Вы находитесь на странице: 1из 14

Energy Conversion and Management 106 (2015) 859–872

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Thermoeconomic comparison between pure and mixture working fluids


of organic Rankine cycles (ORCs) for low temperature waste heat
recovery
Yongqiang Feng a, TzuChen Hung b, Kowalski Greg c, Yaning Zhang a, Bingxi Li a,⇑, Jinfu Yang d
a
School of Energy Science and Engineering, Harbin Institute of Technology, Harbin, China
b
Department of Mechanical Engineering, National Taipei University of Technology, Taipei, China
c
Mechanical and Industrial Engineering, Northeastern University, Boston, United States
d
Institute of Engineering Thermo-physics Chinese Academy of Sciences, Chinese Academy of Sciences, Beijing, China

a r t i c l e i n f o a b s t r a c t

Article history: Based on the thermoeconomic multi-objective optimization, simultaneously considering exergy effi-
Received 10 June 2015 ciency and levelized energy cost (LEC), the thermoeconomic comparisons between pure and mixture
Accepted 15 September 2015 working fluids of organic Rankine cycles (ORCs) have been investigated. Four models are proposed based
Available online 23 October 2015
on the different location of evaporating bubble point temperature or condensing dew point temperature
for mixture working fluids. The effects of mass fraction and four key parameters (evaporator temperature,
Keywords: condenser temperature, pinch point temperature difference and degree of superheat) on exergy efficiency
Organic Rankine cycle (ORC)
and levelized energy cost (LEC) are examined. Pareto-optimal solutions of four models using
Thermoeconomic multi-objective
optimization
0.7R245fa/0.3R227ea are obtained and compared. Taking mass fraction into account, the thermoeco-
Exergy efficiency nomic comparisons between pure and mixture working fluids have been studied. Research demonstrates
Levelized energy cost (LEC) that the mixtures don’t always present better thermodynamic performance and economic performance
Mixture working fluids than pure working fluids. Model 2 (T 7 ¼ T E ; T 3 ¼ T C ) is the favorable operation condition for its highest
thermodynamic performance and relatively low economic factor. Taking mass fraction as decision vari-
able, Pareto-optimal solutions for models 1, 2, 3 and 4 in pairs of (exergy efficiency (%), LEC ($/kW h)) are
(56.71, 0.188), (57.67, 0.192), (57.11, 0.194), and (56.91, 0.192), respectively. Compared with pure work-
ing fluids, the mixture working fluids present better exergy efficiency but worse LEC except model 1.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction [14] and gas turbine [15]. Therefore, it is of primary importance to


improve the cycle performance, select of working fluids and deter-
The recent rising electricity demand and cost, in addition to the mine the optimal operation parameters.
escalating global energy demands, are the main motivations in The selection of working fluids exhibits a significant effect on the
seeking for sustainable new technologies. Advanced technologies efficiency of system, the sizes of the system components, the design
for the conversion of low-grade heat sources into power, such as of expansion machine, and the economic performance. Various
the organic Rankine cycle (ORC), Kalina cycle [1,2], Goswami cycle types of fluids [5,16] such as hydrocarbons [17–20], hydroflurocar-
[3] and trilateral flash cycle [4], were proposed and tested. Because bon [21,22], ammonia [17,19], carbon dioxide [23,24], and mixtures
of the simple structure, high reliability, easy maintenance, as well [25,26] have been investigated. Hung et al. [17,18] analyzed ORC
as the better potential in converting the low-grade waste heat to efficiency using cryogens, e.g. benzene, ammonia, R11, R12, R134a
power than other cycles [5], ORC has been widely applied in recov- and R113, and did a comparative study between wet, dry and isen-
ering different heat sources [6,7], such as industrial waste heat [8], tropic fluids on ORC system. They found that isentropic fluids were
solar energy [9], geothermal energy [10], biomass energy [11], the most preferred working fluids for ORC system. Roy et al. [19]
ocean energy [12], internal combustion engine [13], micro turbine applied the parametric optimization on the non-regenerative ORC
using R12, R123, R134a and ammonia. They found that R123 was
⇑ Corresponding author at: School of Energy Science and Engineering, Harbin
the best working fluids for its higher efficiency and net power out-
Institute of Technology, 92, West Dazhi Street, Harbin 150001, China. Tel./fax: +86 put than other candidate working fluids. Saleh et al. [20] compared
451 86412078. the thermodynamic performances of 31 pure working fluids for
E-mail address: libx@hit.edu.cn (B. Li). ORC system and found that the working fluids with high boiling

http://dx.doi.org/10.1016/j.enconman.2015.09.042
0196-8904/Ó 2015 Elsevier Ltd. All rights reserved.
860 Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872

