Anda di halaman 1dari 35

Accepted Manuscript

Thermo-economic optimization of heat recovery steam generator for a range of


gas turbine exhaust temperatures

Mahmoud Nadir, Adel Ghenaiet, Carlo Carcasci

PII: S1359-4311(16)30938-3
DOI: http://dx.doi.org/10.1016/j.applthermaleng.2016.06.035
Reference: ATE 8446

To appear in: Applied Thermal Engineering

Received Date: 28 April 2015


Revised Date: 21 April 2016
Accepted Date: 6 June 2016

Please cite this article as: M. Nadir, A. Ghenaiet, C. Carcasci, Thermo-economic optimization of heat recovery
steam generator for a range of gas turbine exhaust temperatures, Applied Thermal Engineering (2016), doi: http://
dx.doi.org/10.1016/j.applthermaleng.2016.06.035

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Thermo-economic optimization of heat recovery steam generator
for a range of gas turbine exhaust temperatures
1 2 3
Mahmoud NADIR , Adel GHENAIET and Carlo CARCASCI
1
Laboratory of Energetic Mechanics, Faculty of Engineering, University of Boumerdes, 35000, Boumerdes, Algeria
mahmoudnadir82@hotmail.com
2
Laboratory of Energetics and Conversion Systems, Faculty of Mechanical Engineering, University of Sciences and
Technology, Houari Boumediene, BP32 El-Alia, Bab-Ezzouar, 16111, Algiers, Algeria
ag1964@yahoo.com
3
Department of Industrial Engineering, University of Florence, Via Santa Marta, 3, Florence, Italy
carlo.carcasci@unifi.it

Abstract

This paper illustrates the effect of selling price on the optimum design parameters of a heat recovery steam generator

and selection of its ideal configuration for an outlet temperature range of 350°C - 650°C. The Particle Swarm

Optimization (PSO) method was used, considering the steam cycle specific work as an objective to be maximized, the

net present value as another objective to be maximized for the economic optimization and a combination of both of

them. Three configurations of heat recovery steam generators are considered with one, two and three pressure levels

with reheat. The results show that, the three pressure level system is the best configuration from a thermodynamic point

of view, but with respect to the economical aspect the two pressure levels is the best configuration for the low and

medium selling prices (0.04$/kWh, 0.08$/kWh and 0.2$/kWh), whereas the three pressure level configuration would

only be interesting for a high selling price of 0.3$/kWh and a temperature range 450-600°C. For a temperature of

650°C, the high cost of the three level system leads to net present value decrease. As the selling price increases the

optimized design parameters of the three pressure level HRSG based on an economic or a thermodynamic optimization

are similar. The obtained results are used to elaborate a new correlation relating the net present value with the gas

turbine outlet temperature, gas mass flow rate, number of levels of HRSG and the selling price.

Keywords: Combined cycle; HRSG; Exhaust gas temperature; Net present value; Thermo-economic optimization

Highlights

 Thermo-economic optimization of HRSG configurations

 The maximum value of the net present value was targeted for the economic optimization

 Three level HRSG is the best option in respect of power output and high priced medium

 Two level HRSG is the best for net benefit in low and intermediate priced mediums

1
1 INTRODUCTION

Today, combined cycles (CC) installations are considered as the favorite facilities for electricity power generation,

and their bottoming steam cycle provides about 30-40 % of the overall generated power [1]. Any improvement could be

done mainly through optimizing the heat recovery steam generator (HRSG) according to some manufacturing

specifications or costumers economic requirements, contrary to gas turbines (GT) which are mass manufactured with

imposed design parameters.

Many studies concentrated on CC steam part, started with the optimization of a single pressure level HRSG such as

Valdés and Rapuan [2]. Franco and Casarosa [3] and Franco and Russo [4] compared HRSG systems for one, two and

three pressure levels with and without reheat and demonstrated that it is possible to reach an overall efficiency of 60%

just by optimizing the steam cycle. Xiang and Chen [5] optimized a reheat three pressure levels HRSG both at a full and

a part load, and suggested using a partial recovery of the exhaust gas energy for a temperature exceeding 590°C.

Mohagheghi and Shayegan [6] optimized four types of HRSG of one, two and three pressure levels with a reheat

adapted for the GT Siemens V94.2 for which the turbine outlet temperature (TOT) is equal to 550°C, and showed that

by adding a pressure level always leads to improving the steam cycle performance. Bracco and Siri [7] considered

several objective functions when optimizing a single pressure level HRSG adapted to four types of GTs present on the

market, and outlined the influence of TOT and gas mass flow rate on the steam cycle performance. Kaviri et al. [8]

optimized HRSG with two pressure levels and a supplementary firing designed to work with the gas turbine GE9171E

having a TOT equal to 550°C. They fixed the optimized parameters and examined the effect of TOT variation on

HRSG, and found that by increasing TOT to a value of 650°C leads to an increase in the steam cycle thermal and

exergetic efficiencies, but above this temperature the thermodynamic performance improvement is lower. Bassily [9,

10] optimized the whole CC in which the steam cycle is reheated at two and three pressure levels, resulting in an

efficiency enhancement by 1.9 - 2.1 % as compared to the design case. Godoy et al. [11] optimized a simple steam CC

by maximizing its exergetic efficiency for a wide range of power with the determination of HRSG optimum surface. In

order to increase the specific work of GT and TOT (due to its direct impact on the steam cycle performance), some

studies suggested a reheated expansion and a blade cooling with the steam extracted from HRSG. Polyzakis et al. [12]

optimized and compared between simple, intercooled, reheated and intercooled-reheated GT when adapted to a simple

steam cycle, and concluded that the reheat is the most suitable solution. Khaliq and Kaushik [13] showed that the

reheated expansion of GT improves the CC global performance, especially its specific work. Chiesa and Macchi [14]

have demonstrated that an efficiency higher than 61% can be achieved when GT expansion is reheated and the turbine

blades are cooled with the steam extracted from the bottoming cycle using a “closed loop” configuration. Also Sanjay et

al. [15] have shown that with reheated expansion and steam turbine blade cooling, the thermal efficiency may reach 62

2
%. Boyaghchi and Molaie [16] have suggested an advanced exergy and environmental analyses and a multi objective

optimization of a CC plant with supplementary firing. Their results reveal that the optimum value of the total avoidable

exergy destruction rate and CO2 emission indicates, respectively, 10.6% and 8.3% improvements comparatively to the

basic case.

The above mentioned studies have only addressed the thermodynamic analysis and optimization, but the

maximization of performance may lead to a higher cost of investment and electricity [17, 18], and this is why other

authors have considered the economical aspect. Casarosa et al. [19] made a thermo-economic optimization for TOT

equal 700K, 773K and 823K, and have shown that the optimum pinch points are a consequence of a compromise

between thermodynamic efficiency and investment costs. Bassily [20] optimized an objective function of net additional

revenue for a triple pressure level reheat HRSG adapted to a GT reheated with exhaust gas recuperation. The

optimization resulted in an annual saving of 33.7 million USD for a 481MW power plant. Kotowicz and Bartela [21]

analyzed the influence of fuel price variation on the optimum values of the design parameters of the steam cycle parts,

and found that an increase in fuel price requires higher optimum pressure in both high and intermediate pressure parts

and lesser optimum pinch point and a higher optimum value of steam temperature in the intermediate pressure turbine.

