex
3

_
W
net;out
_ m
2
h
2
h
0
T
0
s
2
s
0
_ m
3
h
3
h
0
T
0
s
3
s
0

,
31
Z
ex;doubleflash

_
W
net;out
_ m
2
ex
2
_ m
3
ex
3
_ m
5
ex
5
_ m
6
ex
6

, (32)
Z
ex;flashbinary1

_
W
net;out
_ m
2
ex
2
_ m
3
ex
3
_ m
3
ex
3
_ m
7
ex
7

,
(33)
where ex is the specic ow exergy of the uid. For a
binary cycle, the exergy efciency may be dened based on
the exergy decrease of geothermal water or the exergy
increase of the binary working uid in the heat exchanger.
That is,
Z
ex;binary1

_
W
net;out
_ m
geo
h
2
h
1
T
0
s
2
s
1

. (34)
Z
ex;binary2

_
W
net;out
_ m
binary
h
4
h
3
T
0
s
4
s
3

. (35)
The difference between the denominators of Eqs. (34)
and (35) is the exergy destruction in the heat exchanger.
The exergy efciency denitions in Eqs. (34) and (35) can
be illustrated by considering the different systems indicated
by the inner and outer dashed lines, respectively, in Fig. 7.
By adapting the approach used in Eq. (35), one may
express the exergy efciency for a combined ash-binary
cycle as
Z
ex;flashbinary2

_
W
net;out
_ m
2
ex
2
_ m
3
ex
3
_ m
binary
ex
11
ex
10

.
(36)
The efciency in Eq. (33) is more advantageous than that
in Eq. (36) because exergy input is expressed by the exergy
change of geothermal water for both the ash and binary
parts of the cycle in Eq. (33), respectively.
ARTICLE IN PRESS
steam
turbine
from
production
well
to
reinjection
well
turbine
heat exchanger
pump
1
separator
condenser
condenser
expansion
valve
2
3
4
5
6
7
8
9
11
12
13
10
Fig. 8. Combined ash-binary geothermal power plant.
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3975
5.2. Illustrative example
Consider a binary geothermal power plant like that in
Fig. 7 using geothermal water at 165 1C with isobutane as
the working uid. The mass ow rate of geothermal water
is 555 kg/s. In this cycle, isobutane is heated and vaporized
in the heat exchanger by geothermal water. Then, the
isobutane ows through the turbine, is condensed and
pumped back to the heat exchanger, completing the binary
cycle. The heat exchanger and condenser pressures are
taken to be 3000 and 400 kPa, respectively, while the
temperature at the turbine inlet (or heat exchanger exit) is
taken to be 150 1C, which is 15 1C lower than the
geothermal water temperature at the heat exchanger inlet.
The isentropic efciencies of the turbine and pump are
taken to be 80% and 70%, respectively. About 10% of
power output is used for internal demands such as
powering fans in the air-cooled condenser. These values
closely correspond to those of an actual power plant
(Kanoglu and Cengel, 1999). Noting that a pinch-point will
occur at the start of vaporization of the working uid in the
heat exchanger, the energy balance relations for the heat
exchanger can be written as
_ m
geo
c
geo
T
1
T
vap
DT
pp
_ m
binary
h
4
h
binary;f
,
(37)
_ m
geo
c
geo
T
vap
DT
pp
T
2
_ m
binary
h
binary;f
h
3
,
(38)
where _ m
binary
is the mass ow rate of the binary uid, c
geo
the specic heat of geothermal water, T
vap
the vaporization
temperature of the binary uid at the heat exchanger
pressure, DT
pp
the pinch-point temperature difference and
h
binary,f
is the specic enthalpy of the binary uid at the
start of vaporization. The pinch-point temperature differ-
ence is assumed to be 6 1C. Eqs. (37) and (38) can be used
to establish the mass ow rate of the binary uid, and the
geothermal water temperature at the heat exchanger exit.
The analysis of the cycle with the stated values produces
the following results:
_
W
net;out
22; 382 kW;
_
E
in
328; 786 kW;
_
Q
in
185; 181 kW; T
2
86:6