Nomenclature

A heat transfer area, m2 k thermal conductivity, W/m2 K


B0 boiling number, mqc    0:5 l dynamic viscosity of liquid, Ns/m2
0:8 qg
C0 convection number, 1x x q DT m logarithmic mean temperature difference
l

C BM bare module equipment cost


CEPCI Chemical Engineering Plant Cost Index Subscripts
CRF capital recovery cost 0 base state
COM cost of operation and maintenance 1–4, i state points
d diameter, m cr critical
E exergy, kW cs cold side
ED Euclidian distance d destruction
Fr Froude number eq equivalent
FM, FP multiplying factors ex exergy efficiency
F BM aggregate multiplication factors g generator
h specific enthalpy, kJ/kg glide glide temperature
i interest rate h hydraulic diameter
LCOE levelized cost of electricity hs hot side
LTpl plant lifetime in inlet
LEC levelized energy cost, $/kW h l liquid
m mass flow rate, kg/s max maximum
Nu Nusselt number min minimum
c l
Pr Prandtl number, pk net net
Q energy, kW; ratio of vapor moles to total moles out outlet
Re Reynolds number pp pinch point temperature difference
s specific entropy, kJ/kg K s isentropic
T temperature, K sup degree of superheat
t op operation time, h tot total
U overall heat transfer coefficient, W/m2 K wf working fluid
W power, kW
x average dryness fraction Acronyms
Y assessment parameter of TOPSIS decision making C condenser
E evaporator
Greek symbols Exp expander
a forced convection heat transfer coefficient, W/m2 K GWP global warming potential
b chevron angle of the plates H heat source
c latent heat, J/kg P pump
q density, kg/m3 PPTD pinch point temperature difference
g efficiency

temperature yielded better thermodynamic performance. More- [29] by comparing the thermodynamic performance between
over, some hydroflurocarbon were also examined and applied in R141b/RC318 and three pure working fluids. Pardeep et al. [30]
the high-temperature ORC system. The siloxanes was proposed on examined the realistic thermodynamic analysis for ORC system
the regenerative ORC for its better thermodynamic performance using isopentane, R245fa and their mixtures. They found that
and thermal stability [21] and cyclopentane was recommended as 0.7isopentane/0.3R245fa was the favorite working fluids and 13%
the best working fluid for its maximum thermal efficiency [22]. of thermal efficiency can be obtained at an optimal expansion ratio
Carbon dioxide [23,24] has been most commonly used as the of 7–10. Chys et al. [31] studied the effects of mixtures on ORC
working fluid for a supercritical Rankine cycle for its lower critical system. They found that the improvement in cycle efficiency for
temperature and good environmental characteristics. 423.15 K and 523.15 K of heat source were 16% and 6%, respec-
Additionally, some researchers have shown their interest in the tively. Yang et al. [32] examined the effect of zeotropic mixtures
mixture working fluids concerning thermodynamic efficiency for ORC system to recover a diesel engine exhaust energy. They
[25–38], mole composition [35], experimental setup [39,40] and found that the optimal evaporating pressure for maximum thermal
economic performance [41–47]. Shu et al. [25] conducted the ther- efficiency of ORC system using zeotropic mixtures was 2.0 MPa,
mal efficiency and exergy loss analysis of ORC system using three and R402B yielded the maximum net power output of 24.65 kW
pure hydrocarbons and two retardants. They found that zeotropic at the engine speed of 2200 r/min. Chen et al. [33] proposed a
mixtures presented better thermodynamic performance than the supercritical Rankine cycle using zeotropic mixtures for the low-
pure working fluids. The similar conclusions were obtained by grade heat. They found that the thermal efficiency of mixtures
Heberle et al. [26] using isobutane/isopentane and R227ea/ (0.7R134a/0.3R32) were 10–30% higher than that of the pure
R245fa to examine the second law efficiency of ORC system and R134a. Lecompte et al. [34] analyzed the second law efficiency of
Zhang et al. [27] using R245fa and 0.7isopentane/0.3R245fa to ORC system with zeotropic working fluids. They found that the
contrast the thermodynamic performance of regenerative ORC. mixture presented 7.1–14.2% higher second law efficiency than
Dong et al. [28] investigated the first law thermal efficiency of the pure working fluids. Zhao et al. [35] discussed the influence
high-temperature ORC system with zeotropic mixtures as working of composition shift on ORC system with zeotropic mixtures. They
fluids. They found that the mixtures widened the range of options found that the composition shift significantly influenced the
for working fluids. The similar results were obtained by Li et al. performance of ORC system. Chen et al. [36] studied the
Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872 861