Ahmadi and Dincer [22] studied the effect of fuel cost on the optimum design variables of CC and concluded that by

increasing the fuel price, the values of the decision variables in the thermo-economically optimum design tend to those

of the thermodynamically optimum design. Carapellucci and Giordano [23] undertook a thermo-economic optimization

of several types of HRSG adapted to three types of GT, and investigated the effect of fuel price and capacity factor on

the electricity cost. Rovira et al. [24] considered the frequent off-design operation of CC and developed a

thermoeconomic optimization model in order to minimize the electricity cost. Bakhshmand et al. [25] have optimized a

CC with supplementary firing leading to 3% decrement in the power plant specific cost and consequently the cost of

electricity. Rovira et al. [26] have optimized several configurations of CC using a partially recuperative gas turbine and

an HRSG of two and three pressure levels, among their main conclusions: the optimum recuperative mass fraction is

approximately 90% for both two and three pressure levels HRSG. By considering the exergo-economic and

environmental aspects, Vandani et al. [27] have shown that the natural gas fuel leads to better performance comparing

to diesel in operation of combined cycle power plants

A synthesis of this review shows that most of previous studies addressed the optimization of one or several HRSG

types intended to work with a specific gas turbine, but there is no study for both thermodynamic and economic aspects

considering the effect of electricity selling price on the optimum design parameters and selection of the best HRSG

configuration for a wide range of TOT. In fact, considering a range of TOT for several HRSGs and selling prices allows

3
generalizing the results that can be correlated thereafter to get the NPV in relation with the selling price, which may in

the case of preliminary. Accordingly, the main objectives of this study are as follows:

 A parametric analysis to illustrate the evolution of the net present value (NPV) according to the electricity selling

price, interest rate and HRSG design variables.

 A thermodynamic optimization of three HRSG configurations namely: HRSG with one pressure level with reheat

(1PRH), two pressure levels with reheat (2PRH) and three pressure levels with reheat (3PRH). The steam cycle

specific work is considered as the objective function for a TOT range 350°C-650°C.

 An economic optimization of the three configurations for several selling prices considering NPV as an objective

function

 A Thermo-economic optimization combining the thermodynamic objective function (steam cycle specific work)

with the economic objective function (NPV).

The highest TOTs which do not exceed 650°C are observed in the case of the most advanced gas turbines operating

under a high turbine inlet temperature (1500°-1600°C [28]). The following examples can be quoted: Westinghouse

501G (597°C) [21], Siemens SGT6-8000H (627°C) [29], GE S107FA (603°C) [30] and GE-P9371FB (644°C) [31]. For

relatively low TOTs the conventional steam cycles do not operate suitably and other cycles are preferred such as the

chemical recuperated gas turbine cycle (CRGT) and the organic Rankine cycle (ORC) [32], at a lower limit of 350°C is

chosen. In this study the range of TOT 350°C-650°C is chosen to cover the majority of gas turbines.

2 THERMODYNAMIC MODELING

An HRSG is used to recover the maximum of heat of a GT exhaust, through the optimization of the produced mass

flow, temperature and pressure of steam taking into consideration some constraints such maximum pressure and

temperature, minimal steam fraction at turbine outlet, minimum stack temperature and minimum pinch point. The

simplest HRSG configuration that can be designed is the one pressure level without re-heat which destroys a

considerable quantity of exergy. The transformation of one pressure level HRSG into two and three pressure levels

HRSG with reheat allows the diminution of destroyed exergy, the recovery of more heat quantity and the reduction of

exit gas temperature at the stack. However, by increasing the number of pressure levels increases the capital cost of

steam cycle, and in this context both thermodynamic and economic aspects have to be considered. The quantity of

steam to be produced and its temperature are more affected by the difference between the gas temperature leaving the

evaporator and the temperature of the water leaving the economizer, and such a difference is composed of the sum of

the approach point and the pinch point. The approach point is the difference between the temperature of saturated steam

4
and the temperature of water at the exit of the economizer and the pinch point is the difference between the gas

temperature leaving the evaporator and the temperature of saturated steam.

The schematics of the three HRSG configurations are presented in Fig. 1, their temperature-transferred heat

diagrams are given by Fig. 2. For an easy presentation, only the modeling of HRSG with three pressure levels and a

reheat is shown. Each heat exchanger is modeled by means of mass and energy balance that can also be explained

schematically by the Fig. 2.

The energy balance for the evaporator, reheater and superheater of the third level, separately one by one, leads to

the following equations respectively:

(1)

(2)

(3)

Through the sum on these three parts equations, the following expression is obtained:

(4)

The effectiveness of superheater and reheater is defined by:

(5)

(6)

The definition of pinch point of the third level:

(7)

The temperature of gas leaving the economizer of third level ( ) is obtained from energy balance:

(8)

Equations from (3) to (8) represent a system of 6 equations with 6 unknowns: and , that

are solved by using a numerical method. The properties of water and steam and the determination of saturation

temperatures are based on relations of the International Association of the Properties of Water and Steam (IAPWS)

[33].

Contrary to the third level pressure (where the equations are coupled), the determination of the unknowns of the

second level may be done explicitly. Knowing the steam temperature at evaporator exit ( wich is obtained

using the IAPWS relations), the steam temperature at the superheater exit is obtained from the definition of

effectiveness:

(9)

5
Gas temperature at the outlet of second level evaporator ( is obtained by using the second level pinch point

definition:

(10)

It is now possible to determine the ratio of steam of the second pressure level from the energy balance across both

evaporator and superheater:

(11)

The gas temperature at the outlet of economizer ( ) is obtained from the second level economizer energy balance:

(12)

For the first pressure level, similar relations as for the second level still apply. The exit temperature of superheater is

obtained from the definition of effectiveness:

(13)

Gas temperature at the exit of the first level evaporator ( is obtained from the first level pinch point definition :

(14)

The steam ratio of the first pressure level is determined from the energy balance across both evaporator and superheater:

(15)

The stack exit temperature ( ) is obtained from the first level economizer energy balance:

(16)

Once the system of equations is resolved, it is possible to determine the specific work produced by the three parts of the

steam turbines (LP, IP, HP) and thus the total steam cycle specific work per 1kg of exhaust gas:

(17)

and represent the ratios of steam for the first, second and third pressure level in relation to the gas mass flow

rate at the inlet of the HRSG.

In general terms, the steady state exergy balance applied for a given control volume is written as follows:

For 1kg of exhaust and considering a control volume that corresponds to the whole steam cycle and by neglecting the

exergy of heat transfer, equation (18) becomes:

(19)

The net exergy carried into the control volume:

(20)

6
The destroyed exergy can be deduced:

(21)

The destroyed exergy rate:

(22)

The validity of the proposed model is confirmed after comparing its results with the real data of three pressure levels

and reheat steam cycle built around the GE-PG9371FB [31]. This comparison is presented in table 1 showing that the

calculated values are close to the real data and the slight difference is due to other non-considered aspects by the present

model such as the auxiliary equipments.

3 ECONOMICAL ASPECTS

The net present value (NPV) is one of the most interesting criteria in the analysis of the economic performance

because it represents the updated net income along the life of the project. Its mathematical formulation [21] is as the

follows:

In equation (23) represents the average service life, is the annual interest rate and is the cost operation and

maintenance. The assumed values of these three parameters in the case of economic optimization are given in table 2.

is the capital charge factor defined as [34]:

The total annual income ( ) is as follows:

(25)

In equation (25), is the electricity selling price in USD per kWh, is the steam cycle produced power in kW

(obtained after a thermodynamic calculation) and is the yearly operating hours.

The cost of the steam cycle ( ) is obtained by the sum of its components for which correlations are given in the table

3 [23, 35].

