You are on page 1of 63

Accepted Manuscript

Research Paper

Design of structure and optimization of Organic Rankine Cycle for heat recovery
from gas turbine: the use of 4E, advanced exergy and advanced exergoeconomic
analysis

Hossein Khosravi, Gholam Reza Salehi, Masoud Torabi Azad

PII: S1359-4311(18)30916-5
DOI: https://doi.org/10.1016/j.applthermaleng.2018.09.128
Reference: ATE 12737

To appear in: Applied Thermal Engineering

Received Date: 8 February 2018


Revised Date: 24 September 2018
Accepted Date: 28 September 2018

Please cite this article as: H. Khosravi, G. Reza Salehi, M. Torabi Azad, Design of structure and optimization of
Organic Rankine Cycle for heat recovery from gas turbine: the use of 4E, advanced exergy and advanced
exergoeconomic analysis, Applied Thermal Engineering (2018), doi: https://doi.org/10.1016/j.applthermaleng.
2018.09.128

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.
Design of structure and optimization of Organic Rankine
Cycle for heat recovery from gas turbine: the use of 4E,
advanced exergy and advanced exergoeconomic analysis
Hossein Khosravi1, Gholam Reza Salehi2*, Masoud Torabi Azad1
1Department of Energy System engineering, North Tehran Branch, Islamic Azad University, Tehran Iran

2Department of Mechanical Engineering, Petroleum University of technology, Abadan, Iran.

Rezasalehi20@gmail.com
Design of structure and optimization of Organic
Rankine Cycle for heat recovery from gas Turbine:
the use of 4E, advanced exergy and advanced
exergoeconomic analysis
Abstract
This paper presents the evaluation and optimization of an organic Rankine cycle (ORC)
approached from four different perspectives: 1) selecting the ORC Cycle; 2) selecting the
working fluid in accordance with the thermodynamic properties and environmental impacts; 3)
analyzing energy and exergoeconomic; and finally, 4) advanced exergy and advanced
exergoeconomic. The working fluid is selected based on the thermal source temperature and the
environmental impacts including the reduction of ozone depletion (ODP) and global warming
potential (GWP). Then, the selected working fluids in different cycles are investigated in terms
of the thermodynamic properties based on energy and exergy concepts. The parameters of the
selected cycles (from the perspective of energy and exergy), including the temperature and
pressure of the input working fluid of the turbine, the pinch and approach temperature, and so on,
are optimized by a genetic algorithm. Two objective functions of price and exergy efficiency are
selected as the objective functions for optimal cycles. The results of this study reveal that single-
pressure and dual-pressure cycles with recuperator and the R123 working fluid have the highest
power and the lowest cost. It is indicated that the net power generation of dual pressure cycle
with recuperator, single pressure with recuperator, and dual pressure cycle with two working
fluids are 2.2 kW, (R123: 2.31 kW, R600: 1.72 kW) and 2.26 kW, respectively. Also, the cost of
generated electricity for dual pressure cycle with recuperator, single pressure cycle with
recuperator, and dual pressure cycle with two working fluids are 14.59, (R123: 13.03, R600:
21.95) and 16.78 (Cent/kWh), respectively. In addition, the results show that the cycles of dual
pressure with recuperator and single pressure with recuperator with R123 as a working fluid have
the highest exergy efficiency. The advanced exergy analysis indicated that the HRSG and turbine
components are important to be improved based on exergetic performance.
Keywords: exergy, exergoeconomic, genetic algorithm, environmental impacts
Nomenclature Subscript superscript
A Heat transfer area(m2) p Pump AV Available
CC Capital cost($) D Destruction UN Unavailable
CRF Capital recovery factor o Dead state EN Endogenous
C Stream cost rate($/s) E Evaporator EX Exogenous
EX Exergy stream(kW) C Condenser
h Specific enthalpy(kJ/kg) IHE Heat exchanger
Annual running
H T Turbine
time(hour)
Logarithm mean
i Interest rate (%) LMTD temperature
difference
m Mass flow rate(kg/s) ℎ𝑖 Hot stream inlet
Nominal life of ℎ𝑜
n Hot stream outlet
cycle(year)
OC Operation cost($) 𝑐𝑜 Cold stream outlet
P Pressure(kPa) 𝑐𝑖 Cold stream inlet
Purchased equipment cost
PEC Recup Recuperator
($)
Q Heat rate(kW) f Working fluid
s Specific entropy(kJ/kg) g gas
T Temperature(K) e output
TAC Total annual cost($/year) i input
W Work(kW) q Heat
x Mole fraction (%) p Product
Normalized equipment
Z F Fuel
purchase($/hour)
η Efficiency (%) L Lost
Φ Maintenance factor (%) Net net
1 Introduction
Owing to their unique attributes, the application of organic fluids in waste heat
has been tremendously increased as is evident in Figure 1, depicting the share of
the organic cycle in different low-grade thermal heat sources. It shows that the
organic Rankine cycle (ORC) has a share of 20% in various waste heat sources [1].
1%

