Está en la página 1de 26

Energy 29 (2004) 389414

www.elsevier.com/locate/energy

Thermoeconomic optimization of heat recovery steam


generators operating parameters for combined plants
C. Casarosa a, F. Donatini b, A. Franco a,
a

Universita` di Pisa, Dipartimento di Energetica, via Diotisalvi, 2-56126 Pisa, Italy


b
Enel Produzione, via A. Pisano, 120-56122 Pisa, Italy

Abstract
The optimization of the heat recovery steam generator (HRSG) is particularly interesting for the combined plants design in order to maximise the work obtained in the vapour cycle. A detailed optimization
of the HRSG is a very difficult problem, depending on several variables. The first step is represented by
the optimization of the operating parameters. These are the number of pressure levels, the pressures, the
mass flow ratio, and the inlet temperatures to the HRSG sections. The operating parameters can be determined by means both of a thermodynamic and of a thermoeconomic analysis, minimising a suitable objective
function by analytical or numerical mathematical methods. In the paper, thermodynamic optimization is
based on the minimization of exergy losses, while the thermoeconomic optimization is based on the minimization of the total HRSG cost, after the reduction to a common monetary base of the costs of exergy
losses and of installation.
2002 Elsevier Ltd. All rights reserved.

1. Introduction
The competitive business environment created by liberalization in several countries of the electricity supply sector has made the combined cycle power plant, with its cost benefits and low
environmental impact, the generating plant of choice for power producers. Many of them produce
standardized combined cycle power facilities based on advanced gas turbines models with an
efficiency in the range between 35 and 42% and exhaust gas temperatures variable in the range
between 700 K and 920 K.
An advanced, high-efficiency combined cycle power plant depends not only on high-quality

Corresponding author. Tel.: +39-50-569632; fax: +39-50-569604.


E-mail address: alessandro.franco@ing.unipi.it (A. Franco).

0360-5442/$ - see front matter 2002 Elsevier Ltd. All rights reserved.
doi:10.1016/S0360-5442(02)00078-6

390

C. Casarosa et al. / Energy 29 (2004) 389414

Nomenclature
b
dimensionless cost parameter
cp
specific heat (J/kg K)
C
specific heat of the gas (J/kg K)
D
economic life of the plant (years)
Ex
exergy (J)
Exr
dimensionless residual exergy at the end of the steam expansion process
H
annual functioning duration (hours/year)
I
exergy losses (J)
I
dimensionless exergy losses
Iesp
dimensionless exergy losses due to the isoentropic efficiency of the steam turbine
kI
specific cost of the exergy losses ($/kWh )
k
specific cost of the HRSG surfaces ($/m2)
K
cost relative to one year ($)
K
dimensionless annual cost
m
mass flow rate of water (kg/s)
M
mass flow rate of the gas (kg/s)
n
number of HRSG sections
N=US/Mcpg number of gas side transfer units
P.P. pinch-point (K)
p
pressure (bar)
S
heat exchange surface (m2)
T
gas temperature (K)
t
water (steam) temperature (K)
ta
environment temperature (K)
tv
vaporization temperatures (K)
tsh
temperature at the end superheater (K)
U
overall heat transfer coefficient (W/m2 K)
x
steam quality at the end of the expansion process
W
output power of the steam cycle (W)
W
dimensionless output power of the steam cycle
r
ratio between heat rates of liquid and gas
h
gas-side effectiveness
Subscripts and superscripts
a
e
esp
exp

environment
economiser
expansion
exponent

C. Casarosa et al. / Energy 29 (2004) 389414

g
HRSG
in
I
out
r
rh
sh
st
v
vap
w

391

gas
heat recovery steam generator
inlet to the HRSG
of the exergy losses
outlet from the HRSG
residual value
reheater
superheater
steam turbine
of the evaporator
of the vapor
of the water

Acronyms and abbreviations


CCP Combined Cycle Plant
ECO Economiser section
EVA Evaporator section
HRSG Heat Recovery Steam Generator
HP
High pressure
IP
Intermediate pressure
LP
Low pressure
Px
Section with two or three water streams
RAN Single pressure HRSG (Rankine cycle HRSG)
RANSH Single pressure HRSG with superheater
SH
Superheater section
THP High Pressure Turbine
TIP
Intermediate Pressure Turbine
TLP Low Pressure Turbine
TMD Thermodynamic
TME Thermoeconomic
2PR Simple two pressure level

equipment but also on optimized matching of the components. The combined cycle process
couples the Brayton cycle with a bottoming Rankine cycle. The basic idea is to utilize the energy
contained in the hot gas-turbine exhaust gases in a steam process connected downstream of the
gas turbine. The gas turbinethe topping cycleprovides, in a process configuration with a
heat recovery steam generator (HRSG), about two thirds of the total useful power. The remaining
third comes from utilization of the waste-heat in the steam process, or bottoming cycle. To
increase the power output, a most promising argument is the optimization of the HRSG, that

392

C. Casarosa et al. / Energy 29 (2004) 389414

provides the critical link between the gas turbine cycle and the steam turbine cycle, with the
objective of increasing the steam turbine output [1].
In a combined cycle power plant (CCP) the HRSG represents the interface element between
the gas turbine and the steam cycle. Here, the gas turbine exhaust gas is cooled down and the
recuperated heat is used to generate steam. In order to provide better heat recovery in the HRSG
more than one pressure level is used. With a single-pressure HRSG about 30% of the total plant
output is generated in the steam turbine. A dual-pressure arrangement can increase the power
output of the steam cycle by up to 10%, and an additional 3% can result with a triple-pressure
cycle [2]. To obtain a higher efficiency value, a steam bottoming cycle with five pressure levels
with steam turbo-chargers has also been proposed [3].
Modern gas turbine plants with a triple-pressure HRSG with steam reheat can reach efficiencies
above 55%. ABB-Alstom claims 58% efficiency of a combined cycle plant built around their
GT24/26 reheat gas turbines [4]; the same efficiency is cited by Siemens, for the Westinghouse
steam-cooled W501G/701G gas turbine or V94.3a gas turbine, combined with a triple pressure
HRSG [5]. A gas turbine, with steam cooling of the turbine blades and nozzles, combined with
an advanced HRSG is expected to operate at an efficiency level of 60% in the near future. However, these high efficiency values can be achieved at large units above 300 MW. The purpose of
the most part of the world manufacturers is to reach overall thermal efficiencies of the combined
plant of 60% in the short period, above all increasing the gas turbine inlet temperature [5].
In the literature, different solutions for the increase of the heat recovery efficiency have been
proposed. Among these is the well known Kalina cycle, where a mixed working fluid of variable
composition is used to provide a better match between the temperatures of the hot and cold
streams. An alternative, the use of a fluid with a high critical temperature, such as mercury or
potassium, could be interesting for the bottoming cycle too [5].
Maintaining the use of water as working fluid for the bottoming cycle, the optimization of the
HRSG operating parameters is considered one of the most interesting strategies to obtain an
increase of the combined cycle plant performance. This has been demonstrated in recent work
about the optimization, both of the HRSG [68] and of the whole combined cycle power plant [9].
In general, thermoeconomic analyses proposed in the literature are focused on the economic
objective. But solutions with high thermodynamic efficiency, in spite of an increase of total costs,
may provide much more interesting results due to changes in energy market prices and in energy
policies. Moreover, the analysis is concentrated on general elements of the system.
Considering that further improvements and refinements on this matter are desirable, mainly in
order to obtain a satisfactory thermoeconomic optimization strategy for the whole combined plant
providing a suitable compromise between thermodynamic and economic aspects, it is also interesting to perform new optimization strategies that permit the definition of the various HRSG
operating parameters and consequently the geometrical variables of the single sections.
Starting from those premises, the objective of the paper is to develop a particular optimization
method, based on a combined thermodynamic and economic analysis, for the design of the HRSG,
that tries to satisfy both thermodynamic and economic objectives in order to define the main
operating parameters of the component. The main operating parameters considered for the HRSG
are the number of pressure levels, the mass ratios between hot and cold stream and the saturation
temperatures. The optimization is performed with the aim of evaluating the possibility of reaching
efficiency of 60% without awaiting for a meaningful increase of the gas turbine performances.