4 OPTIMIZATION

For a given GT the values of TOT and the specific work are fixed, and hence the maximization of the energetic or

exergetic efficiency of CC requires only maximizing the specific work (per unit mass of exhaust gas) of the steam cycle

which is considered as the objective function. Concerning the economic optimization, the NPV is chosen as the

economic objective function that should be maximized. The pinch point of each pressure level, the superheater

7
effectiveness and the pressures at different levels are the optimization variables of HRSG. The steam ratio at each level,

the exit temperatures of superheaters and reheater in addition to the specific work and the NPV represent the

optimization results.

4.1 Constraints analyses

In order to avoid the mechanical and aerodynamic performance degradation of the last stages of the steam turbine,

the steam fraction should be higher than 88% [36]. Ganjeh Kaviri et al. [37] have shown that from a thermo-economic

and environmental point of view, the steam fraction at a turbine outlet should be higher than the value of 88%. The

steam pressure and the temperature at exit of HRSG should not exceed 160 bar and 580°C, in order to maintain a good

operation of HRSG and turbine material [38]. The condensed water in the exhaust gas could form a corrosive sulfuric-

acid, so the stack temperature should be higher than 80°C to avoid water condensation [39]. All these constraints are

taken into account as summarized below:

 Vapor fraction at exit of steam turbine should not drop below 88 %:

 Steam temperature at superheater exit or reheater should not exceed 580°C:

 Exhaust gases temperature must be higher than 80°C to avoid condensation and corrosion:

 Pressure of the third level should not exceed 160 bar:

 Pinch points:

 Superheaters and reheater effectiveness:

4.2 PSO Algorithm

PSO method proposed by Eberhart and Kennedy [40] is inspired from the ability of groups of animals’ species to

work as a whole, e.g. birds flocking to a food source, and this seeking behavior was associated with that of an

optimization search for solutions to non-linear equations in a real-valued search space [41]. PSO algorithm starts with a

population of solutions (taken randomly) and looks for an optimum for the problem, making population individuals

8
evolving over generations and the research for the best solutions by moving the individuals from their previous

positions ( ) to better one ( ) such as: . The role of the velocity ( ) is to adjust the particles

path towards their own previous best positions ( ), and towards the previous best position obtained by any member

of the group ( ), which is expressed as:

The performance of each particle is evaluated using the fitness function. In the case of the thermodynamic optimization,

the fitness function is the steam cycle specific work per kg of exhaust gas ( , while for the economic optimization

the fitness function is the NPV. More details concerning this algorithm can be found in the reference [42].

An example, the objective function of 3PRH at TOT=600°C, in the case of thermodynamic and economic

optimization, is shown (Fig. 3) to converge after about 40 iterations.

4.3 Thermo-economic optimization

In order to consider simultaneously both thermodynamic and economic optimization, a bi-objective

optimization is addressed in the present study. The two objective functions may be merged in a single one:

(26)

Since the orders of magnitude of the two objective functions are different, they were divided by their maximum

values ( and ) obtained through the maximization of each objective. By varying in the range 0-1

and maximizing , the Pareto front that gives a set of the best solutions is obtained. For instance, Fig.4 gives an

example of Pareto front for TOT=600°C and a selling price of 0.04$/kWh. The Pareto front can be used for the

determination of the equilibrium point, however in the present study an equitable influence between the two

objectives was adopted and the weight was taken as .

5 RESULTS AND DISCUSSION

5.1 Parametric analysis

This section presents a parametric study in order to understand the evolution of NPV according to the electricity

selling price, interest rate, HRSG design variables (pressure at HP turbine inlet, pinch points, superheaters and reheater

effectiveness). The TOT, the gas mass flow rate and the pressure at HP turbine inlet are fixed at 600°C and 650 kg/s and

160 bar respectively.

a. Evaluation of NPV depending on the pinch point:

Figure 5 shows NPV depending on the pinch point in the case of 3PRH configuration. The selling price and the

interest rate are considered as parameters. It is well understood that the decrease of the pinch point makes the
9
thermodynamic performance of the steam cycle increasing, but its impact on the net present value (NPV) is different. It

depends greatly on the selling price and the interest rate. In fact, the decrease of the pinch point must ensure a

compromise between the increase of the produced power and the over cost due to an increase in the exchange area

caused by the decrease of the pinch point. The decrease of interest rate and increase of the selling price cause the

reduction of the optimum pinch point. In the case of a relatively high selling price of 0.2$\kWh and a low interest rate

of 0.04 (Fig. 5c), the optimum pinch point maximizing the NPV tends to its maximum value which corresponds, in fact,

to the optimum thermodynamic pinch point. The increase of the interest rate till the value of 0.2 makes the optimum

pinch point increase to reach a value of about 12°C. The optimum pinch point in the case of low interest rates is lower

than the one of high interest rate values. For example, for a selling price of 0.08$\kWh, the optimum pinch point is

about 7°C for i = 0.04 and about 22°C for i = 0.2 (Fig. 5b). The decrease of selling price till 0.04$\kWh, as shown by

Fig. 5a, increases the optimum pinch point till 16°C for i = 0.04 and 31°C for i = 0.2. This is justified by the fact that the

increase of “sp” and the decrease of “i” increase the net present value (NPV) and the over cost caused by the decrease

of the pinch point is lower than the additional profit thanks to the surplus of power due to the same decrease of pinch

point fostering the reduction of the optimum pinch point, and vice-versa. For the low selling prices and high interest

rates, the profits decrease and consequently cannot overcome the over cost due to the reduction of pinch point that must

increase its optimum value in order to reduce its exchange area and then reduce the steam cycle cost and finally allow

the NPV to keep its maximum value.

b. Variation of NPV depending on superheater effectiveness:

For the configuration 3PRH, Fig. 6 shows that the optimum effectiveness of the superheater depends on the selling

price and the interest rate. It increases with the decrease of interest rate and the increase of selling price. In the case of a

combination between a low selling price and a high interest rate, the optimum effectiveness does not tend to the highest

value. As shown by Fig. 6a, if the selling price is 0.04$/kWhr and the interest rate is 0.2, the optimum effectiveness is

approximately 0.75. By decreasing the interest rate till 0.04 it increases the optimum effectiveness to 0.85. On the other

hand, by increasing the selling price also increases the optimum as shown by Figs.6b and 6c revealing that the

combination between the high selling price and the low interest rate makes the optimum effectiveness to reach its

maximum value.

c. NPV variation depending on inlet pressure of HP turbine:

The optimum pressure at the inlet of HP turbine in the case of 3PRH configuration is influenced by the selling price

and the interest rate, but only in the case of a combination between a low sp=0.04$/kWhr and a high interest rate of 0.2.

10
Figure 7a shows that in this case the optimum pressure takes a value of 120 bar. For the same interest rate (i= 0.2), by

increasing the selling price to sp=0.08$/kWhr, the optimum pressure is 140 bar (Fig. 7b). In the other cases, as shown

by Figs. 7b and 7c, the NPV increases when the pressure at the inlet of HP turbine increases.