20% Biomass ORC


48% Geothermal ORC
Weste heat recovery
31% Solar applications

Figure 1. The share of ORC in different low-grade thermal heat sources [1]
Four factors mainly contribute to the extensive application of fluids: (i)
diversity and various temperature ranges of low-grade resources; (ii) diversity of
organic fluids with diverse thermodynamic properties; (iii) productivity
enhancement; and (iv) economic issues and initial investment. Consequently, since
decision making should be based on a comprehensive assessment, extensive
studies have been conducted on organic cycles to explore the impact of different
parameters. Each study has focused on one of these aspects. Also, some literature
has looked at different cycles from multiple perspectives [2-4].
Haghighat Mamaghani et al. [4] worked on 4E (energy, exergy, economic,
environmental) analysis of an integrated cycle of a fuel cell, a gas turbine, and an
ORC. The cycle was optimized in terms of economic and exergy efficiency. They
reported that ORC application increased exergy efficiency by 5%. Mokhtari et al.
[3] conducted a 4E analysis of organic cycle as well as desalinated water
production by reverse osmosis system. In this cycle, organic working fluids were
determined on the basis of environmental impacts. The R123 fluid was used in the
presented optimized cycle.
Toffolo et al. [5] adopted a multi-criteria approach for optimum selection of
ORC. They aimed to pick a configuration with respect to economic criteria and
thermal efficiency. They utilized isobutene working fluid and R134a. The results
showed that the use of the preheating cycle can technically and economically
enhance the power and efficiency. Yu et al. [6] applied a new pinch-based method
and examined the working fluid and operating conditions in an ORC for the
recovery of waste heat. Two factors were found to determine the effectiveness of
working fluid parameters: (i) critical point of working fluid, and (ii) saturated
temperature difference with heat source temperature. Among introduced fluidity,
the R600 and R600a fluids exhibited a considerable power with optimum
environmental impacts. Lecompte et al. [7] carried out a comprehensive study on
ORC architecture. They introduced different architectures for organic cycles. As
well, they briefly reviewed empirical data, temperature range, generation power,
and heat and cold source temperature. It can be inferred from this review which
architectures could be used for heat recovery and which fluid could generate more
power in certain temperature ranges. Zhang et al. [8] used a novel method to assess
working fluids in terms of environmental impacts. They introduced Hasse weight
determination method for the assessment of organic fluids in terms of
environmental impacts, such as ODP, GWP, and ALT. This method has been
regarded as a good reference for fluid selection. The method categorizes the
working liquids by the Montreal Protocol. Chen et al. [9] explored 35 substances
for Rankine cycle and supercritical Rankine from thermodynamic, environment,
safety, and environment-friendliness perspectives. Bao and Zhao [10] worked on
the thermal range and the application type of working fluids. This comprehensive
study reviewed different fluids in various thermal ranges and specified the thermal
ranges for each specific working fluid. Also, extensive research has been done on
the use of ORC in gas turbine exhaust, e.g. Cao et al. [11]. They selected an ORC
with recuperator for gas turbine outlet and performed a sensitivity analysis to
specify its optimum parameters. Carcasci et al. [12] used an ORC for the recovery
of exhaust heat of the gas turbine. Given the foregoing thermal range, they
proposed to use toluene, cyclohexane, and cyclopentane as working fluids for
safety reasons. Clemente et al. [13] reported that the use of an ORC on gas turbine
enhanced the energy efficiency of the system. The optimum cycle was selected
from six working fluids.
As is evident, ORCs and their optimization have been subject to extensive
research in terms of energy, exergy, economy, and environmental impacts. But,
there is no detailed, integrated perspective on a roadmap for decision-making on
working fluid, efficiency constraints, energy quality, and optimum performance of
the individual components of the cycle. The present study first selects two cycles
with respect to the theories and the works already done and then, presents a cycle
as a new design for comparison. Working fluid is selected on the basis of cycle
thermodynamic conditions and environmental impacts. We explore different cycle
arrangements and their exposure to various pressure and thermal conditions in
terms of energy and exergy. Then, power generation and exergy efficiency are
considered to select architecture with fluid type and to optimize cycle parameters
in terms of a genetic algorithm. Optimization essentially seeks two objectives:
maximum power and minimum costs including instrument purchase, maintenance,
and exergy destruction. Finally, each cycle component is analyzed at optimum
state in terms of the advanced exergy and advanced exergoeconomics.
Accordingly, the followings can be listed as the innovations of the present work:
a. The selection of optimum architecture for the recovery of the gas turbine
exhaust.
b. 4E analysis of ORC and its optimization by an evolutionary genetic
algorithm (GA).
c. The examination of individual components of the cycle at optimum state
from advanced exergy and advanced exergoeconomic perspectives.
d. The selection of working fluid in terms of the environment,
thermodynamics, and performance range.
e. The introduction of a dual-pressure cycle in terms of thermal performance
range of organic fluids, 4E analysis, and comparison with other cycles.
2 Working Fluid Selection
The working fluids could be categorized according to the saturation vapor
curve, which is one of the most crucial characteristics of the working fluids in an
ORC. There are generally three types of vapor saturation curves in the
temperature-entropy (T–s) diagram: a dry fluid with positive slopes, a wet fluid
with negative slopes, and an isentropic fluid with nearly infinitely large slopes.
Due to the negative slope of the saturation vapor curve for a wet fluid, outlet
stream of the turbine typically contains a lot of saturated liquids. The presence of
liquid inside the turbine may damage its blades and it also reduces the isentropic
efficiency of the turbine. As well, the optimum efficiency of ORC working with a
dry fluid could be achieved when the fluid operates along the saturation curve
without being superheated. Given the extensiveness of organic fluids in different
thermal ranges, the factors considered in the present study are presented in Table 1.
As is evident, numerous parameters account for the selection of an organic fluid
and sometimes, there is a trade-off between a range and another, complicating the
fluid selection. The most important thermodynamic properties of these fluids are
summarized in Table 2 [14, 15, 16]. The working fluids chosen in this paper based
on Tables 1 and 2 and the abovementioned reasons are R123 and R600. These
fluids have more appropriate in this thermal range than other organic fluids.
We have safety degree = B1, ALT = 1.3 years, ODO = 0.02, and GWP = 77 for
the R123 fluid and safety degree = A3, ALT = 1 year, ODO = 0.00, and GWP = 8
for the R600 fluid. As is evident, R600 is better in terms of environmental impacts.
3 Modeling and Optimization
Table 1. Parameters considered for the selection of organic fluid
Range Selection factors of organic fluid (indices of the impact of these factors)
Environmental ODP, GWP, flammability, chemical stability, fluid corrosion, time to return
issues to nature
The proper performance of the organic fluid in a wide thermal range that
Performance
optimize the cycle performance.
Cycle efficiency or output work
Thermodynamic
Positive condenser pressure vs. atmospheric pressure (no need for
performance
accessories)
Critical temperature (cycle efficiency)
Thermodynamic Density, surface tension, latent heat of vaporization, (parameters affecting
features condenser thermal surface)
Specific heat capacity (affecting turbine dimensions)
Positive or negative slope and/or isentropicity of T-S curve
Inclusion or exclusion of superheater (cost increase and power increase)
Limitation in turbine production (output working fluid quality – condenser
T-S curve
dimensions)
The use of exchanger for output heat recovery from turbine (efficiency
improvement)
Economic Easy access and low price
Table 2. Thermodynamic properties and environmental parameters and safety classification of various
fluids
Global
Critical Critical Ozone
ASHARE Type of warming ASHRAE 34 Safety
temperature pressure depletion
CODE fluid potential -100 Group
(K) (kPa) potential
yr
R11 471.11 4407.6 Isentropic 1 4750 A1
R12 385.12 4136.1 Isentropic 1 10900 A1
R13 302 3879 Wet 1.7 14400 A1
R14 227.51 3750 Wet 0 7390 A1
R21 451.48 5181.2 Wet 0.04 151 B1
R22 369.295 4990 Wet 0.05 1810 A1
R23 299.293 4832 Wet 0 14800 A1
R32 351.255 5782 Wet 0 675 A2L
R113 487.21 3392.2 Dry 1 6130 A1
R114 418.83 3257 Dry 1 10000 A1
R115 353.1 3120 Isentropic 0.44 7370 A1
R116 293.03 3048 Wet 0 12200 A1
R123 456.831 3661.8 Dry 0.02 77 B1
R124 395.425 3624.2 Wet 0.022 609 A1
R125 339.173 3617.7 Isentropic 0 3500 A1
R134a 345.857 3761 Isentropic 0.055 1430 A1
R141b 477.5 4212 Dry 0.12 725 A2
R142b 410.26 4055 Isentropic 0.12 2310 A2
R143a 345.857 3761 Isentropic 0 4470 A2L
R152a 386.411 4516.7 Wet 0 124 A2
R236ea 412.44 3501.9 Dry 0 1370
R236fa 398.07 3200 Dry 0 9810 A1
R245fa 427.2 3640 Dry 0 1030 B1
R417a 360.19 4036 Isentropic 0 2346 A1
R422a 344.9 3747 Isentropic 0 3143 A1
R422d 352.71 3903 Isentropic 0 2729 A1
R423a 372.64 3587 Isentropic 0 2280 A1
R170 305.33 4871.8 Wet 0 5.5 A3
R290 369.825 4247.6 Isentropic 0 3.3 A3
R600 425.125 3796 Dry 0 4 A3
R601 469.7 3370 Dry 0 4 A3

The modeling in the paper is composed of three sections. The main section is the
energy and exergy equations composing the basis for the calculations are of the
other sections. The second section expresses the equations and concept of
exergoeconomic. Finally, the third section represents the equations of advanced
exergy and advanced exergoeconomic.

3.1 Energy and exergy analysis


The present system modeling is presented in Table 3 on the basis of the first
thermodynamic law and exergy equations for each element of a sample
architecture shown in Figure 2. All elements used in an ORC were tried to be
present in the sample and the equations were expressed in broad terms. We used
the database of EES software to evaluate the properties of fluids. Also, to calculate
the mixed fluid properties such as flue gas specific enthalpy, the following
equation is used [21]:

ℎ𝑚𝑖𝑥𝑒𝑑 𝑓𝑙𝑢𝑖𝑑 = ∑𝑥 .ℎ
𝑖 𝑖 (1) wh
ere,
𝑥𝑖 and ℎ𝑖 are the mole fractions and specific enthalpy of the flue gas components,
respectively.
Table 3. Equations governing energy and exergy analysis
Component Energy analysis Exergy analysis
W p  m 5 [(1  y )(h2  h1 )  (h5  h4 )]   Ex
Ex 
II , p  2 1

Pump 1 ( p 2  p1 ) 4 ( p 5  p 4 ) w p
I , p  
h2  h1 h5  h 4 E D , P  T 0 m 5 [(1  y)(s 2  s 1 )  (s5  s 4 )]
  Ex
Ex 
II ,E   6  5
Evaporator Q E  m 5 (h6  h5 )  m 10 (h10  h11 ) Ex 10  Ex 11
E D ,E  T 0 [m 5 (s 6  s 5 )  m 10 (s 11  s 10 )]
  Ex
Ex 
II ,C   13  12
Condenser QC  m 9 (h9  h1 )  m 12 (h13  h12 ) Ex 9  Ex 1
E D ,C  T 0 [m 8 (s1  s 9 )  m 12 (s13  s12 )]
  Ex
Ex 
Heat II ,IHE   3  2
Q IHE  m 8 (h8  h9 )  m 2 (h3  h2 ) Ex 8  Ex 9
exchanger
E D , IHE  T 0 [m 2 (s3  s 2 )  m 8 (s 9  s 8 )]

h 4  h3

Ex
y  II ,OFOH   4

Deaerator h 7  h3 Ex 3  Ex 7

E D ,OFOH  T 0 m 6 [s 4  ys 7  (1  y)s 3 ]
h6  h7 h  h8 w
Turbine or I ,T   7 II ,T   T
  Ex

h6  h7 s h7  h8s Ex 6  Ex 7 8
expander
w T  m 6 [(h6  h7 )  (1  y )(h7  h8 )] E D ,T  T 0 m 6 [ ys 7  (1  y )s 8  s 6 ]
In turbine output, the percentage of working fluid passing through path 7 versus total flow is represented by y in
equations.