C. Casarosa et al. / Energy 29 (2004) 389414

393

2. Analysis of the HRSG


The optimization of a component such as the HRSG can occur at various levels of complexity
with objective functions sequentially defined. The optimization, beginning with the global system
to the component with different hierarchical objectives, determines a quasi-optimum design.
For the HRSG the first step of the optimization is represented by the definition of the main
operating parameters. These are in general the number of pressure levels, the mass flow ratio and
the temperature profiles. The HRSG optimum design in the actual technology is based on the
concepts of pinch-point and approach point, which govern the gas and steam temperature profiles.
Particularly important in order to quantify the operating parameters is the pinch-point. The pinchpoint (Fig. 1) representing the minimum difference between the gas temperature leaving the evaporator and the saturation temperature, takes into account implicitly both thermodynamic and economical points of view. Its values, usually in the range between 10 and 20C, are derived from
practical experience even if it is clear that they are strongly dependent on the economic viewpoint considered.
A different optimization approach can be based on a thermodynamic or thermoeconomic analysis by minimising a suitable objective function with analytical or numerical mathematical methods.
In order to identify the aforesaid objective function and the correct weight of thermodynamic and
economic elements, a necessary first step is the splitting of thermodynamic and economic aspects
and the pinch-point becomes a result of the optimization process.
In a thermodynamic analysis, the attention can be focused on the minimization of the thermal
exergy losses, taking into account only the irreversibility due to the temperature difference
between hot and cold streams. The exergy losses due to the pressure drop, neglected at this step
because they are of an order of magnitude lower, can be considered in a further analysis of the
HRSG ended at the final executive design of the HRSG, requiring for their evaluation a detailed
description of the sections geometry. Although this criterion gives solutions with null pinch- point
and infinite heat exchange surface, it allows us to perform a first selection of the operating parameters: mass flow ratio, exhaust gas and water outlet temperatures from the HRSG and saturation

Fig. 1. Typical HRSG temperature profiles.

394

C. Casarosa et al. / Energy 29 (2004) 389414

temperatures. Moreover, it is suggested that the use of a section with two or three parallel cold
streams is a means to increase the heat transfer effectiveness of the HRSG.
A next step is the tentative of obtaining a general optimization combining thermodynamic and
economic aspects. In this case, the cost related to the exergy losses have to be recast to operating
costs of the HRSG and, combined with the HRSG installation costs, gives a cost that must be
minimized. It is shown among the results how the weight of the economic aspects modify the
results of the thermodynamic analysis, evidencing that in this approach the critical element is the
value of the ratio between the cost of exergy losses and the HRSG installation cost.
The thermoecomomic optimization determines a sensible difference in the results with respect
to the thermodynamic optimization and shows how a pinch-point value corresponds to a well
defined value of the aforesaid ratio between the operating costs and the costs of the exergy losses.
The optimum design of various HRSG configurations of growing complexity is examined furnishing a general strategy to select the independent variables of the optimum design problem. In
all the cases discussed the HRSG inlet mass flow of the gas and its temperature are imposed as
boundary conditions. The application of the proposed methods is also extended to consider some
HRSG configurations with two or three pressure levels, relative to commercialized combined
power plants.
3. Thermodynamic optimization
A thermodynamic optimization is the first step of each optimum design process. Though if
approaches for the HRSG optimization based on the minimization of entropy generation are also
available [7], the criterion considered for the thermodynamic optimization yields the minimisation
of exergy losses, taking into account irreversibility due to the temperature difference between the
hot and the cold stream. In the general case, considering null the exergy flows of the outlet gas,
the exergy balance of the HRSG is given by [10]
Exg,in Exl,in Exvap,out I

(1)

where I is the exergy loss occurring in the HRSG. The exergy loss I is a function that has to be
minimized to obtain the thermodynamic optimization
I Exg,in Exl,inExvap,out.

(2)

The general HRSG configuration can be obtained combining n sections similar to the one
represented in Fig. 2. The analysis is based on the concept of gas side effectiveness of the HRSG
sections instead of the pinch-point method. The gas side effectiveness or temperature effectiveness is derived from the concept of heat transfer effectiveness by Kays and London [11]. For
the single HRSG section, if the specific heat of the exhaust gas and of water/steam can be considered constant in each section, selecting an opportune average value in the corresponding temperature range, the gas side effectiveness can be defined as:
hk

TkTk1
f(N,r) (single phase channel)
Tktk1

(3)

where N is the number of gas-side transfer units, and r is the ratio between the heat rates of
water and gas, defined respectively as

C. Casarosa et al. / Energy 29 (2004) 389414

Fig. 2.

395

Heat exchanger or HRSG section.