thermodynamic first law and second law efficiency of ORC system temperature for the mixture working fluids, four models are intro-
using zeotropic working fluids. They found that the mixture’s over- duced. The effects of key parameters and mass fraction on the
all net power output was lower than that of pure working fluids at exergy efficiency and LEC are first illustrated in Section 4.1. Taking
some certain mole fractions. Li et al. [37] analyzed the thermo- those four key parameters (evaporator temperature, condenser
economic performance of ORC system with zeotropic mixtures. temperature, pinch point temperature difference and degree of
They found that the zeotropic mixtures have a better economic superheat) as decision variables, the Pareto-optimal solutions for
performance than the pure working fluids. Liu et al. [38] investi- four models using non-dominated sorting genetic algorithm-II
gated the effect of condensation temperature glide using mixture (NSGA-II) are conducted and compared in Section 4.2. The mass
on ORC performance. They found that when the increase of cooling fraction of R245fa is used as decision variable in the thermoeco-
water temperature is greater than the condensation temperature nomic optimization for mixture working fluids. Further investiga-
glide, an optimal working fluid mole fraction can be obtained to tion of the thermoeconomic comparisons between pure and
maximum thermodynamic performance. In addition, relevant mixture working fluids is discussed in Section 4.3.
studies [39,40] have experimentally assessed the effect of zeotro-
pic mixtures on system performance. 2. Modeling
Furthermore, great attention has been drawn to the single-
objective optimization for ORC system using mixtures with either 2.1. System description
thermodynamic performance [41–43] (net power output, exergy
efficiency) or economic factors [44–47] (LCOE). Andreasen et al. ORC is primarily made up of evaporator, expander, condenser
[41] applied a single-objective optimization of maximum net power and pump, as shown in Fig. 1. The pump supplies the working fluid
output on ORC system with two cases. They found that 12.9% higher (state 5) to the evaporator where the working fluid is heated and
net power output was obtained for optimized ethane/propane com- vaporized by the heat source. The high pressure vapor (state 1)
pared to ethane. The same objective function of maximum net flows into the expander and its enthalpy is converted into work.
power output was used by Liu et al. [42] via conducting the param- The low pressure vapor (state 2) exits the expander and is led to
eter optimization on geothermal ORC system using R600a/R601a the condenser where it is liquefied by water. The liquid is available
mixtures. They found that R600a/R601 mixtures outputted 11%, at the condenser outlet (state 4), and then it is pumped back to the
7% and 4% more power than R600a for geothermal water inlet tem- evaporator and a new cycle begins. The thermodynamic process for
peratures of 383 K, 403 K and 423 K, respectively. Xiao et al. [43] ORC system using pure working fluids is illustrated on a tempera-
applied multi-objective optimization for subcritical ORC system ture–entropy (T–s) diagram as shown in Fig. 2a. Compared with the
with a multi-objective function by the incorporation of four pure working fluids, the main advantage of mixtures as ORC
single-objective functions. They found that ORC performance of working fluids stems from their non-isothermal phase transi-
mixed working fluids was not always better than that of pure work- tions during vaporization and condensation, and hence effectively
ing fluids. The similar results were obtained by Li et al. [44] via match the temperature change of heat source and cooling water.
applying the thermodynamics and thermoeconomics analysis on The corresponding temperature–entropy diagram for ORC system
ORC systems using zeotropic mixtures. They also found that the using mixture working fluids is shown in Fig. 2b. The evaporator
operation parameters presented a significant impact on the is a heat exchanger characterized by a double-pipe construction
thermoeconomics performance of mixtures. Braimakis et al. [45] which is composed of an outer pipe and an outward convex
compared the thermodynamic performance of subcritical and corrugated inner tube, while the plate heat exchanger is used as
supercritical ORC system using five natural refrigerants and their the condenser in the present study [48,49].
binary mixtures from view of the exergetic efficiency and the
technical evaluation parameters. They found that the mixture 2.2. Thermodynamic and economic modeling
working fluids exhibit maximum exergy efficiency under supercrit-
ical cycle operation with the heat source temperatures of 443 K. The following general assumptions [50] are used in this study:
Mavrou et al. [46] applied a multi-criteria mixture selection
methodology on solar ORC system and found that the favorable (a) The system is under steady state.
mixture was neopentane — 2-fluoromethoxy-2-methylpropane at (b) Heat and friction losses as well as the change of potential
70% neopentane for its higher net power output and thermal and kinetic energies are neglected.
efficiency than other candidate mixtures. Le et al. [47] conducted
the thermodynamic and economic optimization on the subcritical
ORC to maximize exergy efficiency and minimize LCOE (Levelized
Cost of Electricity) using two pure working fluids and their
11 8
mixtures. They found that better thermodynamic and economic
performance was obtained for n-pentane.
From the above literature, it is evident that the single-objective
optimization [44–47] for ORC system using mixture working fluids
from view on either thermodynamic performance or economic
factor have been investigated. However, few of them tried to fulfill
the thermoeconomic optimization using mixture working fluids
considering thermodynamic performance and economic factor
simultaneously. Furthermore, limited works attempted to take into
account the comparison between mixture and pure working fluids
based on thermoeconomic optimization.
Based on these statements, the present work focuses on the
thermoeconomic comparisons between pure and mixture working
12 14
fluids based on thermoeconomic optimization considering exergy
efficiency and LEC simultaneously. Based on the different location
of evaporating bubble point temperature or condensing dew point Fig. 1. Schematic diagram of ORC system for low temperature waste heat.
862 Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872