5.2 Thermodynamic optimization

The optimum specific work obtained for three HRSG configurations is given by Fig. 8. As seen, for all TOT

values, adding a pressure level is always interesting and an HRSG configuration of 3PRH leads to high values of the

specific work than HRSG of 2PRH or 1PRH. This is due to the fact that by adding a pressure level this allows achieving

higher pressures, destroying less exergy and producing more steam quantities. This explanation is argued according to

Fig. 9 showing that at higher pressure level the optimum pressure becomes higher for a given TOT. For example, for

TOT=450°C for 3PRH, the pressure at inlet HP turbine reaches 92 bar, while for 2PRH it is 48 bar and only 22 bar for

1PRH. It is also shown from this figure that the optimum pressure at the inlet of HP turbine cannot reach the limit of

160 bar for given values of TOT: 580°C for 1PRH, 530°C for 2PRH and 500°C for 3PRH, because a low value of TOT

leads to a low superheated steam temperature and subsequently a low steam fraction which must remain above a value

of 0.88. For the values of TOT higher than these latter, the optimum inlet pressure at the HP turbine reaches its limit and

becomes constant due to imposed constraint on the maximum cycle pressure that must be under a value of 160bar as

previously mentioned in section 4.1. This result is practically justified by the fact that the maximum pressure of a steam

cycle is constrained by the steam fraction in the last stages of LP turbine which must be higher than 88 % for a given

inlet turbine temperature. In addition, the use of a high pressure leads to a low steam fraction at the end of expansion.

Consequently, there is recourse to reheat in order to increase the steam fraction (and the pressure at turbine inlet). On

other hand, Fig. 9 shows that the steam temperature after a reheat in 3PRH is higher than for 2PRH or1PRH, which

justifies the fact that for a given TOT, 3PRH pressure level allows reaching high pressures more than in the case of

2PRH or 1PRH. Also Fig. 9 gives the optimum total steam ratio obtained as a function of partial steam ratios of all

levels. HRSG with 3PRH allows producing more steam as compared with 2PRH or 1PRH, which is valid for all values

of TOT. For example, at TOT equal 600°C, the produced optimum total steam ratios are 13.3 %, 14.8 % and 16.4 % for

1PRH, 2PRH and 3PRH, respectively.

Another thermodynamic aspect in the favor of configuration 3PRH, concerns the destroyed exergy. The evolution of

destroyed exergy with TOT is illustrated by Fig. 10, showing that for an HRSG of 3PRH, the destroyed exergy is the

lowest. Thus one may conclude from a thermodynamic point of view that for all TOT values the 3PRH is the best

configuration since it allows achieving a high pressure, destroying less exergy, producing more steam quantities and

more specific work. However, this causes an increase of the steam cycle cost.

11
5.3 Economic optimization

As shown by Fig. 11, for an exhaust gas mass flow rate of 650 kg/s, the cost of a configuration 3PRH is

significantly higher than 2PRH or 1PRH. Also, this figure shows that the steam cycle cost for a given gas mass flow

rate increases with TOT. Furthermore, for a TOT equal to 650°C and above, the cost of steam cycle marks relatively a

rapid increase than for low TOTs. In order to decide between the three types of HRSG, the economical aspect should be

considered. The design parameters maximizing and NPV for several electricity selling prices will be determined.

Figure 12 plots the maximum values of NPV for different values of TOT considered in this study, and as revealed the

selling price influences the optimum configuration. For a low selling price of 0.04 $/kWhr (Fig. 12a) an HRSG with

2PRH represents the best configuration and 1PRH is more profitable than 3PRH for all TOTs. The more the selling

price increases the more the addition of another level of pressure becomes more interesting. Also, the 2PRH seems to be

the best configuration for the relatively medium selling prices (Fig. 12b), but 3PRH leads to higher NPV than 1PRH

whatever the value of TOT. Furthermore, even for a high selling price, for example 0.2 $/kWhr (Fig. 12c), the 2PRH

allows having NPV values similar to those of the configuration 3PRH. For a selling price of 0.3 $/kWhr and values of

TOT lower than 600°C, the configuration of 3PRH allows getting higher NPVs than for 2PRH.

The optimization of the three HRSG configurations to operate with a specific gas turbine such as GE-PG9371FB

(design parameters are listed in table 1) allows obtaining the results reported in table 4, which reveals that for prices

lower than or equal to 0.2 $/kWh, the use of the configuration of 2PRH is more beneficial while the 3PRH is only

interesting for a very high selling price of 0.3$/kWh. The configuration 1PRH is also better than 3PRH for prices equal

0.04 and 0.08 $/kWh while for a very high selling price, the configuration 1PRH gives the lowest NPV. For Sp=0.04 the

net benefit obtained with 2PRH is higher than the one of 3PRH by 42.26 million USD which represents 13.2 % of the

NPV obtained with 3PRH. Increasing the selling price till 0.08 or 0.2 $/kWh decreases the net benefit, respectively, at

4.97% and 2.11%, but 2PRH remains better than 3PRH. Furthermore, for a low price such as 0.04$/kWh, the NPV of

1PRH is higher than 3PRH by 31.47 million USD representing 9.82%.

The results of Fig. 12 can be correlated to express the NPV in relation to TOT, gas mass flow rate and selling price,

thus, the following correlations are suggested for 1PRH, 2PRH and 3PRH:

The mean absolute percentage error (MAPE) is 1.2 %, 1.04 % and 2.3 %, respectively for 1PRH, 2PRH and 3PRH. The

MAPE is defined as follows:

12
where is the actual value (results of Fig. 12), is the value obtained with the correlation and

is the number of times for which is calculated.

The optimum economic design parameters maximizing the NPVs are compared to those of the optimum

thermodynamic design parameters in Fig. 13, Fig. 14 and Fig. 15, plotting for several selling prices the optimum

pressure at the inlet of HP turbine, the pinch point and the effectiveness of the superheater. The evaluated data of these

curves are given in tables from 5 to 7.The analyses of these curves reveal the following aspects:

Inlet pressure at HP turbine: For the configuration 3PRH (Fig. 13a), the more the selling price increases, the more the

optimum economic design inlet pressure tends to the optimum thermodynamic design value, till reaching the constraint

of 160 bar. For the 2PRH (Fig. 13b), even though the decrease in selling price leads to an optimum economic inlet

pressure which is different from that of the optimum thermodynamic design value, but the difference is not as

significant as in the case of 3PRH. This is also confirmed for the 1PRH, as shown by Fig. 13c, for which the optimum

economic design inlet pressure coincides with the thermodynamic design value for all TOTs. It may be concluded that,

in addition to the fact that a decrease in selling price brings the economic and the thermodynamic optimum values close,

the reduction in pressure levels allows making the economic optimum independent of the selling price and close to the

thermodynamic optimum owing to a lesser cost of steam cycle and subsequently an increase in NPV.

Pinch points: Figure14 allows noticing that in case of 3PRH, the pinch point is clearly influenced by the selling price

for which a reduced value moves the optimum thermodynamic design value away from the optimum economic design,

but this is not the case for 2PRH and 1PRH which reveal a similarity between the optimum thermodynamic and the

economic pinch point. For a relatively high selling price (0.2$/kWhr or 0.3$/kWhr), the optimum pinch point of 3PRH

is identical to the thermodynamic optimum, except for TOT equal 650°C where the noticeable difference is attributed to

a rapid increase in the cost of steam cycle for this temperature, as previously illustrated by Fig. 11.

Superheater effectiveness: Figure 15 shows that for the thermodynamic optimization, the superheater effectiveness

tends to its maximum value of 0.85 for all the TOTs except for 650°C, for which a drop is noticed. However, for

TOT=650°C in the case of 3PRH and 2PRH, the reduction of selling price always tends to move the economic optimum

away from the thermodynamic one, but for the 1PRH, as expected, there is no noticeable influence.

In summary, the 3PRH allows destroying less exergy reaching higher optimum inlet pressure and producing more

steam quantity and specific work, thus representing the best configuration from a thermodynamic point of view.