US
mcpw
and r
,
Mcpg
Mcpg

(4)

where cpg and cpw are the specific heat of gas and liquid, U the overall heat transfer coefficient,
S the heat exchange surfaces, m and M liquid and gas mass flow rate.
The concept of gas side effectiveness or temperature effectiveness, defined by Eq. (3), is
only an instrument to conduct the optimization and has not a particular meaning in order to
characterize the efficiency of heat recovery. It is coincident with the usual heat exchanger
efficiency for the evaporator, while for the economiser and superheater sections, it represents the
ratio of the gas temperature drop and the maximum temperature difference between gas and liquid.
Using this concept, the HRSG can be analysed and optimised only analyzing the evolution of the
gas, a fluid that does not change phase.
An interesting solution in order to increase the heat transfer efficiency and reduce the thermal
exergy losses in the HRSG is the use of sections with more than one stream on the liquid side
(parallel flow sections). This solution is already used by the manufacturers mostly in the case of
fluid with the same characteristics, e.g. two liquids or two superheated steam flows. But it seems
interesting to propose an extension of its use also to different conditions.
Parallel flow sections are heat exchangers in which exhaust gas exchanges heat with two or
more water streams at the same time. These kind of sections are very useful when two or more
flows of water are required to accomplish the same temperature rise. In this case exergy losses
due to thermal exchange can be sensibly reduced due to the lower mean temperature difference
between gas and water with respect to the configuration with two simple-flow sections. The presence of two or more separated water streams is possible only if they are at different pressures,
so the number of liquid steams in parallel with the gas flow can be lower than or equal to the
number of pressure levels.
In Fig. 3(a), for example, a section with two water streams, at two different pressures p and
p is considered. It is easy to conclude that the situation minimizing the exergy losses is the one
when the following two conditions simultaneously occur: the first is that on the gas side the two
gas flows Mk and Mk mix at the same temperature Tk1, the second is that on the liquid side
the following conditions are verified
tk1 tk1 tk1

tk tk tk.

(5)

396

C. Casarosa et al. / Energy 29 (2004) 389414

Fig. 3.

General (a) and simplified (b) schematization of a section with two parallel streams.

If the previous conditions occur, the section can be represented by the simplified scheme of
Fig. 3(b) and it is possible to extend to the section with two or more water streams the same
concept of the gas side effectiveness defined by Eq. (3). To satisfy this condition only one value
of the number of gas side thermal units N has to be defined so it is necessary that:
N

US
US

.
Mcpg Mcpg

(6)

This condition can be verified even if the water streams are of different characteristics, i.e. a
liquid and a superheated steam. It is not possible to satisfy the condition when one or both the
water streams change phase. So the evaporators can be represented by elements with only one
water stream like the one of Fig. 2. In general, it is possible to use sections with a maximum
number of streams equal to the number of pressure levels. Moreover, it is important to consider
the connections between sections with a different number of streams, like those represented in
Fig. 4.
3.1. HRSG configuration and number of independent variables of the optimum design problem
Various HRSG configurations can be analyzed starting from the simple one, composed by a
Rankine cycle recovery structure (RAN), to the more complex represented by triple pressure

Fig. 4.

Connection between elements with one and two water streams.

C. Casarosa et al. / Energy 29 (2004) 389414

397

HRSG with reheat sections (3PRSH). In particular the configurations that will be examined are
the following:
1 pressure level
RAN
RANSH
RANSH+1P (reheater inlet temperature higher than saturation one)
RANSH+2P (reheater inlet temperature lower than saturation one)
2 pressure levels
2PR coupled (with a single steam turbine)
2PR uncoupled (with two steam turbines)
2PRSH (double pressure level with high pressure reheat)
3 pressure levels
3PRSH (triple pressure level with high pressure reheat)
In general, only the inlet temperature of the exhaust gas to the HRSG, the inlet temperature
of water and the gas mass flow are given. Each HRSG configuration involves a certain number
of variables that have to be defined (NV). These are, for an arrangement composed by elements
with only one stream, for each section the gas temperatures Tk, the water temperatures tk, the gasside effectiveness hk, the heat flows Qk, the gas mass flow Mk and the water mass flow mk. On
the same time, it is possible to write a certain number of equations: for each HRSG section the
mass conservation equations, the energy conservation equation, the gas side effectiveness definition for each HRSG section and finally interconnection equations. The number of equations
(NE) is always lower than the number of variables (NV) but the real independent variables of the
optimum design problem is sensibly lower than NV. This last varies in the range between 21 for
the RANSH HRSG and 86 for the 3PRSH HRSG. It is of primary importance to define for each
configuration the number of independent variables (NIV) of the optimization process. For HRSGs
with a number of pressure levels lower than or equal to 3, the number of independent variables
of the optimum design is given by the equation
NIV NVNE 3n1 w2 2w3(v g z y mw),

(7)

where n is the number of HRSG sections, w2 is the number of sections with two water streams,
w3 is the number of sections with three water streams, v is the number of evaporators, g are the
connections between the water streams of non-contiguous elements, y is the number of mass
conservation equations on the water side (conditions mk=0), z is the number of mass conservation
equations on the gas side (conditions Mk=0), mw is the number of congruence conditions (e.g.
the coupling of two heat recovery cycles).
In each case, the saturation temperatures and the gas side effectiveness of each evaporator are
considered as independent variables; other independent variables are the water outlet temperatures.
Let us now examine two particular HRSG configurations: the 2PRSH, represented in Fig. 5
and in Fig. 7(a), and the 3PRSH, represented in Fig. 6 and in Fig. 7(b) (for a simplified case
with three parallel sections).
In the schemes of Figs. 5 and 6 the elements referred as P1, P2 and P3 in Fig. 5 and P1, P2,
P3, P4 and P5 in Fig. 6 are sections with two or three parallel water streams. Table 1 provides

398

C. Casarosa et al. / Energy 29 (2004) 389414

Fig. 5. Scheme of a two-pressure level HRSG with reheater (2PRSH).

Fig. 6. Scheme of a three-pressure level HRSG with reheater (3PRSH).

Fig. 7. Thermodynamic vapor cycle related to 2PRSH (a) and 3PRSH (b) HRSG configurations.

C. Casarosa et al. / Energy 29 (2004) 389414

399

Table 1
Number of independent variables for the examined HRSG configurations
HRSG configuration

w2

w3

mw

NIV

RANSH
RAN+1P
RAN+2P
2PR (coup.)
2PR (unc.)
2PRSH
3PRSH

3
4
4
5
5
7
9

0
1
2
0
0
3
4

0
0
0
0
0
0
1

1
1
1
2
2
2
3

0
1
2
1
1
3
5

2
3
3
4
4
6
8

2
3
3
2
2
5
7

0
0
0
1
0
1
2

3
4
4
4
5
6
7

the values of the number of independent variables of the optimum design process for seven different HRSG configurations. After the identification of the number of the optimum design independent variables and after their selection, the optimization process consists in the minimization
of the exergy loss I defined by Eq. (2). The exergy losses I can be transformed and given in
dimensionless form with respect to a conventional enthalpy term. For convenience this reference
term can be obtained as the product of the mass flow of the inlet gas to the HRSG, Mg, with the
environmental temperature ta and with the specific heat of the gas at this temperature cpa. Consequently the function to be minimized is represented by the dimensionless exergy losses I
defined as:
I

I
.
Mgcpata

(8)