Since material elements do not change in the cycle, the chemical exergy of the
system is constant. Hence, we did not express the equations of chemical exergy.
Figure 2. A sample architecture to help showing the equations governing the elements
3.2 Exergoeconomic
Given the importance of economic issues and the need for the estimation of costs
before system design finalization, it is unavoidable to target total annual costs.
Total annual costs are the sum of operation costs and capital cost (annualized
costs) [19]:

TAC  Operation Cost (OC)  Capital Cost (Z r ) (2)

where TAC represents total annual costs (in $/year), OC represents operation costs,
and CC denotes normalized costs of machinery purchase. The highest costs are
incurred by the purchase of waste heat boiler, expander, and condenser. The
present paper excludes fuel costs from operation costs because a gas turbine
system has been already existed and the new systems were mounted to improve the
cycle with no changes in fuel consumption. The only annual costs are the
maintenance costs of the newly mounted devices. Table 4 shows the equations
used to express the prices of main components including waste heat boiler,
expander, air condenser, and recuperator. Furthermore, the equations used to
calculate the heat transfer area of the heat exchangers are expressed in the
following [21]:
QHEX
AHEX = (3)
∆TLMTD
(Th,i ‒ Tc,o) ‒ (Th,o ‒ Tc,i)
∆TLMTD =
Ln (
(Th,i ‒ Tc,o)
(Th,o ‒ Tc,i) ) (4)
Table 4. An estimation of fixed capital costs [19-22]
Component Equation
  Q  Q Eva  QSup 
0.8 0.8 0.8
  
HRSG PEChrsg =6570    Eco
      ...   21276m f  m g 0.2
  T log,Eco 

 T log,Eva 

 T log,Sup  
 
Turbine PECT =W 
  1318.5-98.328  ln  W
 
A.C.C PECCon =8500  406  ACon
0.8

Pump 
PECPunp = 705.48 0.001 W p 
0.71 
1 
0.2
 1   pump
 


Recuperator PECRecup = 231.915 + (309.143 ∗ ARecup)
where Q denotes heat transfer rate, Tlog denotes logarithmic temperature, m f

represents organic fluid flow rate, m g is input gas flow rate into HRSG, W is

power, ACon and ARecup represent air condenser and recuperator area, respectively
and PEC shows per equipment cost expressed in dollars. Also, Th,i, Tc,o, Th,o and

Tc,i are hot and cold streams at the inlet and outlet of the heat exchangers.
Since equipment price functions are in terms of dollars, equipment purchase
costs are converted into annual costs by capital-recovery factor (CRF) [20, 21].

i1  i n
CRF  (5)
1  i n  1

where i is interest rate, and n is nominal life of the equipment. Costs can be
normalized through Equation (5).
Annual running time (H) of the new power plant is considered 8000 hrs assuming
accessibility of about 91%. Thus, we have [22, 23]

CRF .φ r .PECr
Z k.  (6)
3600.N

where φ is the maintenance factor that depends on power plant type. We assume it
to be 1.06 in this paper.
3.3 Economic Balance
The price balance equation can be written in a broad sense for a system that
receives heat and produces work as below:

∑C e,k + Cw,k = Cq,k + ∑C i,k + Zk (7)

where C is the cost rate in $/s and the subscripts e, I, w, and q denote the output
and input of work and heat written for the kth component, respectively. If CP is
related to the product costs and CF is related to the fuel costs of a system, they can
be defined as below [21]:

CF = Ci ‒ Ce (8)

Ex𝐹 = Exi ‒ Exe (9)

The cost per unit of fuel and product exergy can be defined for the system
components using these definitions [21, 23].
(10) CF,k
cf =
Ex𝐹,𝑘
(11) CP,k
cP =
Ex𝐹,𝑘
where CLis the cost rate of exergy loss; i.e., the monetary waste due to the exergy
exit from the system into environment. Obviously, exergy loss rate influences
exergy price rate of the products. Economic analysis aims to calculate the final
product prices for the assessment of the whole system optimization. When the
objective is to analyze the exergoeconomic, all exergy losses should be priced for
which the monetary waste due to exergy losses in the kth component of the system
is estimated. The equations are expanded as below [21]:
(12) CL,k = 𝑐𝐹,𝑘Ex𝐿,𝑘
(13) CP,k = CF,k ‒ CL,k + Zk
(14) 𝑐𝑃,𝑘Ex𝑃,𝑘 = 𝑐𝐹,𝑘Ex𝐹,𝑘 ‒ CL,k + Zk
The following exergoeconomic variables were considered for the assessment of the
system components: (i) mean cost per unit fuel exergy, 𝑐𝐹,𝑘; (ii) mean cost per unit
product exergy, 𝑐𝑃,𝑘; and (iii) cost rate of exergy destruction, CD,k.
(15) Ex𝐹,𝑘 = Ex𝑃,𝑘 + Ex𝐷,𝑘 + Ex𝐿,𝑘
𝑐𝑃,𝑘Ex𝑃,𝑘 = 𝑐𝐹,𝑘Ex𝐹,𝑘 ‒ CL,k + Zk
Removing Ex𝐹,𝑘 from the above equations will yield [19-23]:
(16) 𝑐𝑃,𝑘Ex𝑃,𝑘 = 𝑐𝐹,𝑘Ex𝑃,𝑘 + (𝑐𝐹,𝑘Ex𝐿,𝑘 ‒ CL,k) + Zk + 𝑐𝐹,𝑘Ex𝐷,𝑘
The last term in right side of the equation is exergy destruction rate that gives an
approximate estimation of exergy destruction cost. Assuming that the product
Ex𝑃,𝑘 is constant and the cost per unit exergy of fuel of the kth component of the
system is independent of exergy destruction, the exergy destruction cost can be
defined using the last term on the right side of Equation (11) [19-23]:
(17) CD,k = 𝑐𝐹,𝑘Ex𝐷,𝑘
Considering the equations of cost balance for the estimation of exergy destruction
cost of each individual component and given the fact that some components have
more than one single input and output and some parameters cannot be derived from
these equations, we need some supplementary equations. The placement of these
equations next to each other along with the supplementary equations forms a linear
equation system as below [19-23]:
(18) [Exk] × [ck] = [Zk]
where E xk  , ck , and Z 
k denote the matrix of exergy rate (derived from exergy
analysis), exergy cost vector, and the vectors of factors derived from economic
analysis, respectively.
3.4 Advanced exergy and exregoeconomic analysis
3.4.1 Advanced exergy
Advanced exergy analysis is performed on the basis of the results derived from
exergy analysis. Essentially, it is concerned with the classification of irreversibility
into endogenous and exogenous. Endogenous exergy destruction is a part of
destruction that is related to the intrinsic performance of a component while
exogenous exergy destruction is driven by other components pertaining to the
system. In addition, exergy can be divided into avoidable and unavoidable in terms
of potential. Unavoidable exergy is a part of exergy whose destruction cannot be
hindered due to technical and economic constraints so that it will always occur.
Avoidable exergy is a part that can improve system performance by improving the
involved factors. Avoidable exergy ( E DAV,k ) and unavoidable exergy ( E DUV,k ) are
defined and estimated as below and their sum expresses exergy destruction ( E D,k )
of an individual component [24-28]:
(19) E D ,k  E DUN,k  E DAV,k

 E
UN

(20) E DUN,k  E p ,k  D ,k 
 E
 P ,k 

where E p ,k is the exergy generated by each component. Also, exergy destruction

can be divided into endogenous ( E DEN,k ) and exogenous ( E DEX,k ) parts [24-28]:

(21) E D ,k  E DEN,k  E DEX,k

 E D ,k
UN

(22) E UN
D ,k  E p ,k   
 E P ,k 

Endogenous and exogenous exergies can be readily calculated through the


interaction and relationship of the process components. The constraints for
unavoidable exergy destruction calculation in the cycle state are as follows. It is
assumed that turbine efficiency is 95%, pump efficiency is 99%, and pinch and
approach temperature are 5°C [27].
Avoidable exergy of the kth component can be divided into endogenous and
exogenous. The same classification is applied to avoidable exergy destruction [24-
28].
 E D ,k
UN