However, with respect to the economic aspect, the results are different and the choice for the optimum configuration

depends on the electricity selling price, when the NPV is chosen as a criterion for the economic performance. For low

13
selling prices, the 2PRH represents the best configuration and 1PRH leads to higher NPV than 3PRH, whereas for a

medium selling price, the 2PRH also represents the best configuration and 3PRH leads to a higher NPV than the case of

1PRH. For high selling prices and medium TOTs, the configuration 3PRH has the highest NPV, but for high TOTs the

configuration 2PRH remains the best solution due to the high cost of 3PRH. Concerning the optimum design

parameters, the selling price seems to impact much more on the 3PRH configuration and the more the selling price

drops the more the parameters of the optimum economic design parameters move away from the parameters of the

optimum thermodynamic design. The reduced selling price leads to a decrease in the optimum pressure at the inlet of

HP turbine, an increase in pinch point and a decrease of superheater effectiveness. However, the influence for 2PRH is

less than 3PRH, and it is almost insignificant for 1PRH. This can be explained by the fact that by reducing the number

of pressure levels, this leads to a lesser cost of the steam cycle and therefore an increase in NPV.

5.4 Thermo-economic optimization

As shown in the previous section, in the case of a configuration 3PRH, (when NPV is considered as the objective

function to be maximized), the optimum pressure at HP turbine inlet increases with the selling price, especially for

TOT lower or equal than 550°C, till reaching the optimum pressure of the thermodynamic optimization. For TOT

higher than 550°C, the pressure reaches its limit which is constrained by the value of 160 bar, whatever the selling

price. However, the combination of the two objectives reduces the sensitivity of the optimum pressure toward the

selling price, as shown by Fig.16. For a price higher or equal than 0.08$/kWhr, the optimum pressures correspond

practically to those of the thermodynamic optimization, except for sp=0.04$/kWh where the optimum pressure is

different. Figure 17 reveals that the optimum pinch point of 3PRH shows a sensitivity toward the selling price only for a

low value of 0.04$/kWhr and practically for all TOTs, except 350°C. For the selling price of 0.08$/kWhr, the change in

the optimum values start from a TOT of 550°C, while for the economic optimization, changes are significant for all the

TOTs. This increase in the pinch point may be explained by the increase of the steam cycle capital cost with TOT (as

previously shown by Fig. 11). For the selling prices 0.2$/kWhr and 0.3$/kWhr, the optimum pinch points correspond to

the results obtained in the thermodynamic optimization independently from TOT. For the configuration 2PRH the

optimum pressure (which is slightly dependent on selling price) converges totally to the thermodynamic results, in the

case of thermo-economic optimization, as illustrated by Fig. 16. For 1PRH, since both of the thermodynamic and the

economic optimizations lead to the same parameters. The combination will also give the same results, thus by

considering the two objectives simultaneously is not necessary. Concerning the superheater effectiveness, Fig. 18 shows

that it has no significant effect on the optimum values and all of them take approximately the same value in both

thermodynamic and thermo-economic optimizations.

14
6 CONCLUSION

A thermo-economic parametric study and an optimization were undertaken for HRSG of configurations one, two

and three pressure levels with a reheat, and considering several values of turbine outlet temperature, electricity selling

prices and interest rate. The considered objective functions are the steam cycle specific work and the NPV for

thermodynamic and economic optimizations, respectively. Correlations relating the NPVs of the three configurations

HRSG in relation to TOT, gas mass flow rate and selling price are suggested. The main conclusions to be drawn are:

 From the thermodynamic point of view, the steam cycle with 3PRH and reheat is the best configuration for all

values of TOT because it allows producing more steam quantities, reaching higher pressures, destroying less exergy

and producing more specific work.

 Considering the economical aspect, for the low and medium selling prices (0.04-0.08$/kWh and 0.2$/kWh), 2PRH

is the best configuration.

 The configuration 3PRH would only be interesting for a high selling price of 0.3$/kWh and TOT range 450-600°C,

while for a high TOT (650°C) the high cost of 3PRH leads to NPV decrease.

 The electricity selling price impacts more the 3PRH and its influence on the design parameters of 2PRH is less than

for 3PRH, and becomes almost insignificant for 1PRH.

 For the configuration 3PRH, the less electricity selling price, the more the economic optimum moves away from the

thermodynamic one.

 The two objective functions steam cycle specific work and NPV show that, in the case of 3PRH, the selling price

influences only the design parameters for a low value of 0.04$/kWh, whereas for 2PRH and 1PRH, their optimum

parameters are non sensitive to the selling price variation.

 The interest rate influence has revealed via the parametric analysis an important effect on NPV and on optimum

design parameters.

As a perspective to improve the NPV correlations a wide range of interest rate, system life and additional cost due to

emissions will be included.

Funding

This research has received no funding

Conflict of interest

None declared.
15
REFERENCES

1. Griffin T, Büker D, Pteffer A. Technology Options for Gas Turbine Power Generation With Reduced CO2

Emission. ASME Journal of Engineering for Gas Turbines and Power 130 (2008)1–8

2. Valdés M, Rapuan J L. Optimization of heat recovery steam generator for combined cycle power plants. Applied

Thermal Engineering 21 (2001) 1149–1159

3. Franco A, Casarosa C. On some perspectives for increasing the efficiency of combined cycle power plants.

Applied Thermal Engineering 22 (2002) 1501–1518

4. Franco A, Russo A. Combined cycle plant efficiency increase based on the optimization of the heat recovery

steam generator operating parameters. International Journal of Thermal Sciences 41 (2002) 843–859

5. Xiang W, Chen Y. Performance improvement of combined cycle power plant based on the optimization of the

bottom cycle and heat recuperation. Journal of Thermal Science 16 (2007) 84―89

6. Mohagheghi M, Shayegan J. Thermodynamic optimization of design variables and heat exchangers layout in

HRSGs for CCGT, using genetic algorithm. Applied Thermal Engineering 29 (2009) 290–299

7. Bracco S, Siri S. Exergetic optimization of single level combined gas-steam power plants considering different

objective functions. Energy 35 (2010) 5365-5373

8. Kaviri A G, Jaafar M N M, Lazim T M, Barzegaravval H. Exergoenvironmental optimization of Heat Recovery

Steam Generators in combined cycle power plant through energy and exergy analysis. Energy Conversion and

Management 67 (2013) 27–33

9. Bassily A M. Modeling, numerical optimization, and irreversibility reduction of a dual-pressure reheat combined

cycle. Energy 81 (2005) 127–151

10. Bassily A M. Modeling, numerical optimization, and irreversibility reduction of a triple-pressure reheat combined

cycle. Energy 32 (2007) 778–794

11. Godoy E, Scenna N J, Benz S J. Families of optimal thermodynamic solutions for combined cycle gas turbine

(CCGT) power plants. Applied Thermal Engineering 30 (2010) 569–576

12. Polyzakis A L, Koroneos C, Xydi G. Optimum gas turbine cycle for combined cycle power plant. Energy

Conversion and Management 49 (2008) 551–563

13. Khaliq A, Kaushik S C. Second-law based thermodynamic analysis of Brayton/Rankine combined power cycle

with reheat. Applied Energy 78 (2004) 179–197

14. Chiesa P, Macchi E. A Thermodynamic analysis of different options to break 60% electric efficiency in combined

cycle power plants. ASME Journal of Engineering for Gas Turbines and Power126 (2004) 770-785

16
15. Sanjay Y, Onkar Singh, Prasad B N. Energy and exergy analysis of steam cooled reheat gas–steam combined

cycle. Applied Thermal Engineering 27 (2007) 2779–2790

16. Boyaghchi F A, Molaie H. Advanced exergy and environmental analyses and multi objective optimization of a

real combined cycle power plant with supplementary firing using evolutionary algorithm. Energy 93 (2015) 2267-

2279

17. Valdés M, Duran M D, Rovira A. Thermoeconomic optimization of combined cycle gas turbine power plants

using genetic algorithms. Applied Thermal Engineering 23 (2003) 2169–2182

18. Koch C, Cziesla F, Tsatsaronis G. Optimization of combined cycle power plants using evolutionary algorithms.

Chemical Engineering and Processing 46 (2007) 1151–1159

19. Casarosa C, Donatini F, Franco A. Thermoeconomic optimization of heat recovery steam generators operating

parameters for combined plants. Energy 29 (2004) 389–414.