In general, steam saturation temperatures, superheated steam temperatures and evaporators effectiveness can be selected as independent variables for the thermodynamic optimization, while the
inlet temperature of the exhaust gas to the HRSG is considered as a parameter. A sensitivity
analysis is performed to evaluate the effects of its variation on the optimum design process.
4. Thermoeconomic optimization
It is quite easy to expect that all the solutions obtained by exergy loss minimization furnish
null pinch-point and infinite surface. The results obtained, even if meaningless from a technical
point of view (solutions involving null pinch points with infinite surfaces and costs of the HRSG),
represent a first rough selection criterion for the definition of the HRSG operating parameters. In
order to find a compromise between high thermodynamic efficiency and low cost of the HRSG,
a further analysis development is represented by the thermoeconomic optimization.
The thermoeconomic analysis carried out in the present work is based on the assumption that
the total cost of the HRSG is equal to the sum of the costs related to the exergy losses plus
the operating costs (related to installation and operation) [12]. The thermoeconomic optimization
considers as objective function the minimization of the above defined total HRSG cost. In the
proposed method, the key element for the thermoeconomic optimization is represented by the
definition of the cost of the exergy losses. To destroy exergy means to destroy available mechan-

400

C. Casarosa et al. / Energy 29 (2004) 389414

ical energy that must be computed at a cost sensibly higher than the fuel one. It is simple to
understand that accepting this point of view, the HRSG pinch-point values are a direct consequence of the cost assumed for the exergy losses.
4.1. Definition of a cost structure
The costs related to exergy losses have to be recast to operative costs of the HRSG and combined with the installation costs to obtain the total cost of the HRSG. For this aim the definition
of a true cost structure is of primary importance.
4.1.1. Cost of the exergy losses
The cost of the exergy losses can be expressed in the form
KI kIMgcpataHI.

(9)

In Eq. (9), kI represents the specific cost of the exergy losses, cpa is the reference specific heat
of the exhaust gas, at the environmental temperature, and H is the functioning duration of the
plant. For the definition of the specific cost of the exergy loss kI various strategies can be assumed.
To consider it as the cost of the fuel, but this seems to be the less realistic one.
To consider the cost of the exergy losses as the cost of the fuel divided for the efficiency of
the plant (in the case of the combined plants typically about 0.55).
To consider that exergy losses correspond to a lower energy availability, and they are equivalent
to the cost of the fuel divided for the average efficiency of the installed plants (typically of the
order of 0.350.40).
Another possibility, referred to in the examples carried out in the paper, is to consider the
exergy losses equal to an average value of the selling price of the electrical energy. This last
option derives from the consideration that an exergy loss in the HRSG corresponds to a lower
output of the plant and to a lower amount of energy that can be sold.
4.1.2. Cost of the HRSG sections
While the definition of the cost of the exergy losses and the data required for its evaluation
are quite simple to find, the definition of a HRSG cost structure is more complex. A simple
structure that can be proposed is that the cost of the single HRSG section must be proportional
to the surface, as well as the total cost of the HRSG must be equal to the sum of the costs of its
sections, so that for an HRSG composed of n sections

KHRSG

Kk.

(10)

k1

The cost of the single HRSG section can be expressed in the form
Kk ksSexp

(11)

where ks is the specific cost of the surface of the single HRSG section, a function of pressure
and temperature, S is the heat exchange surface and exp an opportune exponent. Concerning the

C. Casarosa et al. / Energy 29 (2004) 389414

401

cost of the single HRSG section, the main problem is the correct value to attribute to the specific
cost ks and to the exponent exp.
A first hypothesis states in considering the cost of the HRSG section directly proportional to
the surface (exp=1), being this protective with respect to the position usually suggested in the
literature stating that the exponent is equal to 0.8. Considering the general HRSG configuration,
it is possible to write that the total cost is equal to the sum of the costs of the various sections.
Four different kinds of sections are distinguished: economisers, evaporators, superheaters and
reheaters, so that
KHRSG

keSe

kvSv

kshSsh

sh

krhSrh.

(12)

rh

4.2. The HRSG total cost function


After the separate definition of the costs of the exergy losses and of the HRSG sections, the
total annualized cost of the HRSG to be minimized is expressed in the form
K KI KHRSG kIMgcpataHI

1
(
D

keSe

kvSv

kshSsh

sh

krhSrh)

(13)

rh

where D is the economic life of the plant. By means of the definition of the number of gas side
transfer units, Nk, the heat exchange surface S can be written for each one of the k sections in
the form
Sk

cpkMgNk
.
Uk

(14)

The total cost of the HRSG, can be expressed in dimensionless form homogeneously with I
defined by Eq. (8) as a function of the various N:

K I b

cpe
ke Ne
cpa

cpv
N
cpa v

sh

cpsh
ksh Nsh
cpa

rh

cprh
krh Nrh
cpa

(15)

where

and

kv
,
kItaHDUv

ke

k e Uv
,
kvUe

(17a)

ksh

ksh Uv
,
kv Ush

(17b)

(16)

402

C. Casarosa et al. / Energy 29 (2004) 389414

krh

krh Uv
,
kv Urh

(17c)

so that K represents the objective function to be minimized to have the thermoeconomic optimization and the structure of Eq. (15) permits a direct comparison between the cost of the exergy
losses and the cost of the sections of the HRSG. The thermoeconomic optimization will select
the most economic configuration and set of operating parameters, relative to the cost structure
assumed. These parameters will then constitute the input data for a more detailed design procedure
where the geometric variables of the HRSG sections can also be considered in a detailed optimization also concerning the minimization of exergy loss due to the pressure drop.

5. Mathematical method to solve the optimization problem


From a mathematical point of view, the optimum design problem consists in finding the minimum of a function, [F(X)], represented alternatively by the exergy loss I, defined by Eq. (8),
or by the total annual HRSG cost K, defined by Eq. (15), where X is a vector containing the
NIV independent variables. It is the classical problem of constrained minimization. The aim of
the mathematical procedure of optimization, after the appropriate selection of the vector X, is to
define a combination of the variables so that a certain value of the vector X will minimize the
objective function. The minimization of the objective function must occur simultaneously with
respect to the full set of equations describing the HRSG as well as the conditions that stand for
the physical sense of all the variables (e.g. temperature and mass flows greater than zero). This
problem of non-linear constrained minimization can be solved with one of the various methods
available.
Using the definition of the heat transfer effectiveness, given by Eq. (3), the problem has been
transformed in a quasi-linear constrained optimization problem. Just in some simple cases
(CARNOT, RAN and RANSH HRSG configurations) the problem can be solved analytically. In
general it has been solved by means of the Simplex method (in the version called the Complex
method), appropriately implemented on a computational code. For a detailed treatment of the
aforesaid Simplex (Complex) method, the reader is referred to [13]. The main advantages in using
the Simplex method are that it does not require:
the evaluation of the derivatives of the function F;
the function F to be continuous;
to know the dominion of existence of F.
A first step in the optimization consists in the definition by a random search procedure of an
initial vector of the variables X where the function to be minimized assumes the minimum value
among the various values obtained. Then, starting from this initial vector, the final optimization
is obtained by means of the direct application of the Simplex method.