(23) E DUN,k ,EX  E EN 
p ,k   
 E P ,k 

(24) E DUN,k ,EX  E DUN,k  E DUN,k ,EN

(25) E DAV,k ,EX  E DAV,k  E DUN,k ,EN

3.4.2 Advanced exregoeconomic


In advanced exergoeconomic analysis, the economic performance of the
individual components is examined in more comprehensive sense under interaction
with other components. This analysis improves researchers’ understanding of
system improvement trend. Also, it explores exergy cost and investment costs
pertaining to process components in these two perspectives. Exergy costs and
investment costs related to the kth component are divided into endogenous ( C DEN,k
and Z kEN ) and exogenous ( C DEX,k and Z kEX ) parts. Endogenous cost rate of the kth
component is calculated as below [25]:
(26) C DEN,k  c F ,k E DEN,k where
Z k
(27)  Z k 
real

Z kEN  E EN
 p ,k  E 
denot
 P ,k 
es the
kth equipment cost and c F ,k denotes the kth equipment fuel cost. The exogenous
exergy cost rate ( C DEX,k ) of the kth component is as follows [29].
(28) C DEX,k  c F ,k E DEX,k
(29) Z kEX  Z k  Z kEN Ex
ergy
destruction costs and the kth component costs investment are classified into
avoidable ( C DUN,k ) and unavoidable ( C DAV,k ) parts. The avoidable part depends on the
unavoidable part. The kth component cost rate is yielded from the following
equations:
(30) C DUN,k  c F ,k E DUN,k Th

(31) e rate
 Z
UN

Z UN
 E p ,k  k 
 k  E for the
 P ,k 
avoid
able part ( C DAV,k ) is, also, calculated in the same way as the other part [25]:
(32) C DAV,k  c F ,k E DAV,k To

(33) Z kAV  Z k  Z kUN gain


better
information than exergy cost rate and investment pertaining to the process, the
unavoidable and avoidable exergy costs and investment cost should be broken into
smaller parts. The cost of endogenous cost rate ( C DUN,k,EN ) pertaining to the kth
component is expressed as below given the technical and economic limitations:
(34) C DUN,k ,EN  c F ,k E DUN,k ,EN where

(35) the
 Z k
UN

Z UN , EN
 E EN
p ,k  
 k  E fuel
 P ,k 
cost is
the exogenous avoidable cost of equipment purchase ( Z kUN ,EN ). For exogenous
unavoidable part ( C DUN,k,EX ), we have:
(36) C DUN,k ,EX  c F ,k E DUN,k ,EX
(37) Z kUN ,EX  Z kUN  Z kUN ,EX And it
is
expressed as below for avoidable part [20, 24]:
(38) C DAV,k ,EN  c F ,k E DAV,k ,EN where

(39) Z kAV ,EN  Z EN  UN ,EN Z kAV ,EX


p ,k  Z k

(40) C DAV,k ,EX  c F ,k E DAV,k ,EX


denot
es the
(41) Z kAV ,EX  Z kEX  Z kUN ,EX
avoid
able endogenous cost of the equipment purchase, Z kUN ,EX represents the avoidable
exogenous cost of the equipment purchase, and Z kEX shows the cost of avoidable
part of the system.
4. Optimization
The optimization was performed by a genetic algorithm in this work. Readers
are cited to [30]-[35] for more details as so how genetic algorithm works. In
optimization, we assumed the variation ranges for the decision parameters as
expressed in Table 5. Also, the objective functions of this optimization were
defined as below. The first function includes system costs and the second one is
total exergy efficiency of the system.
(42) Object 1 Min  Cos t  TAC  C D

(43) Ex D Total  Ex L Total


Object  2  Max  II ,Total  1 
Ex f Total
Table 5. The variation ranges of the decision parameters

Parameter Unit Low boundary High boundary


Condenser pressure kPa 100 200
Input working fluid pressure kPa 1400 2100
Input working fluid temperature °C 110 155
Pinch temperature °C 5 100
Approach temperature °C 5 100
5. Case study
The case was a gas turbine in NGL Power Plant of Siri Island in the Persian
Gulf. The environmental conditions of the power plant where the turbine is located
are summarized in Table 6. In this site, the turbine exhaust gas temperature equals
272°C after flowing into a heat exchanger. The plant uses a part of this heat for its
own process in the exchanger cycle (HEX 1), which reduces the temperature of the
combustion products down to 272°C. Nominal parameters of commercial gas
turbines are presented in Table 7.
Table 6. Environmental conditions of the study site
Parameter Unit Value
Mean maximum temperature °C 40
Mean minimum temperature °C 15
Highest recorded temperature °C 45
Lowest recorded temperature °C 8
Maximum moisture % 100
Minimum moisture % 50
Altitude m 1.25
Table 7. Nominal parameters of commercial gas turbine.
THM1304-11
Manufacture MAN
Shaft power output (kW) 11200
Efficiency (%) 30.8
6. Validation
The model of a conventional Rankine cycle with an internal recuperator (cycle
b, Figure 3) can be developed based on the above equations while its reliability
should be confirmed. The reliability of the cycle is validated by comparing with
results of other publications. Table 8 presents validation results for the
conventional Rankine cycle with internal recuperator. Ref. [31] investigated a
Rankine cycle with and without internal recuperator for waste heat recovery. To
validate the computations, a set of energy and exergy balance equations were
solved using the same input conditions as those in Saleh et al. [31]; 𝑇1 = 303.15 𝐾
, 𝑇7 = 373.15 𝐾, 𝑇9 = 303.15 𝐾, 𝑊 = 1000 𝑘𝑊, 𝜂𝑝𝑢𝑚𝑝 = 65%, and 𝜂𝑡𝑢𝑟𝑏𝑖𝑛𝑒

= 85%. This cycle is built by our modeling method, the relative error is less than
6% (Table 8). This difference may be induced by the approaches used to calculate
the properties of the working fluid. These indicate that simulation results of the
present study can be consistent with the published data. Therefore, it seems that the
proposed model can evaluate the performance of the conventional Rankine cycle
with internal recuperator within acceptable error range.
Table 8. Results for model validation of organic Rankine cycle with internal recuperator
Thermal efficiency (%) Mass flow rate (kg/s)
Fluid Present Work [31] Present Work [31]
iso-Butane 12.44 12.43 19.121 20.423
n-Butane 13.01 13.07 17.213 17.746
R-245fa 13.01 13.07 32.556 33.424

Tg 15
5

Eva

Tg 4

Tg 12

7
Sup 6
3
Turbine
4
Eco
8 Generator
Heat
Exchanger
3

9
2
Condenser
Tg 13
1

Figure 3. A schematic view of a cycle modeled in EES Software Package

7. Solution Algorithm

The flowchart for full comparison of cycle and working fluid is followed as
depicted in Figure 4. First, the working fluid is selected with respect to the
environmental impacts and the related heat source temperature. Then, the optimum
parameters are selected with respect to the objective functions of cost and exergy
efficiency using a genetic algorithm, taking into account the constraints governing
the cycle and the output conditions of combustion products exiting the furnace
(including heat capacity, enthalpy, exhaust gas flow rate, and its compounds
percentage). Higher exergy efficiency results in the quality improvement of cycle
energy. Given the equal temperature of hot and cold heat sources and so, their
equal Carnot efficiency, the selection of a single parameter as the objective
function would improve system performance at the same conditions of heat
sources. Eventually, the cycle exhibiting higher exergy efficiency and lower costs
would be recommended as the cycle to be used in the output of heat exchangers.
Environmental Impact

ODP
Yes
ON

GWP

Selection fluid

R123

150 2100 kPa

100
parameters GA OPT Cycle
T [°C]

50

0
90 kPa
0.2

0.4 0.6 0.8

1.00 1.25 1.50 1.75 2.00 2.25 2.50

s [kJ/kg-K]