330 E_ ¼ m½ðh
_ i  h0 Þ  T 0 ðsi  s0 Þ ð7Þ

300
Evaporator model heating evaporating superheating The equation of the exergy destruction rate (E_ d ) can be
expressed as
9 8 X X
270 11 10
1 E_ d ¼ E_ in  E_ out  W
_ ð8Þ
Temperature (K)

Heat source Tpp,E Tsup


The exergy destruction rate of ORC system (E_ d ) can be
6 7
240
2 expressed:
5
210 5s
3
2s E_ d ¼ E_ d;E þ E_ d;EXP þ E_ d;C þ E_ d;P ð9Þ
4
180 13 Tpp,C 14
The exergy efficiency of ORC can be expressed as:
Cooling water 12
_
W
150 condensation cooling
g_ ex ¼ _ net_ ð10Þ
Condenser model Ed þ W net
120 The pinch point temperature difference in the evaporator
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
(DT pp;E ) and condenser (DT pp;C ) can be expressed as:
s (kJ/kgK)
DT pp;E ¼ T 10  T 6 ð11Þ
Fig. 2a. Typical T–s diagram for ORC system using pure working fluids, as well as
DT pp;C ¼ T 3  T 13 ð12Þ
evaporator and condenser model.
The degree of superheat (DT sup ) is the temperature difference
between the evaporator outlet temperature (T 1 ) and saturation
330
vapor temperature (T 7 ) corresponding to evaporator outlet pres-
300
Evaporator model heating evaporating superheating sure, which is given by:

9 8 DT sup ¼ T 1  T 7 ð13Þ
270 11 10
1
Temperature (K)