20. Bassily A M. Analysis and cost optimization of the triple-pressure steam-reheat gas-reheat gas-recuperated

combined power cycle. International Journal of Energy Research 32 (2008) 116–134

21. Kotowicz J, Bartela Q. The influence of economic parameters on the optimal values of the design variables of a

combined cycle plant. Energy 35 (2010) 911–919

22. Ahmadi P, Dincer I. Thermodynamic analysis and thermoeconomic optimization of a dual pressure combined

cycle power plant with a supplementary firing unit. Energy Conversion and Management 52 (2011) 2296–2308

23. Carapellucci R, Giordano L. A comparison between exergetic and economic criteria for optimizing the heat

recovery steam generators of gas-steam power plants. Energy 58 (2013) 458–472

24. Rovira A, Sánchez C, Muñoz M, Valdés M, Durán MD. Thermoeconomic optimization of heat recovery steam

generators of combined cycle gas turbine power plants considering off-design operation. Energy Conversion and

Management 52 (2011) 1840–1849

25. Bakhshmand S K, Saray R K, Bahlouli K, Eftekhari H, Ebrahimi A. Exergo-economic analysis and optimization

of a triple-pressure combined cycle plant using evolutionary algorithm Energy 93 (2015) 555-567

26. Rovira A, Sánchez C, Muñoz M. Analysis and optimisation of combined cycles gas turbines working with partial

recuperation. Energy Conversion and Management 106 (2015) 1097–1108

27. Vandani A M K, Joda F, Boozarjomehry R B B. Exergic, economic and environmental impacts of natural gas and

diesel in operation of combined cycle power plants. Energy Conversion and Management 109 (2016) 103–112

28. Taamallah S, Vogiatzaki K, Alzahrani F M, Mokheimer E.M.A, Habib M A, Ghoniem A F. Fuel flexibility,

stability and emissions in premixed hydrogen-rich gas turbine combustion: Technology, fundamentals, and

numerical simulations. Applied Energy 154 (2015) 1020–1047

17
29. http://www.energy.siemens.com/hq/en/fossil-power-generation/gas-turbines/sgt5-

8000h.htm#content=Technical%20data

30. J. Ramachandran and M.C. Conway. MS6001FA – An advanced-technology 70-MW class 50/60... Hz gas turbine.

Report GER 3765B GE Power Systems. Schenectady. NY. 1996

31. Sánchez D, Chacartegui R, Muñoz J M, Muñoz A, Sanchez T. Performance analysis of a heavy duty combined

cycle power plant burning various syngas fuels. International Journal of Hydrogen Energy 2010;35:337–345

32. Carlo Carcasci, Riccardo Ferraro, Edoardo Miliotti Thermodynamic analysis of an organic Rankine cycle for

waste heat recovery from gas turbines Energy 65 (2014) 91-100

33. Wagner W, Kretzschmar H J. International Steam Tables Properties of Water and Steam Based on the Industrial

Formulation IAPWS-IF97 Springer 2008

34. Ganjeh kaviri A, Mohd Jaafar M N, Ahmadi P, Barzegaravval H. Modelling and optimization of combined cycle

power plant based on exergoeconomic and environmental analyses. Applied Thermal Engineering 67 (2014) 566-

578

35. Roosen P, Uhlenbruck S, Lucas K. Pareto optimization of a combined cycle power system as a decision support

tool for trading off investment vs. operating costs. International Journal of Thermal Science 42 (2003) 553-60.

36. Sonntag R E, Borgenakke C, Van Wylen G J. Fundamentals of thermodynamics, 5th ed., vol. 356–57. New York.

John Wiley & Sons, Inc.1998.

37. Ganjeh kaviri A, Mohd Jaafar MN, Hosseini S E. Optimization and the effect of steam turbine outlet quality on the

output power of a combined cycle power plant. Energy Conversion and Management. 2015. 89: 231-243

38. Chase D L, Kehoe P T. GE combined-cycle product line and performance. Report GER-3574g. GE Power

Systems. Schenectady. NY. 2000.

39. Dick E, De Paepe M. Steam-injected gas-turbine: cycle analysis and feasibility of water recuperation. Eur J Mech

Eng 1997.42(2)67–77

40. Eberhart R, Kennedy J. A new optimizer using particle swarm theory. MHS 95. Proceedings of the Sixth

International Symposium on Micro Machine and Human Science, pages 39-43, October 1995

41. Bratton D, Kennedy J. Defining a Standard for Particle Swarm Optimization. Proceedings of the 2007 IEEE

Swarm Intelligence Symposium (SIS 2007)

42. Nadir M, Ghenaiet A. Thermodynamic optimization of several HRSG (Heat Recovery Steam Generator)

configurations for a range of exhaust gas temperatures. Energy 86 (2015) 685-695

NOMENCLATURE

18
Heat transfert area [ ]

Cost [US$]

Specific heat capacity [kJ/kg K]

Effectiveness of exchanger

Ex Exergy [kJ]

Specific exergy [kJ/kg]

Fitness function

Global best solution

Steam or water specific enthalpy [kJ/kg]

Yearly operating hours

Annual interest rate

Total annual income

Mass flow rate [kg/s]

Condenser cooling water mass flow rate [kg/s]

Number of particles in population

Average service life

Pressure [bar]

Power [kW]

Best solution for each particle

Population

Selling price

Temperature [K]

speed of particle vector

Inertia weight

Specific work [kJ/kg]

Vapor fraction at turbine exit

Optimization variables vector

Logarithmic mean temperature difference [K]

Pinch point [K]

Efficiency

19
Isentropic efficiency

Steam to gas mass ratio

SUBSCRIPTS

Condenser

Exit

Economizer

Evaporator

Gas

High pressure

Inlet

Intermediate pressure

Particle

Low pressure

Reheater

Steam

Steam cycle

Superheater

Steam turbine

Thermal

First level

Second level

Third level

Gas station at different heat exchangers

ABBREVIATION

GT Gas turbine

CC Combined cycle

HRSG Heat recovery steam generator

IAPWS International Association for the Properties of Water and Steam

NPV Net present value

TOT Turbine outlet temperature

20
1PRH Reheat one pressure level HSRG

2PRH Reheat two pressures level HSRG

3PRH Reheat three pressures level HSRG

21
(a) (b)

(c)

Figure 1: Three configurations of HRSG: a) 1PRH, b) 2PRH, c) 3PRH

(a) (b)

22
(c)

Figure 2: Temperature-transferred heat diagram: a) HRSG 1PRH, b) HRSG 2PRH, c) HRSG 3PRH

222 8
6.3x10

220
8
6.3x10
218
Wsc (kJ/kg)

8
6.2x10
NPV ($)