Cost

Exergy Selection OPT cycle End


Object
Efficiency

Power

Figure 4. The flowchart to determine optimum cycle


8. Results
8.1 Architecture Selection
Given the existence of various architectures in organic cycles and a brief
review of literature, the cycles were picked up with a focus on cutting the
calculations and modeling on the basis of the research results and studies. The
cycles were examined with respect to the desired thermal ranges. Considering the
thermal range, the architectures were compared to figure out which had higher
output power and efficiency. Also, among the selected cycles, a dual-pressure
cycle with two different fruits was compared as an innovative cycle with other
ones.
As is evident in [8-10], a simple cycle with a recuperator has been mostly used in
the thermal range close to what considered in our study. The point to note about
these cycles is the cold source temperature that is determined on the basis of the
fluid type and ambient conditions. The cooling system is air-based in most organic
cycles. Thus, mean ambient temperature is an important parameter in organic fluid
determination.
As was mentioned, cycles introduced in [18] have extensive applications. [15-18]
have stated that the cycles have been mostly used in exploiting the temperature of
the exhaust gases in the thermal recovery of cycles with pre-heating.
Cycle 5a is a dual-pressure cycle with the R123 fluid in which the fluid is
superheated in high-pressure (HP) section and flows into the turbine. Considering
the T-S curve of this fluid, the turbine output in low-pressure (LP) section will be
in superheat zone. So, a heat exchanger can be applied for pre-heating of the
organic fluid.
The next recommended cycle is presented in Figure 5b. In this cycle, once R123
and then R600 are selected as the working fluid. They flow into the turbine in
superheated state and since the turbine output is in the superheated zone, it is fed
into heat exchanger so as to preheat the organic cycle flowing into HRSG and to
cool the superheat fluid down to saturated steam point.
The cycle presented in Figure 5c is based on various literature. Research has
shown that in different thermal ranges for heat sources, R123 is appropriate for
high temperatures and the other fluids including R600 are appropriate for lower
temperatures. In this cycle, R123 has been selected as the working fluid in HP
section and R600 as the working fluid in LP section.
These three cycles are contrasted in heat efficiency, power generation, total price
of the produced electricity, and exergy efficiency. The cycle that could optimize
these parameters under the existing and optimum conditions will be recommended
as the optimum cycle for the turbine.

Figure 5. A schematic illustration of the recommended cycles modeled on the bottoming cycle
of a gas turbine

8.2 Energy, exergy and exregoeconomic analysis


The general specifications that are common in tables include these: condenser
pressure is above atmospheric pressure in all three fluids. Then, there is no need to
mount an ejector, cutting the costs because higher pressure in condenser than the
atmosphere hinders the penetration of air into the condenser and the disruption of
the condensation. One point that can be observed in all three figures is the quality
of output working fluid. This fluid quality reduces condenser pressure and prevents
the breakdown of steam turbine blades. Given the constraints considered for
condenser, its pressure cannot be decreased. This influences power generation and
exergy destruction rate of the turbine. A point that should be noted about the
quality of output working fluid is its superheater zone in which a pre-heating is
applied to enhance the efficiency and also, to reduce the cooling fluid flow rate in
condenser.
One another advantage of the use of organic fluids is the application of low-
volume turbine, which is a competitive advantage for organic fluids because of the
space limitations in gas turbine cycles.
8.2.1 Cycle a
Table 10 presents the results of Figure 5a. Considering the local temperature
limitation, the condenser pressure is set to 130 kPa in this case. Then, the net
generated power would be 2.2 MW. With respect to the limitation of acidifying
temperature of combustion products that destroys economizer, acidifying
temperature for gas fuel was set to 105°C according to [2]. To this end and
considering input conditions, the pinch and approach temperatures are 27.01 and
73.55 in HP section and 7 and 31.7 in LP section, respectively.
Table 10. Results of optimization of Figure 5a
Point Fluid T (°C) P (kPa) h (kJ/kg) s (kJ/kg.K) Ex (kW)
1 R123 34.85 130 236.5 1.125 10.97
2 R123 35.65 1590 237.7 1.126 62.04
3 R123 60.76 1590 264.5 1.212 120.8
4 R123 60.94 1590 264.7 1.211 134.8
5 R123 134.5 1590 351.2 1.441 1026
6 R123 134.5 1590 459.7 1.708 2476
7 R123 140.8 1590 466.6 1.724 2573
8 R123 70.75 130 430.8 1.751 398
9 R123 34.85 130 404 1.669 277.8
10 R123 51.39 1064 254.3 1.181 23.44
11 R123 82.25 1064 288.2 1.279 101.4
12 R123 114 1064 325.4 1.378 235.1
13 R123 114 1064 449.9 1.7 725.2
8.2.2 Cycle b (R123)

The parameters of the cycle modeled in Figure 5b are presented in Table 11


and Figure 7. The interesting point about this cycle is that the output temperature
of the superheater is for organic working fluid.

hgas (kJ/kg)
120 150 180 210 240 270 300 330
300
Hot Gas
250 R123 (HP)

200 R123 (LP)

150 5 1590 kPa 6 7


T (°C)

W T,H P = 1.47MW

100 Tgo u t =110 °C 1064 kPa 13


4 12
W T,LP = 0.49 MW
3 11
50 W P ,H P = 0.06MW 2 10 8
1
W P ,LP = 0.01MW 130 kPa
0 1 9

-50
0.8
-100
0.75 1.00 1.25 1.50 1.75

s (kJ/kg.K)
Figure 6. T-S diagram for Figure 5a cycle

As working fluid temperature rises, the heat transferred to working fluid in pre-
heating section increases. Consequently, the energy in HRSG section will be spent
more on increasing mass flow rate and power. Sensitivity analysis depicted in
Figure 8 reveals that pinch and approach temperatures become decisive in this
state. As output temperature of the superheater is increased, more heat is taken
from the combustion products in HRSG section. But, the recovery from smoke is
not appropriate because of the constant pinch and approach temperatures. So, the
HRSG production rate is lost and the temperature of output smoke of HRSG is
increased. Wnet vs. T7 curve is drawn with respect to the increase in superheater
output temperature with the decrease in flow and their contrast. In this case, this
cycle would have a net power generation of 2.31 MW.
Table 11. Results of Figure 5b
Point Fluid T (°C) P (kPa) h (kJ/kg) s (kJ/kg.K) Ex (kW)
1 R123 34.85 130 236.5 1.125 13.88
2 R123 35.77 1823 237.9 1.126 88.77
3 R123 59.17 1823 262.8 1.207 140.7
4 R123 82.13 1823 288.1 1.279 377.4
5 R123 141.9 1823 361 1.465 1482
6 R123 141.9 1823 462.7 1.71 3285
7 R123 144.8 1823 466.1 1.718 3346
8 R123 68.29 130 428.9 1.745 487.5
9 R123 34.85 130 404 1.669 351.5
hgas (kJ/kg)
125 165 205 245 285 325

250

200

150 5 7
T (°C)

1822.77 kPa 6
WT= 2.34 MW
100 Tg ,o u t = 115 °C 4
3 8
50 Wp= 0.098 MW
2

130 kPa 9
0 1

-50
0.8

0.75 1.00 1.25 1.50 1.75

s (kJ/kg.K)

Figure 7. T-S diagram for the cycle depicted in Figure 5b


8.2.3 Cycle b (R600)
Following the examination of the cycle depicted in Figure 5b with the R600 fluid,
the results of cycle energy analysis are tabulated in Table 12. The T-S diagram for
the R600 fluid is illustrated in Figure 9. The shaded parts of this figure display
entropy generation in each part of HRSG. In the R600 fluid, the lowest entropy
generation is related to superheat and then, to evaporator.
5.8 190

5.6 W net
180
Tg,out
5.4
170
5.2
160
Wnet (MW)

Tg,out (°C)
5
150
4.8
140
4.6
130
4.4

4.2 120

4 110
140 160 180 200 220 240 260
T7 (°C)

Figure 8. Sensitivity analysis of net output power and output temperature of HRSG with the
variations of superheat working fluid temperature.
Table 12. Results of Figure 5b
Point Fluid T (°C) P (kPa) h (kJ/kg) s (kJ/kg.K) Ex (kW)
1 R600 32.96 310 278.9 1.271 805.7
2 R600 34.33 2300 283.4 1.275 885
3 R600 78.23 2300 398.4 1.623 1128
4 R600 107.2 2300 484.1 1.863 1447
5 R600 122.2 2300 534.7 1.988 1745
6 R600 122.2 2300 742.5 2.514 2885
7 R600 150 2300 823.6 2.712 3375
8 R600 91.93 310 746.4 2.766 1294
9 R600 32.96 310 631.5 2.423 1011
Unlike water that constitutes the main part of the evaporator, this component in
organic fluids exhibits lower exergy destruction than control due to its
thermodynamic properties and the loss of latent heat of vaporization. As can be
seen, the net power of this cycle is 1.72 MW which is lower than that of the similar
R123 cycle.