Heat source Tpp,E Tsup Because of the non-isothermal phase transitions of mixture
240 6 7
Tglide,E working fluids, the temperature glide of evaporator (DT glide;E ) and
2 condenser (DT glide;C ) can be defined as:
5 2s
210 5s 3 DT glide;E ¼ T 7  T 6 ð14Þ
4 Tpp,C Tglide,C
14
DT glide;C ¼ T 3  T 4 ð15Þ
180 Cooling water 12 13
Based on the variation of heat transfer coefficient caused by dif-
150 Condenser model condensation cooling ferent phase state, the evaporator is divided into three sections
(heating section (5–6), evaporating section (6–7) and superheating
120 section (7–1)), while the condenser is divided into two sections
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
(cooling section (2–3) and condensation section (3–4)). The
s (kJ/kgK)
detailed sections for the pure and mixture working fluids are
Fig. 2b. Typical T–s diagram for ORC system using mixture working fluids, as well shown in Figs. 2a and 2b, respectively. The heat transfer area (Ai )
as evaporator and condenser model. for each section can be expressed as:
Q i lnðDT max =DT min Þ
Ai ¼ ð16Þ
U i ðDT max  DT min Þ
(c) Saturated liquid is supposed at the condenser exit. The heat transfer coefficient (U i ) for each section can be
(d) There are no pressure drops in the heat exchangers, con- expressed as:
densers and pipes. 1 1 di 1
¼ þ þ ð17Þ
(e) The isentropic efficiencies of the expander and the pump are U i ai;hs ki ai;cs
both set to be 0.8.
The heat-transfer coefficient of heating process (5–6) and
superheating process (7–1) of the evaporator can be calculated
Based on the assumptions and combined with the T–s plot (as
using the Dittus–Boelter correlation [51].
_ P ), heat transfer rate (Q_ E ),
shown in Fig. 2), the pump power (W
k
expander power (W t ) and condenser heat rate (Q_ C ) can be
_ ai;cs ¼ 0:023 Re0:8 Prn ð18Þ
d
expressed as follows:
The Chisholm and Wanniarachchi correlation [52] is employed for
_ P¼m the cooling process (2–3) of condenser, which is calculated as follows:
W _ wf ðh5  h4 ÞgP ð1Þ  0:646
h  h4 6b
gs;P ¼ 5s ð2Þ Nu ¼ 0:724 Re0:583 Pr1=3 ð19Þ
h5  h4 p
Q_ E ¼ m
_ wf ðh1  h5 Þ ð3Þ For the pure working fluids, the Kandlikar correlation [53] is
_ t¼m used for the evaporating process (6–7), which can be expressed as:
W _ wf ðh1  h2s Þgs;t gg ¼ m
_ wf ðh1  h2 Þgg ð4Þ ai;cs
¼ C 1 ðC 0 ÞC 2 ð25Frl ÞC5 þ C 3 ðB0 ÞC4 F fl ð20Þ
Q_ C ¼ m
_ wf ðh2  h4 Þ ð5Þ al
 0:8
_ net ) can be written as: mð1  xÞd Pr0:4
l kl
And therefore, the net power output (W al ¼ 0:023 ð21Þ
ll d
_ net ¼ W
_ tW
_P  0:8  0:5
W ð6Þ 1  x qg
C0 ¼ ð22Þ
The exergy of the state point can be considered as:
x ql
Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872 863

q
B0 ¼ ð23Þ 0.110 0.15
mc
m2 0.105 0.14
Fr l ¼ ð24Þ
9:8q2l d

Thermal efficiency
0.100 0.13
where al is the forced convection heat transfer coefficient for liquid

LEC ($/kWh)
single-phase; C 0 and B0 are the convection number and the boiling
0.095 0.12
number, respectively; Frl is the Froude number of liquid; F fl is the
fluid-dependent parameter; x is the average dryness fraction; c
0.090 0.11
denotes latent heat; and q is density. The values of C 1 , C 2 , C 3 , and
Thermal efficiency of Ref. [44]
C 4 are set to 1.136, 0.9, 667.2, and 0.7, respectively. The value of Thermal efficiency of present simulation
0.085 0.10
C 5 is 0.3 in convective region and nucleate boiling region. However, LEC of Ref. [44]
the value of C5 is 0 for vertical tube and for horizontal tubes with LEC of present simulation

Frl > 0:04. 0.080 0.09


350 355 360 365 370 375
The heat transfer coefficient [54] of the condensation process
Evaporator temperature (K)
(3–4) for pure working fluids is calculated by:
Fig. 3. Validation of the numerical model with the previously published data [44]
hdh 1=3
Nu ¼ ¼ 4:118Re0:4
eq Pr l ð25Þ for ORC system using R245fa/R600a (0.7/0.3).
k
_ h
md
Reeq ¼ ð26Þ
ll 2.3. Accuracy and validation
As stated by Li et al. [44], the two-phase heat transfer coeffi-
In order to validate the computation of the present model, the
cients of mixture working fluids are much higher than those of
energy balance equations are solved with the same operating condi-
gas side in the heat exchanger (evaporator and condenser), and
tions as in Ref. [44] using R245fa/R600a (0.7/0.3) as the working
the overall heat transfer coefficient is mainly dependent on the
fluid. The comparison between the simulation result and the result
convective heat transfer coefficient of gas side. Therefore, the over-
reported by Li et al. [44] is shown in Fig. 3. It can be observed that the
all heat transfer coefficient is estimated to be approximately equal
simulation result of thermal efficiency agrees well with that in the
to the heat transfer coefficients of gas side in the evaporator and
reference, whose maximum derivation from the date in reference
condenser. In the present study, similar method is applied on the
is 0.9%. For LEC, the maximum derivation between present study
overall heat transfer coefficients of the evaporating process (6–7)
and Ref. [44] is 5.6%, which attributed to the different calculation
and condensation process (3–4).
for the overall heat transfer coefficients of the heat exchangers.
The total heat transfer area in the evaporator (AE;tot ) and
condenser (AC;tot ) can be obtained as:
3. Multi-objective optimization and decision making
AE;tot ¼ A56 þ A67 þ A71 ð27Þ
AC;tot ¼ A23 þ A34 ð28Þ 3.1. Multi-objective optimization
The levelized energy cost (LEC) [48,49,55] is defined as the
system cost to total net power output, which is selected as the 3.1.1. Multi-objective optimization solution
evaluation criterion for economic factor. A multi-objective optimization has been widely applied in var-
ious disciplines to minimize or maximize simultaneously two or
2
lg C p;i ¼ K 1 þ K 2 lg Ai þ K 3 ðlg Ai Þ ð29Þ more functions. The multi-objective optimization can be expressed
lg F p;i ¼ C 1 þ C 2 lg Pi þ C 3 ðlg Pi Þ
2
ð30Þ using mathematical terms [59]:
C BM;i ¼ C P;i F BM;i ¼ C P;i ðB1 þ B2 F M F P;i Þ ð31Þ T
min FðXÞ ¼ ½f 1 ðXÞ; f 2 ðXÞ; f 3 ðXÞ; . . . ; f n ðXÞ ð36Þ
C BM;i;2008 ¼ C BM;i;1996 CEPCI2011 =CEPCI1996 ð32Þ
X
C tot ¼ C BM;i ð33Þ subject to