C. Casarosa et al. / Energy 29 (2004) 389414

403

6. Results and discussion


6.1. Results of the thermodynamic optimization
By applying the above-mentioned thermodynamic optimization method to the single or multipressure evaporator, a simple criterion for the selection of the saturation temperatures ts as a
function of inlet temperature of the exhaust gas to the HRSG, and environmental temperature can
be given.
Using the definition of the evaporator effectiveness given in Eq. (3), where the dimensionless
exergy losses I, for the single pressure evaporator, can be analytically obtained as shown in
[12]. Minimizing the expression of the exergy loss with reference to the saturation temperature
tv, the result hv=1 (i.e. null pinch-point) can be obtained and the optimum saturation temperature
is given by:
tv,opt Tginta / Tgin.

(18)

The same optimization method applied to a multi-pressure HRSG obtained with a number n of
evaporators furnished for the ith level:
tv,i Tgin(ta / Tgin)i/(n+1),

(19)

with n the number of pressure levels. Assuming as reference environmental temperature ta=293
K, the results of Table 2 for different inlet temperature of the exhaust gas to the HRSG and
independently on the operating fluid, are obtained. By applying the thermodynamic optimization to
the configuration with economizer and evaporator (RAN), economiser, evaporator and superheater
(RANSH), writing a more complex function, an analogous result can be obtained. The minimum
of the exergy losses always occurs for null pinch-point values (hv=1). The optimization method
has been used to define the optimal operating parameters of the HRSG configurations enumerated
in 3.1.
In the Tables 36 the results of the optimization are shown for the various analyzed configurations, obtained assuming the following parameters:
reference mass flow of gas, Mg=386.7 kg/s;
reference ambient temperature, ta=293 K;
Table 2
Temperatures of the optimum Carnot HRSG
1 Carnot

2 Carnot

3 Carnot

Tgin[K]

ts [K]

ts1 [K]

ts2 [K]

ts1 [K]

ts2 [K]

ts3 [K]

700
773
823
850

453
476
491
499

523
561
585
598

391
407
416
421

563
606
636
647

453
475
491
499

364
373
379
382

404

C. Casarosa et al. / Energy 29 (2004) 389414

Table 3
Minimum inlet gas temperature to obtain critical conditions for the steam (ta=293 K and tlin=313 K)
HRSG configuration

Tgin for critical condition [K]

RANSH
RANSH+1P
RANSH+2P
2PR coupled
2PR uncoupled
2PRSH
3PRSH

823
790
773
830
820
725
760

Table 4
Optimum operating parameters for a 2PR HRSG
HP

LP

Tgin [K]

tv[K]

pv [bar]

tsh[K]

m [kg/s] x

tv [K]

pv [bar]

m[kg/s] x

700
725
750
773
800
810
820
823

0.08926
0.09053
0.08872
0.08428
0.07322
0.06524
0.04802
0.03064

555.5
585.7
609.2
621.2
635.3
640.4
645.6
647.0

66.5
102.2
139.0
161.5
191.5
203.8
216.7
220.4

700
704.7
750.0
773
800
810
820
822.8

31.91
34.76
36.99
41.35
47.64
50.72
55.49
58.95

412.9
425.5
433.4
434.8
431.1
426.0
410.4
392.7

3.585
5.066
6.216
6.440
5.873
5.131
3.344
1.959

18.44
19.73
20.44
19.25
16.67
14.96
11.75
9.01

0.831
0.8
0.8
0.8
0.8
0.8
0.8
0.8

0.8513
0.8384
0.8307
0.8293
0.8328
0.8379
0.8539
0.8738

Table 5
Dimensionless exergy losses for the optimized HRSG in the various configurations
RANSH

RAN+1P

RAN+2P

2PR (cou.)

2PR (unc.)

2PRSH

3PRSH

Tgin [K]

700
750
773
800
823

0.15632
0.17235
0.17487
0.17474
0.09069

0.14935
0.16513
0.17121
0.13211
0.07464

0.14466
0.15613
0.13892
0.11626
0.08866

0.09584
0.09808
0.09605
0.08751
0.07214

0.08926
0.08872
0.08428
0.07322
0.03064

0.08635
0.07271
0.05806
0.05467
0.04321

0.05974
0.05978
0.04892
0.04121
0.03025

liquid inlet temperature and condensation temperature, tlin=313 K;


steam turbine isoentropic efficiency=90%;
minimum value of steam quality at the end of expansion, x=0.8;
specific heat of gas at environmental temperature, cpa=1.0645 kJ/kg K

C. Casarosa et al. / Energy 29 (2004) 389414

405

Table 6
Dimensionless total exergy losses for HRSG and bottoming cycle
RANSH

RAN+1P

RAN+2P

2PR (cou.)

2PR (unc.)

2PRSH

3PRSH

Tgin [K]

I+Iesp+Exr

I+Iesp+Exr

I+Iesp+Exr

I+Iesp+Exr

I+Iesp+Exr

I+Iesp+Exr

I+Iesp+Exr

700
750
773
800
823

0.2354
0.2649
0.2771
0.2858
0.2152

0.2272
0.2549
0.2668
0.2410
0.2008

0.2225
0.2471
0.2298
0.2191
0.2037

0.1990
0.2186
0.2239
0.2235
0.2181

0.1807
0.1957
0.1995
0.1989
0.1691

0.1722
0.1841
0.1670
0.1718
0.1663

0.1485
0.1611
0.1531
0.1542
0.1521

Table 3 provides the inlet temperature of the exhaust gas to the HRSG, determining a critical
value for the high pressure level (assumed in our calculation as 220 bar and 647 K). With HRSG
composed of evaporators only, the aforesaid critical value of the water saturation temperature is
obtained only with three pressure levels and an inlet temperature of the exhaust gas of 850 K
(Table 2). On the contrary, with the other configurations examined, the inlet temperature Tgin, for
which the HRSG high pressure, approximates the critical value of 220 bar, is lower than 850 K.
Table 4 provides the optimized operating parameters of the HRSG and the corresponding
dimensionless exergy losses I for the case of a simple two pressure HRSG (2PR), at different
values of the inlet gas temperature Tgin.
The results obtained by means of this analysis, though if only ideal, determining infinite heat
exchange surfaces, represents a possible first selection criterion for the operating parameters. In
the Tables 5 and 6 the results of the exergy losses relative to the whole analysis are reported. In
Table 5 only the irreversibility related to the HRSG is considered. Table 6 provides the term of
total exergy losses, considering also those due to the isentropic efficiency of the expansion and
the residual exergy of the steam at the exit of the turbine, both given in dimensionless form, with
reference to the term Mgcpata, as Iesp and Exr. In Tables 5 and 6 it seems interesting to underline
that for the aims of the optimization, in the greater part of the cases, it seems equivalent to
consider as objective function the minimization of the HRSG exergy losses I or of the total
exergy losses Itot obtained as the sum of the three terms I+Iesp+Exr.
6.2. Results of thermoeconomic optimization for various HRSG configurations
Although the thermodynamic optimization gives null pinch-points and unlimited heat exchange
surfaces, it permits the knowledge of an extreme point of the spectrum of the optimal solution
for the HRSG operating parameters. A thermoeconomic optimization, obtained with the minimization of the costs defined by Eq. (15), gives more realistic results for the HRSG design. To
understand the differences determined by the thermoeconomic optimization with respect to the
thermodynamic optimization, it is interesting to consider the case of the Rankine cycle for what
analytic solution is possible. For a simple Rankine cycle HRSG (with only an economizer and
an evaporator), assuming the simplified case in which the cost of the economizers ke and the cost
of the evaporators kv be equal as well as the specific heat of the gas and the overall heat transfer