8.2.4 Cycle C
Table 13 presents the results of Figure 5c. With respect to the suitable thermal
range for the performance of the selected fluids, R123 was used in HP section of
the cycle.
hgas (kJ/kg)

100 150 200 250 300 350


300
R600
Hot Gas
250

200

150 7
Tpinch = 25.95 °C
T (°C)

100 5 2300 kPa 6 W T= 1.72 MW


Tout = 112°C 4 8

3
50 2 310 kPa
W T= 0.1 MW
1 9
0

-50
0.2 0.4 0.6 0.8

-100
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
s (kJ/kg.K)

Figure 9. The T-S diagram of the R600 fluid in cycle b

Also, R600 was used in LP section given the temperature loss of combustion-
generated gases. This cycle, finally, generated 2.26 MW of power.
Table 13. Results inferred for Figure 5c
Point Fluid T (°C) P (kPa) h (kJ/kg) s (kJ/kg.K) Ex (kW)
1 R123 34.85 130 236.5 1.125 11.62
2 R123 35.65 1590 237.7 1.126 65.66
3 R123 60.09 1590 263.8 1.21 123.4
4 R123 60.94 1590 264.7 1.211 142.7
5 R123 134.5 1590 351.2 1.441 1086
6 R123 134.5 1590 459.7 1.708 2621
7 R123 140 1590 465.7 1.722 2710
8 R123 69.8 130 430 1.749 416
9 R123 34.85 130 404 1.669 294.1
10 R600 31.85 300 276.2 1.262 272.7
11 R600 33.06 2064 280.1 1.265 296.4
12 R600 50.43 2064 323.9 1.413 294
13 R600 84.34 2064 415.7 1.68 387.2
14 R600 116.1 2064 514 1.937 551
15 R600 116.1 2064 737.5 2.511 947.2
16 R600 55.04 300 673.6 2.56 352.1
17 R600 34.85 130 236.5 1.125 11.62
hgas (kJ/kg)
180 210 240 270 300 330
300
R123 (HP) Tg2

250 Hot Gas Tg1

200
Tg3
Tpinch = 21.5 °C
150 7
T (°C)

Tg4 6 WT ,HP =1.8 MW


100 5 1590 kPa
4
3
50 WP,HP =0.06 MW 2 8
9
0 1

-50
0.8

0.75 1.00 1.25 1.50 1.75 2.00

s (kJ/kg.K)

Figure 10. The T-S diagram for the HP section of the cycle depicted in Figure 5c

Figure 10 shows the T-S diagram for the HP section and Figure 11 for the LP
section. As a result, Table 14 can be presented for the three cycles from energy and
economy analysis perspective. As indicated in this table, cycles a and b have the
desired power output at an appropriate expense.

Cycle a, cycle b with two fluids of R123 and R600, and cycle c are compared
in terms of exergy destruction in Figure 12. Low power generation in cycle b-R600
and lower exergy destruction than other cycles has caused the interaction of these
two parameters to slightly increase the exergy efficiency of this cycle as compared
to other cycles. The comparison reveals that cycles a and b with R123 have the
highest exergy efficiency.
hgas (kJ/kg)
80 100 120 140 160 180 200 220

R600 (LP)
Hot Gas
150
15
Tpinch = 7 °C
100 13 14 2064 kPa WT ,L P =0.5 MW
Tgo u t = 110.1 °C
T (°C)

12
50 WP,L P =0.02 MW 11 16

10 300 kPa 17
0

-50
0.8

-100
0.5 1.0 1.5 2.0 2.5 3.0

s (kJ/kg.K)
Figure 11. T-S diagram for LP section of the cycle depicted in Figure 5c
Table 14. Results of energy and economy analysis of the studied cycles
Power Thermal efficiency (%) Price of generated
Working
Cycle generation With Gas turbine electricity
fluid
(MW) recovery cycle (cent/kW)
Cycle a (dual-pressure with
R123 2.2 36.56 28.8 14.59
preheating)
Cycle b (single-pressure with R123 2.31 36.90 28.8 13.03
preheating) R600 1.72 35.07 28.8 21.95
Cycle c
R123 &
(dual-pressure with two 2.26 36.75 28.8 16.78
R600
working fluids)
According to the currents in the compared cycles, the cost of each current and the
rate of exergy destruction are expressed for each individual component of each
cycle. It is observed that the cost of exergy destruction for each component is
affected by the rate of exergy destruction of that component. Two HRSG and
turbine components have the highest rate of exergy destruction among cycle
components. As is evident in Tables 15 to 17, the values of Ck and the cost of
exergy degradation of each component are calculated by solving an equations
system.

Figure 12. Comparison of total exergy destruction and exergy efficiency of cycles examined
with different working fluid
Table 15. The cost rate of each current and the cost of exergy destruction and fuel of each individual
component in cycle a
The studied fluid R123
C k C D C F
Component Flow name
($/h) ($/h) ($/h)
Input to pump 4.92
Pump Consumed work 43.42 7.91 0.61
Output from pump 48.31
Heat exchanger Output from exchanged (cold flow) 119.30 46.51 0.33
Input to LP section 19.39
Input to HP section 99.90
HRSG 2953.44 0.57
Input gas to HRSG 21.54
Output gas from HRSG 3.29
Input to HP turbine 1421.57
Input to HP turbine 309.17
Turbine 250.70 0.43
Produced work 1268.28
Turbine output 171.58
Output air from condenser 185.04
Condenser 37.62 0.33
Input to condenser 124.52
Table 13 illustrates cycle b in two fluid states. As can be observed, R600 has lower
exergy destruction cost than R123. Table 14 presents cycle c with two fluids: R600
for LP section and R123 for HP section.
Table 16. The cost rate of each current and the exergy destruction cost of each individual component in
cycle b
Studied fluid R600 R123
C k C D C k C D
Component Flow name
($/h) ($/h) ($/h) ($/h)
Input to pump 5.52 0.37 0.19
Pump Output from pump 7.16 2.48 0.43
Consumed work 1.64 2.29
Heat exchanger Output from exchanged (cold flow) 13.65 4.76 8.86 1.84
Input gas to HRSG 21.54 91.11 21.54
HRSG 344.44
Output gas from HRSG 26.24 12.88
Input to HP turbine 23.13 7.66 45.29
Turbine Produced work 8.87 6.60 8.26
Turbine output 28.37 57.17
Output air from condenser 105.01 1.32 66.71
Condenser 1.43
Input to condenser 6.93 4.76
Table 17. The cost rate of each current and the exergy destruction cost of each individual component in
cycle c
R600 R123
The studied fluid
(Part LP) (Part HP) C D
C k C k ($/h)
Component Flow name
($/h) ($/h)
Input to pump 3.007 0.111
Pump Output from pump 4.5252 1.47 0.65
Consumed work 1.518 1.362
Heat exchanger Output from exchanged (cold flow) 17.14 7.18 16.00
Input gas to HRSG 14.60 21.54
HRSG 646.2
Output gas from HRSG 12.70 14.60
Input to HP turbine 30.93 26.00
Turbine Produced work 24.16 37.87 21.67
Turbine output 11.84 3.99
Output air from condenser 3.77 2.821
Condenser 1.3
Input to condenser 43.05 55.8
Figure 13. Comparison of the exergy destruction cost of the studied cycles

Finally, Figure 13 compares the cost of exergy destruction of the studied cycles. In
this figure, cycle a and cycle b with two fluids of R600 and R123 have the lowest
destruction cost. The figure shows that the cost of exergy destruction with R600 is
the lowest. To select the optimized cycle, the results of energy, exergy and
exergoeconomic analysis are compared and explored:

- The cost of exergy destruction among the cycles with the highest generated
power is the lowest for R123. Cycle a and cycle b have lower costs of exergy
destruction.

- Exergy efficiency is the highest for cycle b with the R123 fluid and cycle a.

- Thermal efficiency and total price of electricity is more suitable in cycle a


and cycle b with R123 than other cycles.
According to the results, the parameters are optimized for two cycles a and b
with the R123 working fluid so as to enhance the total price of electricity and
exergy efficiency by reducing the cost of exergy destruction and selecting the
appropriate parameters. These two cycles are optimized with a genetic
algorithm. Figure 14 depicts two Pareto curve derived from the optimization of
these two cycles. To determine the ideal point from two points A and C, which
indicate the lowest exergy efficiency and cost and the highest efficiency and
cost, respectively, two tangent lines are plotted on the curve, and the intersection
of these two points show the ideal point with the lowest cost and highest
efficiency. A point of the curve that is closer to this point (B) is the best point of
Pareto curve.