where K 1 , K 2 , K 3 , C 1 , C 2 , and C 3 are coefficients for the cost evalua- g i ðXÞ 6 0; i ¼ 1; . . . ; m


tion of system component with the respective value of 3.2138, hi ðXÞ 6 0; j ¼ 1; . . . ; n
0.2688, 0.07961, 0.06499, 0.05025, and 0.01474; B1 and B2 are xk;min 6 xk 6 xk;max
dependent on the type of heat exchanger with the respective value
of 1.8 and 1.5; F M account for the effects of construction material on where X denotes the vector of decision variables to be optimized; F
costs with the constant value of 1.25 [49]. (X) is the vector of objective function; g i ðXÞ is the inequality con-
The capital recovery cost (CRF) is estimated based on the follow- straints while hi ðXÞ is the equality constraints, respectively; xk;min
ing relation [56,57]: and xk;max are the bottom and top bounds of the decision variables,
LTpl LTpl respectively.
CRF ¼ ið1 þ iÞ =½ið1 þ iÞ  1 ð34Þ
NSGA-II (Non-dominated Sorting Genetic Algorithm II), an
where i is the interest rate and set at 5%, and LTpl is the plant life- improved version of NSGA, is applied to conduct the multi-
time and set as 20 years. objective optimization of ORC system in this study. The genetic
Therefore, LEC is simplified by ignoring the field capital cost and algorithm repeatedly modifies a population of individual solutions.
the field O&M cost. It can be expressed as [44,49,55]: At each step, the genetic algorithm selects individuals at random
from the current population to be parents and uses them to pro-
LEC ¼ ðCRF  C tot þ COMÞ=ðt op  W net Þ ð35Þ
duce the children for the next generation. Over successive genera-
where COM is the cost of operation and maintenance which is 1.5% tions, the population evolves toward an optimal solution [60].
of the investment cost, and the t op is the operation time and set as Details about the procedure adopted by NSGA-II can be found in
8000 h [58]. Refs. [61,62].
864 Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872

rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Table 1 Xn  ideal
2
Decision variables and their lower and upper boundaries. EDiþ ¼ j¼1
f ij  f j ð38Þ
Decision variables Lower bound Upper bound
ideal
where f j is the ideal solution of jth objective in a single-objective
Evaporator temperature (K) 330 370
Condenser temperature (K) 303 312 optimization.
Degree of superheat (K) 3 10 The Euclidian distance of each solution from the nadir solution
Pinch point temperature difference (K) 3 10 [67,68] denoted EDi is defined as:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Xn  nadir
2
EDi ¼ j¼1
f ij  f j ð39Þ
3.1.2. Objective functions and decision variables
Two objective functions are considered for the multi-objective A new assessment parameter is defined as follows:
optimization: exergy efficiency and LEC. The following four key
EDi
system parameters are selected as the decision variables: evapora- Yi ¼ ð40Þ
EDi þ EDiþ
tor temperature, condenser temperature, degree of superheat, and
pinch point temperature difference. The lower and upper bound of In TOPSIS decision making, a solution with maximum Yi is
these constraints are listed in Table 1. selected as a desired final solution. Therefore, ifinal is
ifinal ¼ i 2 maxðY i Þ ð41Þ
3.1.3. Constraints
Based on the energy balance and the definition of pinch point
temperature difference, the following constraints should be 4. Results and discussion
satisfied.
As stated by Li et al. [44], the comparative results between the
T 1 þ DT pp;E < T 8 ; T 7 þ DT pp;E < T 9 ; T 5 þ DT pp;E
pure and mixture working fluids are based on the location of evap-
< T 11 ; T 14 þ DT pp;C < T 2 ; T 12 þ DT pp;C < T 4 ð37Þ orating bubble point temperature or condensing dew point tem-
perature for the mixture working fluids. In Fig. 2b, points 6 and 7
are the potential location of evaporating bubble point temperature,
3.2. Decision-making in multi-objective optimization
while the condensing dew point temperature can be located at
points 3 and 4. Based on the location of evaporating bubble point
In multi-objective optimization problems, the selection proce-
temperature and condensing dew point temperature, four models
dure is more complicated than in single-objective optimization
for mixture working fluids are introduced. It should be noted that
problems. Each Pareto-optimal solution represents a compromise
the constrains (Eq. (37)), as discussed in Section 3.3, ensure the
considering different objectives, such that the component of the
location of pinch point temperature difference. Combined with
corresponding vector of objectives cannot simultaneously be
the T–s plot of mixture working fluids (as shown in Fig. 2b), the
improved. Any improvement in an objective requires at least one
detailed four models (as shown in Fig. 5) are described as follows:
of the other objectives to be scarified, and therefore a set of
trade-offs exists [63]. TOPSIS (Technique for Order Preference by
Model 1: T 7 ¼ T E ; T 4 ¼ T C
Similarity to Ideal Situation), is a simple ranking method in con-
ception and application. The standard TOPSIS method attempts T 6 ¼ refpropmð‘T’; ‘P’; PE ; ‘Q ’; 0; FLUID1; FLUID2; ½x1 x2Þ
to choose alternatives that simultaneously have the shortest dis- ð42Þ
tance from the positive ideal solution and the farthest distance
T 3 ¼ refpropmð‘T’; ‘P’; PC ; ‘Q ’; 1; FLUID1; FLUID2; ½x1 x2Þ
from the negative-ideal solution. The TOPSIS decision making is
used in the present study. The detailed parameters for decision
ð43Þ
making are shown in Fig. 4. Model 2: T 7 ¼ T E ; T 3 ¼ T C
The Euclidian distance of every solution on the Pareto frontier
from the ideal solution marked by EDiþ is defined as [64–66]: T 6 ¼ refpropmð‘T’; ‘P’; PE ; ‘Q ’; 0; FLUID1; FLUID2; ½x1 x2Þ
ð44Þ
T 4 ¼ refpropmð‘T’; ‘P’; PC ; ‘Q ’; 0; FLUID1; FLUID2; ½x1 x2Þ
ð45Þ
fj
Model 3: T 6 ¼ T E ; T 4 ¼ T C

T 7 ¼ refpropmð‘T’; ‘P’; PE ; ‘Q ’; 1; FLUID1; FLUID2; ½x1 x2Þ


Ideal
fj
Ideal Ideal ð46Þ
Ideal solution ( fi , fj )
T 3 ¼ refpropmð‘T’; ‘P’; PC ; ‘Q ’; 1; FLUID1; FLUID2; ½x1 x2Þ
fij ð47Þ
Model 4: T 6 ¼ T E ; T 3 ¼ T C

T 7 ¼ refpropmð‘T’; ‘P’; PE ; ‘Q ’; 1; FLUID1; FLUID2; ½x1 x2Þ


Pareto-optimal frontier
ð48Þ
Nadir solution ( fi
Nadir
, fj
Nadir
) fi
Ideal
T 4 ¼ refpropmð‘T’; ‘P’; PC ; ‘Q ’; 0; FLUID1; FLUID2; ½x1 x2Þ
ð49Þ

fi The system parameters (evaporator temperature, condenser


temperature, pinch point temperature difference and degree of
Fig. 4. Pareto frontier of a multi-objective optimization. superheat) and mass fraction of mixture have significant influence
Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872 865

330 330

Model 1: T7=TE, T4=TC Model 2: T7=TE, T3=TC


300 300

270 1 270
Temperature (K)

Temperature (K)
6 7 6 7 1

240 240
5 2 2
5
210 5s 210 5s
2s 3 2s
4 3 4
180 180

150 150

120 120
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
s (kJ/kgK) s (kJ/kgK)

330 330

Model 3: T6=TE, T4=TC Model 4: T6=TE, T3=TC


300 300

270 270
Temperature (K)