C;
_
Ex
in
60; 014 kW; D
_
Ex
12
46; 904 kW;
D
_
Ex
34
37; 316 kW; Z
th1
6:8% Eq:23;
Z
th2
12:1% Eq:27; Z
ex
37:3% Eq:30;
Z
ex;binary1
47:7% Eq:34; Z
ex;binary2
60:0% Eq:35.
It is clear that using different denitions leads to
signicantly different thermal and exergy efciencies. This
is typical of geothermal power plants. The results are
presented in a combined energy and exergy diagram in
Fig. 9. Because of the large range of values involved, the
exergy of geothermal water at the heat exchanger inlet is
normalized to 100 units of energy and other values are
modied accordingly. The energy and exergy efciencies
can be obtained using terms in this diagram. The thermal
and exergy efciencies are 6.8% and 37.3%, respectively,
based on the energy and exergy of geothermal water at the
heat exchanger inlet. The thermal efciency increases from
6.8% to 12.1% when the heat input to the binary uid in
the heat exchanger is used as the energy input to the cycle
instead of the energy of the geothermal water at the inlet of
the plant. This is analogous to using the heating value of
fuel versus using the heat transferred to the steam in the
boiler as the heat input to a steam power plant.
The exergy efciency is only 37.3% when the exergy of
geothermal water at the plant inlet is used. Using the
exergy decrease of geothermal water in the heat exchanger
as the exergy input to the cycle yields an exergy efciency
of 47.7% while using the exergy increase of the binary uid
yields an exergy efciency of 60.0%. These three ap-
proaches are analogous to using the exergy of the fuel
(Eq. (8)), the exergy transfer to the steam accompanying
the heat input to the cycle (Eq. (7)) and the exergy increase
of the steam in the boiler (Eq. (6)). The exergy of the
geothermal water at the exit of the heat exchanger, which is
ARTICLE IN PRESS
ex
reinject
=21.8
ex
dest
=62.7
ex
1-2
=78.2
ex
in
=100
e
in
=547.8
q
in
=308.6
ex
3-4
=62.2
w
net,out
=37.3
Fig. 9. Combined energy and exergy diagram for the binary geothermal power plant considered.
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3976
reinjected to the ground, represents 21.8% of the exergy
input to the cycle. This signicant percentage is due to the
relatively high temperature of the geothermal water
(86.6 1C). The exergy destruction in the heat exchanger
accounts for 16.0% of the exergy input. The remaining
exergy destructions (10021.816.037.3 24.9) are due
to irreversibilities in the turbine, pump and condenser.
6. Conclusions
Insights are presented into various energy and exergy
efciencies used in power plants. The efciencies for the
plants considered as examples yield some important
information on the relative magnitudes of heat losses and
exergy destructions in the plants. Combined energy and
exergy diagrams present the results concisely and clearly.
The efciencies for power cycles not specically discussed
in this article can be deduced from the relations given for
the cycles considered.
For the current state of thermodynamics, it seems almost
impossible to have a common efciency denition for all
energy systems. Therefore, the best way of avoiding misuse
and misunderstanding is to dene the efciency used in any
application carefully. An understanding of both energy and
exergy efciencies is essential for designing, analyzing,
optimizing and improving energy systems through appro-
priate energy policies and strategies. If such policies and
strategies are in place, numerous measures can be applied
to improve the efciency of electrical generating plants.
These measures should be weighed against other factors
and, where appropriate, implemented. It should be under-
stood that decisions on power plant operations are
normally based primarily on economic criteria. Often
other criteria such as environmental considerations are
also important. Economic and exergy analyses can be
combined by means of exergoeconomic analyses, which can
include exergetic life cycle assessment. A rational efciency
denition should accompany such an analysis. It is more
appropriate to use an exergy efciency based on the exergy
of the fuel in a fossil-fuel power or cogeneration system
since an important part of exergy costing involves fuel cost.
In this approach, all exergy losses are accounted for. For
renewable energy systems, it is more appropriate to use the
exergy of an energy source as the exergy input to the
system. This approach allows all exergy destructions to be
accounted for, including those in heat exchange equipment.
All losses are ultimately related to the economics of the
system operation. The difference between efciency deni-
tions often relates to the selection of different system