406

C. Casarosa et al. / Energy 29 (2004) 389414

Fig. 8. Correlation between pinch-point and cost parameter b (a) and between pinch-point and saturation temperature
ts (b).

coefficient in the various sections remains constant, the total dimensionless cost can be
expressed as
K I b(Nv Ne).

(20)

Minimizing the previous expression analytically, using I defined by the authors [12], the pinchpoint value (P.P.) is a function of b, as well as the saturation temperature, as shown in Fig. 8(a)
for the particular values of gas and water inlet temperatures.
Assuming the values Tgin=700 K it is easy to obtain the results presented in Table 7. In this
table it is clearly shown that the pinch-point (P.P.) is a function of the economic parameter b,
and can vary over a wide range. From the previous considerations and the results of Table 7 and
Fig. 8(a), the economical meaning of the pinch-point appears evident, a concept sometimes forgotten in the literature, and it is clear how its value is strongly dependent on the economic scenario
and cannot be selected on the basis of practical considerations only.
In the next calculations, maintaining the parameters already considered for what concerns Mg,
Ta, tlin, x and steam turbine isoentropic efficiency, the following cost structure has been assumed.
As cost of the exergy losses the value
kI 0.068 $ / kWh
Table 7
Sensitivity analysis of the thermoeconomic optimization of a RANSH HRSG with respect to the parameters b
b

pv [bar]

tv [K]

he

hv

hsh

P.P. [K]

0.001
0.01
0.03
0.05

15
8
7
7

472.6
442.4
436.8
436.8

0.407
0.374
0.288
0.239

0.999
0.946
0.835
0.753

0.206
0.196
0.168
0.145

0
11
36
55

C. Casarosa et al. / Energy 29 (2004) 389414

407

that represents an average value for the selling price of the electrical energy in Italy has been
adopted. The average costs of the four different kind of sections, extrapolated by the data of a
manufacturer, [14], are:
Evaporators, kv=34.9 $/m2,
Economizers, ke=45.7 $/m2,
Reheaters, krh=56.2 $/m2,
Superheaters, ksh=96.2 $/m2.
With the previously defined structure, the effect of temperature and pressure are considered in
the different costs of the sections, so the cost of the superheater sections is sensibly higher than
those of the reheaters, mainly for the different pressure, higher for the superheaters. Due to the
fact that the approach point influences only the allocation of the surfaces on the HRSG sections
(economizer and evaporator), its value has been assumed null for simplicity. For what concerns
the overall heat transfer coefficient the following values have been considered [14]:
Ue=42.6 W/m2 K,
Uv= 43.7 W/m2 K,
Ush=Urh= 50 W/m2 K.
In addition, the following values have been assumed for the parameters H and D:
functioning duration of the plant, H=8000 hour/year,
economic life of the plant, D=10 years.
With reference to the afore-mentioned values, the order of magnitude of the dimensionless parameter b is about 0.5103.
The aforesaid optimum design method with the cost structure defined in the previous paragraph
was applied to all the HRSG configurations. In particular, in Tables 8 and 9 the detailed results
of the HRSG thermoeconomic optimization in the two particular cases of the configurations
described in Figs 5 and 6 are reported, for three different values of the inlet temperature of the
Table 8
Optimized operating parameters for a thermoeconomically optimized 2PRSH HRSG
HP

LP

Tgin[K]

tv [K]

pv[bar]

tsh[K]

m [kg/s] tv [K]

pv[bar]

tsh[K]

m [kg/s]

700

0.09847

0.08680

0.07323

0.08635
0.08818
0.05806
0.06639
0.04321
0.04837

549.8
554.3
645.7
644.6
645.0
645.3

61.0
65.3
217.0
214.1
215.2
215.9

674.0
672.6
746.5
743.7
807.9
799.9

29.52
29.04
40.40
39.29
45.90
46.02

2.62
2.50
2.78
3.22
1.33
1.77

483.5
479.4
410.1
434.0
462.0
490.0

16.39
16.94
16.10
16.03
12.35
11.34

773
823

402.1
400.6
404.2
434.0
380.9
389.6

408

C. Casarosa et al. / Energy 29 (2004) 389414

Table 9
Optimized operating parameters for a thermoeconomically optimized 3PRSH HRSG
HP

IP

LP

Tgin
[K]

tv [K] pv
[bar]

tsh[K] m
tv [K] pv
[kg/s]
[bar]

m
tv [K] pv
tsh [K]
[kg/s]
[bar]

m
[kg/s]

700

0.07571

0.07593

0.06254

0.05974
0.06330
0.04892
0.05907
0.03025
0.03716

564.3
561.5
646.9
629.4
647.0
646.9

699.5
695.4
772.8
766.9
823.0
807.6

12.46
12.15
3.78
10.21
1.19
2.89

8.04
8.10
9.50
9.22
9.05
8.82

773
823

75.7
72.6
220.3
178.5
220.4
220.1

25.34
25.55
37.50
32.90
45.58
45.02

449.8
448.6
511.5
490.6
476.8
486.5

9.26
9.02
32.52
22.07
16.78
20.40

374.3
374.0
418.6
399.1
374.3
386.3

1.06
1.04
4.21
2.39
1.06
1.59

450.0
448.6
511.7
490.6
476.8
486.5

exhaust gas. In each case, the results of the thermoeconomic optimization are compared with
those of the thermodynamic optimization (identified by K=). In general, the results obtained
with the thermodynamic optimization have been substantially modified by the introduction of
economic elements. For each HRSG configuration, positive pinch-point values have been obtained,
as described in the Table 10.
As already stated the values of the pinch-points depends on the cost hypothesis; due to the
approach considered it is, however, possible to use different cost structures so that the method is
completely general. The cost hypothesis used in the paper to test the method is relative to a case
in which the cost of the exergy losses is high (it is equal to the selling price of energy). Consequently a low value of the parameter b results, so that the pinch-points obtained could seem to
be surprisingly low.
Moreover, Tables 11 and 12 provide the costs of the optimized HRSG configurations and the
sum of the costs of the optimized HRSG plus the cost related to the irreversibility related to the
steam expansion and of the residual exergy of the vapor at the end of the expansion. The results
of Tables 11 and 12 are the homologues of those reported in Tables 5 and 6 in the case of the
thermodynamic optimization. Figure 9 summarizes the results obtained in the thermoeconomic
optimization. In addition to those results Table 13 provides the surfaces of the optimized HRSG
configuration in the two meaningful cases of 2PRSH and 3PRSH illustrated in Figs 5 and 6.
Figure 9(a) provides, for the various analyzed HRSG configurations, a graphical analysis of
Table 10
Pinch-point obtained for optimized HRSG configurations
RANSH

RAN+1P

RAN+2P

2PR (coup.)