Figure 14. The Pareto curve derived from the optimization of two cycles on the basis of the
objective functions of the price of the generated electricity and exergy efficiency using
MATLAB Software Package
The comparison of two Pareto curves of the two cycles shows that cycle b with the
optimum selected parameters has the best performance as compared to cycle a.
According to Figure 14, it is seen that in the same efficiency, cycle b achieves this
efficiency with lower cost, which includes the cost of exergy destruction of
maintenance and equipment purchases. Table 18 indicates that the optimum
parameters selected by the genetic algorithm.
Table 18. The values of the optimum selected parameters of point B on the Pareto curve of cycle b-R123

Parameter Unit Optimum value Pre-optimum


Condenser pressure kPa 125.02 130
Input pressure to turbine kPa 1722.22 1822.77
Input temperature of turbine °C 146.33 144
Pinch temperature °C 12.65 11.21
Approach temperature °C 38.55 59.8

Figure 15 compares cycle b under optimum state and the designed state in terms of
exergy destruction of each individual component. It shows that the genetic
algorithm has reduced exergy destruction in HRSG by selecting appropriate pinch
and approach temperature and has also reduced entropy generation in the turbine
by improving its performance. Overall, it has decreased the exergy destruction of
the cycle.

The cost of exergy destruction has been reduced from $346 to $233 under optimum
state, which is influenced by the decrease in exergy destruction rate of the cycle.

8.3 Advanced exergy and exergoeconomic analysis


Dividing the exergy loss within the kth component into the endogenous and
exogenous parts enables us to estimate the loss of exergy induced by the
inefficiency of the component itself and that induced by the inefficiency of other
components separately. Endogenous exergy loss occurs only due to irreversibility
in the kth component. This loss of exergy occurs when other components are acting
in their ideal states and the intended component is acting with its exact efficiency.
The separation of exergy losses into unavoidable and avoidable parts in the kth
component provides a realistic measure of the ability to improve the
thermodynamic efficiency of a component. The rate of exergy destruction that
cannot be reduced due to the technological constraints such as material availability
and cost and production methods is unavoidable.
The endogenous avoidable part of exergy destruction can be reduced by changing
the desired component, while the exogenous avoidable part of exergy destruction
describes a part of the exergy destruction in the kth component that can be reduced
by improving the productivity of the remaining components and/or the overall
structure of the system.
Given the fact that it is important to divide exergy degradation and determine the
avoidance factors for economic savings, this section explores the organic cycle
using two concepts of advanced exergy and advanced exergoeconomic in Table 18
and 19 for the first time.
Table 18. The results of advanced exergy for the optimized cycle

Ex D Ex UN
D ,k
Ex AV
D ,k
Ex EN
D ,k
Ex EX
D ,k
Ex UN,EN
D ,k
Ex UN,EX
D ,k
Ex AV,EN
D ,k
Ex AV,EX
D ,k

(kW) (kW) (kW) (kW) (kW) (kW) (kW) (kW) (kW)


Con 93.8 92 1.8 93.8 0.0047 83.42 8.57 1.632 0.167
HRSG 5053 3413 1640 2805 2248 3095.72 317.27 1487.5 152.45
Per 145.8 119.2 26.58 91.79 54.01 100.40 18.79 19.58 6.99
Pump 15.74 12.64 3.1 0.5591 15.18 11.55 1.089 2.83 0.267
Turbine 538 275.3 260.1 48.98 486.4 196.9 78.33 186.08 74.01
Table 19. Normalized cost of equipment on the basis of advanced economics

Z  EN Z EN  EX  EX
Z  UN  UN
Z  AV
Z  AV  UN,EN
Z  UN,EX
C  UN,EX  UN,EX
C  AV,EN  AV,EN
C  AV,EX  AV,EX
k C D k C D k C D k k C D k D Z k D Z k D Z k
C D
($/h) ($/h) ($/h) ($/h) ($/h) ($/h) ($/h) ($/h) ($/h) ($/h) ($/h) ($/h) ($/h)
($/h) ($/h) ($/h) ($/h)

Con
67.52 1.35 61.23 0.00 6.29 1.33 67.48 0.03 0.03 61.20 0.12 6.29 0.12 0.03 0.02 0.00 0.00

HRSG
25.26 68.39 22.91 54.81 2.35 83.22 25.25 0.01 39.99 22.91 7.74 2.35 7.74 0.01 36.27 0.00 3.72

Per
4.54 1.33 3.83 0.78 0.72 1.72 3.92 0.62 0.38 3.31 0.27 0.62 0.27 0.52 0.28 0.10 0.10

Pump
0.00 0.00 0.00 0.07 0.00 0.05 0.00 0.00 0.01 0.00 0.00 0.00 0.00 0.00 0.01 0.00 0.00

Turbine
16.92 0.71 12.11 7.02 4.81 3.98 14.88 2.04 3.76 10.64 1.13 4.23 1.13 1.46 2.69 0.58 1.07

Consequently, advanced exergoeconomic and advanced exergy only determine


that, in the optimal states of the system, whether the exergy destruction and the
exergy destruction cost are related to avoidable or unavoidable part and/or if they
are exogenous or endogenous. This is just a parameter for system analysis. Thus,
this analysis is applied only in system optimization in order to make optimum
system analysis possible. As can be observed, the share of exergy destruction
induced by other components has decreased. Referring to Table 18, it can be seen
that the value of ExEN EX
D is greater than the value of Ex D in all system components

except the pump and turbine. This means that the main part of exergy destruction
rate in each of these components is originated from irreversibility of the
component itself. The pump and turbine have the highest exogenous exergy
destruction rates. Therefore, a modification in the other component efficiencies can
lead to a reduction in these exergy destruction rates and brings about an
improvement in cycle efficiency. Table 18 indicates that the avoidable exergy
destruction is less than the unavoidable part for all components. This means that
the efficiencies of these components cannot be improved by using technical
modifications and modern technologies or by replacing these components with
more efficient ones. Splitting exergy destruction rates of the main system
components into endogenous, exogenous, avoidable, unavoidable, endogenous-
avoidable, endogenous-unavoidable, exogenous-avoidable and exogenous-
unavoidable parts are shown in Figure 16. Also, contributions of the components in
total endogenous, exogenous, avoidable, unavoidable, endogenous-avoidable,
endogenous-unavoidable, exogenous-avoidable and exogenous-unavoidable exergy
destruction rates of the system are indicated in Figure 17.
Condenser

HRSG

Recuperator
Pump

Turbine

Fig 16. Splitting exergy destruction rates of the system components into endogenous, exogenous,
avoidable, unavoidable, endogenous-avoidable, endogenous-unavoidable, exogenous-avoidable
and exogenous-unavoidable parts
Fig 17. Contributions of the system components in the total endogenous, exogenous, avoidable,
unavoidable, endogenous- avoidable, endogenous-unavoidable, exogenous-avoidable and
exogenous-unavoidable exergy destruction rates of the system

According to Figures 16 and 17 and Table 18, the HRSG and turbine should be
first focused on in order to optimize the cycle exergetic performance. This can be
explained if we consider that from 5053 kW of total exergy destruction rate in the
HRSG, 2805 kW (55.51% ofExD,HRSG) is related to the component itself and at

most, 29.4379% of ExD,HRSG (ExEN,AV


D 1487.5kW) can be reduced by technological

improvement or technical modifications. Also, ExEN,AV


D value for turbine is 34.587
kW (34.587% of ExD,Turbine). With the same analysis and Table 19, from 16.92 $/h
of exergy destruction cost rate, 12.11 $/h is related to the component itself and at
most, 1.46 $/h can be reduced by technological improvement or technical
modifications. Also, 𝑍𝐸𝑁,𝐴𝑉
𝐾 value for recuperator is 0.52 $/h.

9. Conclusion
In this paper, we tried to select an organic cycle from the existing organic
cycles with different functions in order to use it for heat recovery. The cycle was
selected according to the review of literature based on different thermal sources.
The results indicate the extensive application of the organic cycle with pre-heating.
The application of this cycle will improve efficiency and reduce system costs. On
the other hand, the choice of the working fluid for this cycle is important given the
turbine exhaust conditions. This paper picked up two fluids of R600 and R123 on
the basis of the thermometric properties of organic fluids, as well as environmental
impacts such as ozone depletion and heating, and compared them in a simple cycle
with pre-heating and single-pressure and dual-pressure cycle. Although the R123
fluid did not have optimum environmental conditions as compared to the R600, it
was selected for comparison due to its prominent features, such as its extensive use
in the heat recovery cycles and the temperature range of its proper operation (200-
270°C).
According to the exergy analysis, exergoeconomic analysis, and cost estimation
including the costs of purchasing, maintenance, and exergy destruction, cycles a
and b were selected for optimization based on the following two criteria: 1) high
exergy efficiency, 2) low cost of power generation.
According to the Pareto curve, cycle b was selected as the optimum cycle with the
R123 fluid because of its optimum efficiency and lower costs. Exergy destruction
cost was reduced from $346 to $233 per hour in optimum state.
Finally, this optimum cycle was analyzed for advanced exergoeconomic and
advanced exergy. The results showed that the genetic algorithm reduces the
avoidable exergy of the system to the extent possible, which implies the optimal
performance of the components relative to each other. In addition, bottoming
organic Rankine cycles are favorable compared to single stage cycles. But, their
main disadvantage is their complex layout and increased costs since they
essentially consist of many equipment components. As a result, sufficiently
significant thermodynamic performance enhancement is a prerequisite to justify
their application.