boundaries. Depending on the selection, losses occurring at
a particular site may be accounted for in a denition or
excluded.
It is expected that this article will help engineers,
researchers and policy makers in the area of energy
management for power plants appreciate that exergy
analysis is a useful tool and make better use of energy
and exergy efciencies in energy management.
Acknowledgments
The authors acknowledge the support provided by
University of Gaziantep in Turkey, and in Canada by
University of Ontario Institute of Technology and the
Natural Sciences and Engineering Research Council.
References
Barbier, E., 1997. Nature and technology of geothermal energy: a review.
Renewable and Sustainable Energy ReviewsAn International
Journal 1 (2), 169.
Bejan, A., 2006. Advanced Engineering Thermodynamics, third ed. Wiley,
New York.
Bisio, G., Rubatto, G., 2001. Comparing hydraulic and polytropic efciencies
with exergy efciency. Exergy International Journal 1, 193201.
Brookes, L., 2004. Energy efciency fallacies-a postscript. Energy Policy
32, 945947.
Brzustowski, T.A., Brena, A., 1986. Second law analyses of energy
processes. IVthe exergy of hydrocarbon fuels. Transactions of the
Canadian Society for Mechanical Engineering 10 (3), 121128.
Cengel, Y.A., Boles, M.A., 2006. Thermodynamics: An Engineering
Approach, fth ed. McGraw-Hill, New York.
Consoli, F., 1993. Guidelines for Life-cycle Assessment: A Code of
Practice. SETAC, Brussels, Belgium.
Cornelissen, R.L., 1997. Thermodynamics and sustainable development:
the use of exergy analysis and the reduction of irreversibility, Ph.D.
Thesis, University of Twente.
Cornelissen, R.L., Hirs, G.G., 2002. The value of the exergetic life cycle
assessment besides the LCA. Energy Conversion and Management 43
(9-12), 14171424.
Cornelissen, R.L., Hirs, G.G., Kotas, T.J., 1995. Different denitions of
exergetic efciencies. In: Proceedings of JETC IV, Nancy, France,
pp. 335344.
Dincer, I., Rosen, M.A., 2005. Thermodynamic aspects of renewables and
sustainable development. Renewable and Sustainable Energy Reviews
9, 169189.
Dincer, I., Hussain, M.M., Al-Zaharnah, I., 2004. Energy and exergy use
in public and private sector of Saudi Arabia. Energy Policy 32,
16151624.
DiPippo, R., 1994. Second law analysis of ash-binary and multilevel
binary geothermal power plants. Geothermal Resources Council Trans
18, 505510.
Kanoglu, M., 2001. Cryogenic turbine efciencies. Exergy: an Interna-
tional Journal 1, 202208.
Kanoglu, M., 2004. Exergy analysis of a dual-level binary geothermal
power plant. Geothermics 31, 709724.
Kanoglu, M., Cengel, Y.A., 1999. Improving the performance of an
existing binary geothermal power plant: a case study. Transactions of
the ASME, Journal of Energy Resources Technology 121 (3), 196202.
Kanoglu, M., Isik, S.K., Abusoglu, A., 2005. Performance characteristics
of a diesel engine power plant. Energy Conversion and Management
46, 16921702.
Kestin, J., 1980. Available work in geothermal energy. In: Sourcebook on
the Production of Electricity from Geothermal Energy. US Depart-
ment of Energy, Washington, DC.
Kotas, T.J., 1995. The Exergy Method in Thermal Plant Analysis, second
ed. Krieger, Malabar.
Lombardi, L., 2003. Life cycle assessment comparison of technical
solutions for CO
2
emissions reduction in power generation. Energy
Conversion and Management 44 (1), 93108.
Pulkrabek, W., 2004. Engineering Fundamentals of the Internal Combus-
tion Engine, second ed. Prentice-Hall, New York.
Rosen, M.A., Le, M.N., Dincer, I., 2005. Efciency analysis of a
cogeneration and district energy system. Applied Thermal Engineering
25, 147159.
ARTICLE IN PRESS
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3977
Sonntag, R.E., Borgnakke, C., Van Wylen, G.J., 2002. Fundamentals of
Thermodynamics, sixth ed. Wiley, New York.
Struchtrup, H., Rosen, M.A., 2002. How much work is lost in
an irreversible turbine? Exergy, an International Journal 2,
152158.
Szargut, J., Morris, D.R., Steward, F.R., 1988. Exergy Analysis of
Thermal, Chemical, and Metallurgical Processes. Hemisphere Publish-
ing Corp., New York.
Tsatsaronis, G., 2007. Denitions and nomenclature in exergy analysis
and exergoeconomics. Energy 32 (4), 249253.
Tsatsaronis, G., Lin, L., Pisa, J., 1993. Exergy costing in exergoeconomics.
Journal of Energy Resources Technology, Transactions of the ASME
115 (1), 916.
Zaleta, A., Royo, J., Valero, A., 2001. Thermodynamic model of the loss
factor applied to steam turbines. International Journal of Applied
Thermodynamics 4 (3), 127133.
ARTICLE IN PRESS
M. Kanoglu et al. / Energy Policy 35 (2007) 39673978 3978

Das könnte Ihnen auch gefallen