2PR (unc.)

2PRSH

Tgin [K] P.P. [K]

P.P. [K]

P.P. [K]

P.P.
H.P.

[K]
L.P.

P.P.
H.P.

[K]
L.P.

P.P.
H.P.

[K]
L.P.

P.P.
H.P.

[K]
I.P.

L.P.

1.7
2.2
5.5

0.9
0.9
1.0

1.2
1.3
1.8

1.1
1.3
1.9

1.3
3.1
4.7

1.1
2.4
2.1

1.7
2.1
9.0

2.1
2.9
2.6

1.3
1.2
2.6

700
773
823

0.6
0.7
0.3

0.6
0.6
0.6

0.7
3.7
7.1

3PRSH

C. Casarosa et al. / Energy 29 (2004) 389414

409

Table 11
Dimensionless cost of the optimized HRSG configurations
RANSH

RAN+1P

RAN+2P

2PR (cou.)

2PR (un.)

2PRSH

3PRSH

Tgin [K]

700
773
823

0.16423
0.18430
0.14330

0.15741
0.18115
0.10643

0.15218
0.17232
0.13552

0.10725
0.11094
0.09599

0.10155
0.10184
0.07894

0.09847
0.08680
0.07323

0.07571
0.07593
0.06254

Table 12
Dimensionless total costs for HRSG and steam cycle
RANSH

RAN+1P

RAN+2P

2PR (cou.)

2PR (un.)

2PRSH

3PRSH

Tgin [K]

K+Iesp+Exr

K+Iesp+Exr

K+Iesp+Exr

K+Iesp+Exr

K+Iesp+Exr

K+Iesp+Exr

K+Iesp+Exr

700
773
823

0.24370
0.28621
0.26417

0.23390
0.27670
0.22967

0.22997
0.26252
0.24836

0.21006
0.23906
0.24193

0.19251
0.21671
0.21517

0.18466
0.19519
0.19359

0.16416
0.18221
0.18414

the transformation of the exergy available in hot gas in the four dimensionless quantities W, I,
Iesp, Exr, power, exergy losses in the HRSG, exergy losses in the steam expansion and the residual
exergy of the steam, all in homogeneous dimensionless form, respectively. Figure 9(b) provides
a comparison between the total dimensionless cost of the HRSG, K, and the cost of the exergy
losses, I. The graphics of Fig. 9(a) and (b) are relative to three values of the inlet temperature
of the exhaust gas: Tgin=700 K, 773 K and 823 K. From Fig. 9(a) it clearly appears how the
configurations 2PRSH and 3PRSH permit us to reduce in a very sensible way the HRSG exergy
losses I, that are of the same order of magnitude as the exergy losses in the steam turbine Iesp,
especially for the higher value of the inlet gas temperature Tgin. So it appears evident how
important it will be to work also on the increase of the isoentropic efficiency of the steam turbine
at the same time as on the reduction of the HRSG exergy losses to increase the combined cycle
efficiency. Furthermore, from Fig. 9(b) it is possible to understand how, with the considered
structure, the cost of the exergy losses represent a significant amount of the total.
6.3. Application of the thermoeconomic optimization of the HRSG to existing plants
The optimization procedures were run on the 2PRSH and 3PRSH HRSG configurations also
using the same input data that are characteristic of two existing plants that use similar HRSG
configurations. The plants are produced by ABB-Alstom for a 280 MW utility (the KA24-1 ICS
combined cycle power plant based on the GT24 gas turbine) and Siemens for a 380 MW utility
(based on the Siemens V94.3a gas turbine). The input data for the HRSG are the mass flow
characteristic at the exhaust of the two turbines, respectively

410

C. Casarosa et al. / Energy 29 (2004) 389414

Fig. 9. Destination of the initial exhaust gas exergy (a) and comparison of exergy losses and dimensionless HRSG
cost (b) for different configurations and inlet gas temperatures Tgin.

C. Casarosa et al. / Energy 29 (2004) 389414

411

Table 13
Surfaces relative to the thermoeconomically optimized HRSG configurations (Tgin=823 K)

ECO
EVA
EVA
ECO
EVA
P1
P2
P3
P4
P5

LP
LP
IP
HP
HP

2PRSH

3PRSH

33777.6 m2
29162.3 m2

234.3 m2
10061.4 m2
18062.2 m2 (ECO) 1555.1 m2 (RH)
130160.4 m2 (ECO) 39520.6 m2 (RH)
48702.9 m2 (ECO) 14657.1 m2 (RH)

40999.1 m2
33761.8 m2
20001.2 m2

10215.9 m2
34573.3 m2 (ECO) 2501.7 m2 (RH)
not present in the optimized solution
not present in the optimized solution
100011.3 m2 (ECO) 2761.7 m2 (RH)
52855.9 m2 (ECO) 17252.7 m2 (RH)

Mg=386.7 kg/s, Tgin=920 K, for the GT24 gas turbine;


Mg=653.1 kg/s, Tgin=853 K, for the V94.3a gas turbine.
The condensation temperature in both cases is 305 K.
The HRSG configurations adopted and optimized here are quite different from those really
existing, above all for the use of the parallel type exchanger sections anywhere it is possible.
Both the thermodynamic and thermoeconomic optimization lead to a meaningful reduction of
the exergy losses, an increase of the value of the HRSG high pressure and to a general reduction
of the pinch-points. It can be supposed that all the other differences that arise from the comparison
(above all, the pressures and the flow rates in the IP and LP sections) are a consequence of the
primary characteristics of accepting the increase of the high pressure to a value close to the critical
one. This possibility deeply influences the properties of the thermal cycles, with respect to the
usually adopted configurations, and shows how the exergy losses in the HRSG become similar
to those relative to the non-isentropic expansion in the steam turbine.
In general, the thermoeconomic optimization seems to promise an increase of the power output
of about 9% for the 2PRSH cycle (the ABB-Alstom plant) and of 11% for the 3PRSH one (the
Siemens plant), with respect to their real applications. Thus the overall efficiencies of the plants
can be brought from 57.6%, declared by the manufacturer, to 59.5% in the former case, and from
57.1%, declared by the manufacturer, to 59.4% in the latter one [15]. Table 14 provides the results
of the two optimizations for the second case (the Siemens plant).
This result shows the possibility of breaking the 60% efficiency barrier, as sought by all the
manufacturers, only by best fitting the existing technology, though with a rise in the HRSG investment.
7. Conclusions
In the paper a thermoeconomic optimization of the operating parameters of the heat recovery
steam generator (HRSG), for combined cycle plants, is performed. The method is an alternative