216
8
6.2x10
214

8
212 6.2x10

8
210 6.2x10
10 20 30 40 50 60 70 80 90 100 10 20 30 40 50 60 70 80 90 100
Iterations iterations

(a) (b)

Figure 3: Objective function convergence for 3PRH at TOT=600°C: a) maximization, b) NPV maximization

280

278
NPV (Million US$)

276

274

272

270

207 208 209 210 211 212 213 214 215


Ws (kJ/kg)

Figure 4: Pareto front in the case of 3PRH, TOT=600°C and sp=0.04

23
260.0
112
586.5
262.5
110
259.5
585.0

NPV (million US$)


108 260.0
NPV (million US$)

NPV (million US$)


NPV (million US$)
259.0
106 583.5 257.5

258.5 104
582.0 255.0
102
258.0 580.5
100 252.5

i=0.04 98 i=0.04
257.5 579.0 250.0
i=0.2 i=0.2
96
3 6 9 12 15 18 21 24 27 30 3 6 9 12 15 18 21 24 27 30
Pinch point (K) Pinch point (K)

(a) (b)
726
1575 i=0.04
i=0.2
1570
724
NPV (million US$)

NPV (million US$)


1565

1560 722

1555
720
1550

1545 718

3 6 9 12 15 18 21 24 27 30

Pinch point (K)

(c)

Figure 5: NPV variation for 3PRH versus pinch point, selling price (sp) and interest rate (i) as parameters:

a) sp=0.04$/kWh, b) sp=0.08$/kWh, c) sp=0.2$/kWh

718 320
i=0.04
i=0.04
i=0.2
352 i=0.2 148.0 716
NPV (million US$)

319
NPV (million US$)

NPV (million US$)


NPV (million US$)

351 714
147.5
350 712
318

147.0 710
349

317
348 708
146.5

0.4 0.5 0.6 0.7 0.8 0.9 0.4 0.5 0.6 0.7 0.8 0.9

Superheater effectiveness Superheater effectiveness

(a) (b)

24
1810 i=0.04
i=0.2 836

1805

NPV (million US$)


834

NPV (million US$)


1800 832

1795 830

1790 828

826
1785
0.4 0.5 0.6 0.7 0.8 0.9

Superheater effectiveness

(c)

Figure 6: NPV variation for 3PRH versus superheater effectiveness: a) sp=0.04$/kWh, b) sp=0.08$/kWh, c)

sp=0.2$/kWh

353
319
715
147.5
351
318
NPV (million US$)

NPV (million US$)


NPV (million US$)

NPV (million US$)

350 710
147.0
317
348
146.5 705
315
347

146.0
345 700 i=0.04 314
i=0.04
i=0.2 i=0.2
344 145.5
40 60 80 100 120 140 160 40 60 80 100 120 140 160

Pressure at HP turbine inlet (bar) Pressure at HP turbine inlet (bar)

(a) (b)

1810 835

1800
NPV (million US$)

NPV (million US$)

830
1790

825
1780

1770 820

i=0.04
1760 i=0.2 815

40 60 80 100 120 140 160

Pressure at HP turbine inlet (bar)

(c)

Figure 7: NPV variation for 3PRH versus the pressure at HP turbine inlet: a) sp=0.04$/kWh, b) sp=0.08$/kWh, c)

sp=0.2$/kWh

25
220 1PRH
2PRH
200
3PRH
180

160

Wsc (kJ/kg)
140

120

100

80

60

40
350 400 450 500 550 600 650
TOT(°C)

Figure 8: Steam cycle specific work versus TOT

18 500
160 1PRH
Pressure at HP turbine inlet (bar)

140
2PRH
3PRH 16 450
120
14 400
100

80 12 350
60
pressure at HP
10
turbine inlet (bar) 300
40
steam ratio (%)
20 steam temperature at 8
inlet of reheater (°C) 250
0
350 400 450 500 550 600 650
TOT(°C)

Figure 9: Optimum pressure at inlet of HP turbine, optimum total steam ratio and steam temperature at exit of reheater

versus TOT

50
1P
45 2P
3P
40
exdr (%)

35

30

25

20
350 400 450 500 550 600 650
TOT(C°)

Figure 10: Destroyed exergy of steam cycle

26
130
3PRH
120 2PRH

Steam cycle cost (Million US $)


110 1PRH
100
90
80
70
60
50
40
30
20
350 400 450 500 550 600 650
TOT

Figure 11: Steam cycle cost corresponding to the optimum thermodynamic parameters of steam cycle

8 8
4.0x10 8x10
1PRH 1PRH
3.5x10
8
2PRH 7x10
8
2PRH
3PRH 3PRH
8 8
3.0x10 6x10

8 8
2.5x10 5x10
NPV ($)

NPV ($)

8 8
2.0x10 4x10

8 8
1.5x10 3x10

8 8
1.0x10 2x10

7 8
5.0x10 1x10
350 400 450 500 550 600 650 350 400 450 500 550 600 650
TOT (°C) TOT (°C)

(a) (b)
9 9
2.0x10 3.0x10
9
1PRH 1PRH
1.8x10 2PRH 2PRH
9
9 3PRH 2.5x10 3PRH
1.6x10
9
1.4x10 9
2.0x10
NPV ($)
NPV ($)

9
1.2x10
9 9
1.0x10 1.5x10
8
8.0x10
9
6.0x10
8 1.0x10

8
4.0x10
8
5.0x10
350 400 450 500 550 600 650 350 400 450 500 550 600 650
TOT (°C) TOT (°C)

(c) (d)

Figure 12: Optimum NPVs versus TOT for several electricity selling price: a) sp=0.04$/kWhr, b) sp=0.08$/kWhr,

c) sp=0.2$/kWhr, d) sp=0.3$/kWhr

27
160 160

Pressure at HP turbine inlet (bar)

Pressure at HP turbine inlet (bar)


140 140

120 120

100 100

80 80
sp=0.04 $/kWh sp=0.04 $/kWh
60 sp=0.08 $/kWh 60 sp=0.08 $/kWh
sp=0.2 $/kWh sp=0.2 $/kWh
40 sp=0.3 $/kWh 40 sp=0.3 $/kWh
Thermodynamic Thermodynamic
20 20 optimization
optimization

350 400 450 500 550 600 650 350 400 450 500 550 600 650
TOT (°C) TOT (°C)

(a) (b)

160
sp=0.04 $/kWh
Pressure at HP turbine inlet (bar)

140 sp=0.08 $/kWh


sp=0.2 $/kWh
120 sp=0.3 $/kWh
Thermodynamic
100 optimization
80

60

40

20

350 400 450 500 550 600 650


TOT (°C)

(c)

Figure 13: Effect of selling price on optimum inlet pressure at HP turbine: a) 3PRH, b) 2PRH, c) 1PRH

20 20
sp=0.04 $/kWh
18 18
sp=0.08 $/kWh
16 16 sp=0.2 $/kWh
sp=0.3 $/kWh
Pinch point (°C)
Pinch point (°C)

14 14 Thermodynamic
sp=0.04 $/kWh, sp=0.08 $/kWh optimization
12 sp=0.2 $/kWh, sp=0.3 $/kWh 12
Thermodynamic optimization
10 10

8 8

6 6

4 4
350 400 450 500 550 600 650 350 400 450 500 550 600 650
TOT (°C) TOT (°C)

(a) (b)

Figure 14: Effect of electricity selling price on optimum pinch point: a) 3PRH, b) 2PRH and 1PRH