References
[1] Wei D, Lu X, Lu Z, Gu J. “Performance analysis and optimization of organic Rankine Cycle
(ORC) for waste heat recovery”. Energy Convers Manage 2007; 48:1113–9.
[2] Hamid Mokhtari, Hossein Ahmadi sedigh, Mohammad Ameri, The optimal design and 4E
analysis of double pressure HRSG utilizing steam injection for Damavand power plant,
Energy 118 (2017) 399-413.
[3] Hamid Mokhtari, Hossein Ahmadi sedigh, Iman Ebrahimi, Comparative 4E analysis for solar
desalinated water production by utilizing organic fluid and water, Desalination 377 (2016)
108–122
[4] Alireza Haghighat Mamaghani, Behzad Najafi, Ali Shirazi, Fabio Rinaldi, 4E analysis and
multi-objective optimization of an integrated MCFC (molten carbonate fuel cell) and ORC
)organic Rankine cycle) system, Energy 82 (2015) 650-663.
[5] Toffolo, Andrea Lazzaretto, Giovanni Manente, Marco Paci , A multi-criteria approach for
the optimal selection of working fluid and design parameters in Organic Rankine Cycle
systems, Andrea, Applied Energy 121 (2014) 219–232.
[6] Haoshui Yu, Xiao Feng, Yufei Wang, A new pinch based method for simultaneous selection
of working fluid and operating conditions in an ORC (Organic Rankine Cycle) recovering
waste heat, Energy xxx (2015) 1-11.
[7] Lecompte, Henk Huisseune, Martijn van den Broek, Bruno Vanslambrouck, Michel De
Paepe, Review of organic Rankine cycle (ORC) architectures for waste heat recovery Steven,
Renewable and Sustainable EnergyReviews 47 (2015) 448–461.
[8] Xinxin Zhang, Maogang He, Jingfu Wang, A new method used to evaluate organic working
fluids, Energy 67 (2014) 363-369
[9] Huijuan Chen, D. Yogi Goswami, Elias K. Stefanakos , A review of thermodynamic cycles
and working fluids for the conversion of low-grade heat, Renewable and Sustainable Energy
Reviews 14 (2010) 3059–3067.
[10] Junjiang Bao,Li Zhao , A review of working fluid and expander selections for organic
Rankine cycle Renewable and Sustainable Energy Reviews 24 (2013) 325–342.
[11] Yue Cao, Yike Gao, Ya Zheng, Yiping Dai ,Optimum design and thermodynamic analysis
of a gas turbine and ORC combined cycle with recuperators , Energy Conversion and
Management 116 (2016) 32–41.
[12] 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
[13] Stefano Clemente Diego Micheli, Mauro Reini, Rodolfo Taccani , Bottoming organic
Rankine cycle for a small scale gas turbine: A comparison of different solutions, , Applied
Energy 106 (2013) 355–364.
[14] NIST National Institute of Standards and Technology, Thermophysical properties of fluid
systems, USA, 2016 (http://webbook.nist.gov/chemistry/fluid/).
[15] Thermodynamics properties of DuPontTM ISCEON® MO29, DuPontTM ISCEON® 9
Series:
Refrigerants, 2005 (https://www.chemours.com).
[16] Thermodynamics properties of DuPontTM ISCEON® MO59 (R417a), DuPontTM
ISCEON® 9
Series: Refrigerants, 2005 (https://www.chemours.com).
[17] Thermodynamics properties of DuPontTM ISCEON® MO79 (R422a), DuPontTM
ISCEON® 9
Series: Refrigerants, 2005 (https://www.chemours.com).
[18] Thermodynamics properties of DuPontTM ISCEON® 39TC (R423a), DuPontTM
ISCEON® 9
Series: Refrigerants, 2005 (https://www.chemours.com).
[19] M. Ameri, P. Ahmadi, A. Hamidi, Energy, exergy and exergoeconomic analysis of a steam
power plant: a case study, Int. J. Energy Res. 33 (2009) 499–512.
[20] Hamid Mokhtari, Hasti Hadiannasab, Mostafa Mostafavi, Ali Ahmadibeni, Behrooz
Shahriari, Determination of optimum geothermal Rankine cycle parameters utilizing coaxial
heat exchanger, Energy 102 (2016) 260-275.
[21] Bejan A, Tsatsaronis G, Moran M. Thermal design and optimization. New York: Wiley;
1996.
[22] El-Emam RS, Dincer I. Exergy and exergoeconomic analyses and optimization of
geothermal organic Rankine cycle. Appl Therm Eng 2013;59:435-44
[23] Pouria Ahmadi, Ibrahim Dincer, Marc A. Rosen, Exergy, exergoeconomic and
environmental analyses and evolutionary algorithm based multi-objective optimization of
combined cycle power plants, Energy 36 (2011) 5886e5898
[24] Fateme Ahmadi Boyaghchi, Hanieh MolaieInvestigating the effect of duct burner fuel mass
flow rate on exergy, destruction of a real combined cycle power plant components based on
advanced exergy analysis, Energy Conversion and Management 103 (2015) 827–835.
[25] M.H. Khoshgoftar Manesh , P. Navida, A.M. Blanco Marigorta, M. Amidpour, M.H.
Hamedi, New procedure for optimal design and evaluation of cogeneration system based on
advanced exergoeconomic and exergoenvironmental analyses, Energy 59 (2013) 314-333.
[26] M. Fallah, S. Mohammad S. Mahmoudi, M. Yari, R. Akbarpour Ghiasi ,Advanced exergy
analysis of the Kalina cycle applied for low temperature enhanced geothermal system,
Energy 59 (2013) 314-333.
[27] Fateme Ahmadi Boyaghchi, Hanieh Molaie, Sensitivity analysis of exergy destruction in a
real combined cycle power plant based on advanced exergy method, Energy Conversion and
Management 99 (2015) 374–386.
[28] Yasin ¸ Eohret, Emin Açıkkalp , Arif Hepbasli , T. Hikmet Karakoc, Advanced exergy
analysis of an aircraft gas turbine engine: Splitting exergy destructions into parts, Energy xxx
(2015) 1-10.
[29] A. Lazzaretto, and G. Tsatsaronis, “On the Quest for Objective Equations in Exergy
Costing, Proceedings of the ASME Advanced Energy Systems Division”. ASME, New York,
Vol. 37, pp. 197-210, 1997.
[30] A. Nafey, M. Sharaf, L. García-Rodríguez, Thermo-economic analysis of a combined solar
organic Rankine cycle-RO desalination process with different energy recovery
configurations, Desalination 261 (2010) 138–147.
[31] Saleh B, Koglbauer G, Wendland M, Fischer J. Working fluids for low temperature organic
Rankine cycles. Energy 2007;32:1210e21.
[32] A. Ghaffarizadeh, Investigation on Evolutionary Algorithms Emphasizing Mass
ExtinctionB.Sc. Thesis Shiraz University of Technology, Shiraz, Iran, 2006.
[33] D. Goldberg, E. Genetic, Algorithms in Search, Optimization and Machine Learning,
Addison-Wesley, Reading, MA, 1989.
[34] J.D. Schaffer, Multiple objective optimizations with vector evaluated genetic
algorithmsPh.D. Thesis Department of Computer Science, Vanderbilt University, Nashville,
Tennessee, 1984.
[35] V. Srinivas, K. Deb, Multiobjective optimization using no dominated sorting in genetic
algorithms, J. Evol. Comput. 2 (1994) 221–248.
[36] E. Wang, H. Zhang, B. Fan, M. Ouyang, Y. Zhao, Q. Mu, Study of working fluid selection
of organic Rankine cycle (ORC) for engine waste heat recovery, Energy, 36 (2011) 3406-
3418.
 Optimum architecture of ORC cycles for the recovery of gas turbine exhaust selected
 working fluid selected in terms of environment, thermodynamics and performance range
 4E analysis of ORC and its optimization by genetic algorithm has been done.
 All components of the cycle at optimum state examined from advanced exergy and advanced
exergoeconomic perspectives.

You might also like