412

C. Casarosa et al. / Energy 29 (2004) 389414

Table 14
Application of the optimization methodology to a commercial plant configuration (case Siemens, Mg=653.1 kg/s,
Tgin=853 K, tlin=305 K)
HP
K

REAL
TMD
TME

p
tsh
[bar] [K]

117

0.03412 217
0.07712 0.05882 216

813
852
835

IP

LP

m
P
tsh
[kg/s] [bar] [K]

m
P
tsh
[kg/s] [bar] [K]

m
LP
[kg/s] [K]

IP
[K]

HP
[K]

65.7
79.8
76.4

22.1
4.0
8.2

8.9
14.3
16.3

4.2

8.5

131

150
14.6 146

15.2 573
17.4 502
28.2 542

3.7
1.2
2.2

PP

615
503
504

Wst
[MW]

HRSG cost calculated after the application of the optimization procedure: 20.3 M$; HRSG cost in real conditions:
10 M$.

to the usual pinch-point one, largely used in practical applications. It represents an attempt to
find a compromise between economic and thermodynamic analysis, based on incorporating
exergy-based production costs with economic evaluations.
The analysis is based on the gas-side effectiveness of the sections of the HRSG instead of the
usual pinch-point method. Simple configurations such as a single pressure evaporator and a more
realistic one (like double and triple pressure HRSGs with reheat sections) have been considered.
The thermodynamic optimization, carried out by means of the minimisation of exergy losses,
taking into account only the irreversibility due to the temperature difference between the hot and
the cold stream, determines solutions with null pinch-point and infinite surfaces and gives indications based on the maximum gas temperature to obtain the critical value of the steam pressure
and temperature.
It is shown how, using water as the fluid for the bottoming cycle, it is not convenient for an
unconditional increase of the inlet temperature of the exhaust gas to the HRSG, because the
efficiency of the steam bottoming cycle shows an asymptotic trend and it appears reasonable to
assume an upper limit value for the inlet temperature of the exhaust gas to the HRSG. The trend
of the HRSG optimization leads to the supercritical cycle for the inlet temperature of the gas to
the HRSG, higher than 850 K. The results of the thermodynamic optimization clearly demonstrate
that the use of the steam water HRSG cycle in bottoming applications has drawbacks due to the
constant temperature of isobaric vaporization and to the low critical value of water, equal to 647 K.
The economical considerations change the results obtained with only the minimisation of exergy
losses and, depending on the HRSG operating costs, the pinch-points will assume values different
from the null one, in the range between 0.3 and 9 K depending on the HRSG configuration. By
applying the optimum design strategy pointed out in the paper, it would seem possible to obtain
combined cycle power plants with efficiency very close to 60%.
Summarizing:
The optimization method proposed can be used instead of that based on the aprioristic choice
of the pinch-points, for the design of a HRSG, as their values are the result of an optimization process;
The use of a heat exchanger with more than one parallel water streams, together with the shift

C. Casarosa et al. / Energy 29 (2004) 389414

413

of the high pressure level towards the critical condition, is a key element to obtain sensible
increases in the exergetic efficiency;
Thermodynamic analysis provides null pinch-points (i.e. infinite evaporators surfaces), as
expected;
The pinch-points rise when economic analysis is carried out, proving that their choice is the
consequence of a compromise between thermodynamic efficiency and investment costs;
The values of the pinch-points depend on the cost hypothesis done; due to the approach adopted
it is however possible to consider different cost functions, so that the method is completely general;
Beyond certain values of the inlet temperature of the exhaust gas to the HRSG (different among
the cycles), the higher pressure of steam trends upwards close to the critical condition;
With this method it seems possible to reach overall combined cycle efficiencies close to 60%
on existing plants, just by optimizing the heat recovery and the steam cycle operating parameters,
without modifying the gas turbine characteristics or searching for an increase in the steam turbine
isentropic efficiency;
Further developments of gas turbine technology joined with the proposed HRSG optimization
method can well lead to combined plant efficiencies higher than 60%.

Acknowledgements
The authors would like to acknowledge Dr Ing. Alessandro Russo for his valid collaboration
during the development of this research work.

References
[1] Horlock JH. Combined power plantspast, present, and future. ASME Journal of Engineering for Gas Turbines
and Power 1995;117:60816.
[2] Deschamps PJ. Advanced combined cycle alternatives with the latest gas turbines. ASME Journal of Engineering
for Gas Turbines and Power 1998;120:3507.
[3] Jericha H, Fesharaki M, Seyr A. Multiple Evaporation Steam Bottoming Cycle. ASME Paper 97-GT-287, 1997.
[4] Schultz R, Bachmann R. KA24-1CSTMarket success for a standardized power plant. ABB Internal Report
M489, 1999.
[5] Korobitsyn MA. New and advanced energy conversion technologies. Analysis of cogeneration, combined and
integrated cycles. Ph.D. Thesis, University of Twente, Enschede, The Netherlands, 1998.
[6] Deschamps PJ. Incremental cost optimization of heat recovery steam generators. In: Cogen-Turbo Power Conference, Vienna, Austria, CTP-11, ASME, 1995.
[7] Nag PK, De S. Design and operation of a heat recovery steam generator with minimum irreversibility. Applied
Thermal Engineering 1997;17(4):38591.
[8] Valdes M, Rapun JL. Optimization of heat recovery steam generators for combined cycle gas turbine power plants.
Applied Thermal Engineering 2001;21:114959.
[9] Bandyopadhay S, Bera NC, Bhattachara S. Thermoeconomic optimization of combined cycle power plants. Energy
Conversion and Management 2001;42:35971.
[10] Kotas TJ. The exergy method of thermal plant analysis. London: Butterworths, 1985.
[11] Kays WM, London AL. Compact heat exchangers. New York: McGraw-Hill, 1984.

414

C. Casarosa et al. / Energy 29 (2004) 389414

[12] Casarosa C, Franco A. Thermodynamic optimization of the operative parameters for the heat recovery in combined
plants. International Journal of Applied Thermodynamics 2001;4(1):4352.
[13] Rao SS. Engineering optimization. Chichester: John Wiley and Sons, 1996.
[14] Boyen JL, Waters MH. Use of the HXDSN program to size and estimate the cost of a heat recovery steam
generator. Report EPI TB-102, 2000 (on www.energyplan.com).
[15] Russo A. Ottimizzazione termoeconomica dei parametri operativi di generatori di vapore a recupero. Master Thesis, University of Pisa, Engineering Faculty (in Italian), 2001.

También podría gustarte