28
0.90 0.90

0.85 0.85
Effectiveness of superheater

Superheater effectiveness
0.80 0.80

0.75 0.75
sp=0.04 $/kWh sp=0.04 $/kWh
0.70 sp=0.08 $/kWh 0.70
sp=0.08 $/kWh
sp=0.2 $/kWh sp=0.2 $/kWh
sp=0.3 $/kWh sp=0.3 $/kWh
0.65 0.65 Thermodynamic
Thermodynamic
optimization optimization
0.60 0.60
350 400 450 500 550 600 650 350 400 450 500 550 600 650
TOT (°C) TOT (°C)

(a) (b)
0.90

0.85
Superheater effectiveness

0.80

0.75
sp=0.04 $/kWh
0.70 sp=0.08 $/kWh
sp=0.2 $/kWh
sp=0.3 $/kWh
0.65
Thermodynamic
optimization
0.60
350 400 450 500 550 600 650
TOT (°C)

(c)

Figure 15: Effect of selling price on optimum effectiveness: a) 3PRH, b) 2PRH, c) 1PRH

160
Pressure at HP turbine inlet (bar)

140
3PRH
120

100

80 sp=0.04$/kWh
sp=0.08$/kWh
60
sp=0.2$/kWh
40 sp=0.3$/kWh
Thermodynamic
20
2PRH optimization

350 400 450 500 550 600 650


TOT (°C)

Figure 16: Selling price effect on optimum pressure at HP turbine inlet of 2PRH and 3PRH in case of thermo-

economic optimization. Dashed lines: 3PRH, continuous lines: 2PRH

29
20
sp=0.04 $/kWh, sp=0.08 $/kWh,
sp=0.2 $/kWh, sp=0.3 $/kWh,
15 Thermodynamic
optimization

Pinch point (°C)


10

0
350 400 450 500 550 600 650
TOT (°C)

Figure 17: Selling price effect on optimum pinch point of 3PRH in case of thermo-economic optimization

0.88
0.84
Effectiveness of superheater

0.80

0.76
0.72
0.68 sp=0.04 $/kWh
0.64 sp=0.08 $/kWh
sp=0.2 $/kWh
0.60
sp=0.3 $/kWh
0.56 Thermodynamic
0.52 optimization

350 400 450 500 550 600 650


TOT (°C)

Figure 18: Selling price effect on optimum 3PRH superheater effectiveness in case of thermo-economic optimization

30
Table 1: Model validation

Parameter (unit) Real data [31] Model results


Power ( MW) 145.5 146.21
1st level mass flow rate (kg/s) 12.63 13.01
2nd level mass flow rate (kg/s) 12.76 13.51
3rd level mass flow rate (kg/s) 86.7 88.07
Maximum steam cycle temperature (°C) 565 564.9
Stack temperature (°C) 96 94.8

Table 2 : Economic assumptions

Economic parameters Assumptions


Average service life ( ) 12 years
Yearly operating hours ( ) 8000 hours [16]
The annual interest rate ( ) 8% [19]
Electricity selling price ( ) 0.04 - 0.3 $/kWh
Operation and maintenance cost ( ) 10% of [16]

Table 3: Steam cycle components costs [23, 35]

Number of
Cost correlations
elements

Steam
3
turbines
corresponds to LP turbine, IP turbine and HP turbine

( 990)500 0.8+17500.2 =1100.097 30+0.9 +1948.4


1.2
HRSG heat 10
exchangers
corresponds to the heat exchanger in the HRSG namely: 1st, 2nd and 3rd level
economizers, evaporators and superheaters.

Condenser 1

3
Pumps
31
corresponds to the pump of each level (pump of 1st , 2nd and 3rd level).

Table 4: Optimized NPV in case of the gas turbine GE-PG9371FB


Sp ($/kWh) 1PRH (million US$) 2PRH (million US$) 3PRH (million US$)
0.04 346.7046 362.565 320.306
0.08 701.339 727.842 693.406
0.2 1773.725 1863.937 1825.420
0.3 2667.792 2790.898 2809.531

Table 5: 3PRH optimum design variables

TOT Thermodynamic Economic optimization


(°C) optimization Sp=0.04 Sp=0.08 Sp=0.2 Sp=0.03

(bar) (K) (bar) (K) (bar) (K) (bar) (K) (bar) (K)
350 41.93 5 0.85 31.35 14.19 0.85 33.43 6.28 0.85 33.44 5 0.85 33.62 5 0.85
400 63.10 5 0.85 50.05 14.28 0.85 54.88 6.36 0.85 55.30 5 0.85 55.91 5 0.85
450 139.56 5 0.85 78.66 14.35 0.85 82.20 7.21 0.85 92.80 5 0.85 103.37 5 0.85
500 160 5 0.85 102.72 14.39 0.85 138.10 7.67 0.85 160 5 0.85 160 5 0.85
550 160 5 0.85 160 15.83 0.85 160 8.46 0.85 160 5 0.85 160 5 0.85
600 160 5 0.85 160 17.62 0.85 160 9.42 0.85 160 5 0.85 160 5 0.85
650 160 5 0.76 160 19.04 0.64 160 11.11 0.69 160 6.02 0.75 160 5.73 0.75

Table 6: 2PRH optimum design variables

TOT Thermodynamic Economic optimization


(°C) optimization Sp=0.04 Sp=0.08 Sp=0.2 Sp=0.03

(bar) (K) (bar) (K) (bar) (K) (bar) (K) (bar) (K)
350 21.34 5 0.85 21.12 5 0.85 21.20 5 0.85 21.11 5 0.85 21.90 5 0.85
400 36.34 5 0.85 32.37 5 0.85 32.85 5 0.85 32.26 5 0.85 32.94 5 0.85
450 55.69 5 0.85 53.85 5 0.85 54.59 5 0.85 54.67 5 0.85 54.013 5 0.85
500 95.00 5 0.85 78.02 5 0.85 87.52 5 0.85 87.49 5 0.85 87.04 5 0.85
550 160 5 0.85 160 5 0.85 160 5 0.85 160 5 0.85 160 5 0.85
600 160 5 0.85 160 5 0.85 160 5 0.85 160 5 0.85 160 5 0.85
650 160 5 0.76 160 5 0.73 160 5 0.75 160 5 0.76 160 5 0.76

Table 7: 1PRH optimum design variables

TOT Thermodynamic Economic optimization


(°C) optimization Sp=0.04 Sp=0.08 Sp=0.2 Sp=0.03

(bar) (K) (bar) (K) (bar) (K) (bar) (K) (bar) (K)
350 11.021 5 0.85 11.022 5 0.85 11.021 5 0.85 11.022 5 0.85 11.020 5 0.85
400 16.754 5 0.85 16.754 5 0.85 16.754 5 0.85 16.754 5 0.85 16.749 5 0.85
450 24.356 5 0.85 24.354 5 0.85 24.355 5 0.85 24.356 5 0.85 24.343 5 0.85
500 38.006 5 0.85 38.006 5 0.85 38.006 5 0.85 38.008 5 0.85 38.004 5 0.85
550 59.014 5 0.85 59.014 5 0.85 59.014 5 0.85 59.034 5 0.85 59.018 5 0.85
600 160 5 0.85 160 5 0.85 160 5 0.85 160 5 0.85 160 5 0.85
650 160 5 0.75 160 5 0.75 160 5 0.75 160 5 0.75 160 5 0.75

32
33
Highlights

 Thermo-economic optimization of HRSG configurations

 The maximum value of the net present value was targeted for the economic optimization

 Three level HRSG is the best option in respect of power output and high priced medium

 Two level HRSG is the best for net benefit in low and intermediate priced mediums

34

Anda mungkin juga menyukai