Temperature (K)
6 6
7 1 7 1
240 240
5 2 5 2
210 5s 210 5s 2s
2s
4 3 4 3
180 180

150 150

120 120
0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
s (kJ/kgK) s (kJ/kgK)

Fig. 5. The T–s plot of four models for mixture working fluids.

on the exergy efficiency and LEC, which are carried out firstly in Table 2
Section 4.1. Taking those four key parameters (evaporator temper- Main assumptions for ORC system.
ature, condenser temperature, pinch point temperature difference
Item Unit Value
and degree of superheat) as decision variables, the Pareto-optimal
Heat sources temperature K 393
solutions for four models are conducted and compared in Sec-
Expander isentropic efficiency (%) 80
tion 4.2. The mass fraction of R245fa is used as decision variable Generator efficiency (%) 90
in the thermoeconomic optimization for mixture working fluids, Pump isentropic efficiency (%) 80
and the thermoeconomic comparisons between pure and mixture Pump efficiency (%) 90
working fluids is discussed in Section 4.3. ORC simulation is per- Mass flow of heat sources kg/s 0.33
Cooling water temperature K 293
formed in combination of Matlab and Refprop [69]. The main Degree of superheat (DT sup ) K 5
assumptions for ORC system are listed in Table 2. It should be PPDT in condenser (DT pp;C ) K 5
noted that two pure working fluids (R245fa and R227ea) and their PPDT in evaporator (DT pp;E ) K 10
mixtures [26] are selected as the working fluids in the present Condenser temperature K 303
study. The main properties of two pure working fluids are listed Environmental temperature K 293

in Table 3. Fig. 6 shows the T–s diagrams of R245fa, R227ea and


their mixtures.
Table 3
4.1. Parametric analysis on system performance Main properties of the working fluids [26].

Working fluids M (kg/kmol) Tcr (K) Pcr (MPa) GWP 100 years
4.1.1. Effect of mass fraction on system performance
R245fa 134.05 427.16 3.65 1030
In this part, the evaporator temperature and condenser temper- R227ea 170.03 374.90 2.93 3220
ature are 363 K and 303 K, respectively. The variations of exergy
efficiencies and LEC for four models with mass fraction of R245fa
are displayed in Figs. 7 and 8.
Fig. 7 shows the variations of exergy efficiencies for four models with mass fraction of R245fa. For the same mass fraction of R245fa,
with mass fraction of R245fa. As observed, as the mass fraction of the highest exergy efficiency is obtained for model 3, followed by
R245fa increases, the exergy efficiencies of models 1 and 3 initially model 1 and model 4, model 2 yields the lowest exergy efficiency.
increase and then decrease, while that of model 2 presents a When the mass fraction of R245fa is 0.5, the exergy efficiencies are
reverse trend. The exergy efficiency of model 4 increases smoothly 52.65% for model 1, 45.69% for model 2, 56.30% for model 3, and
866 Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872

450 R245fa (51.17% of exergy efficiency), while that of models 2 and


Mass fraction of R245fa 4 yield lower exergy efficiency compared to the pure working fluid
0.9 of R227ea (48.29% of exergy efficiency). The variations of LEC for
425 0.8
0.7 four models with mass fraction of R245fa are demonstrated in
0.6 Fig. 8. The LEC of four models have a same behavior, increasing first
Temperature (K)

400 0.5 and then decreasing. When the mass fraction of R245fa is 0.5, the
0.4 R245fa LEC for models 1, 2, 3 and 4 are 0.261, 0.265, 0.248 and 0.251 in
0.3
375
0.2
unit of $/kW h, respectively. It is also found that for most case of
0.1 models 1, 2, 3 and 4, the mixture working fluids own worse eco-
350
nomic performance than the pure working fluids.
R227ea As observed in Figs. 7 and 8, the mixture don’t always present
better thermodynamic performance and economic performance
325 than the pure working fluids. Whether the mixtures exhibit better
thermodynamic and economic performance than the pure working
300 fluids depend on the operation parameters and mass fraction of
1.0 1.2 1.4 1.6 1.8 2.0
mixtures.
s (kJ/kgK)

Fig. 6. The T–s diagrams of R245fa, R227ea and their mixtures.


4.1.2. Parameters on system performance
The variations of exergy efficiency for four models with
evaporator temperature and condenser temperature using
0.7R245fa/0.3R227ea are displayed in Fig. 9. The pinch point
0.58
Model 1 temperature difference and degree of superheat are 10 K and 5 K,
Model 2 respectively. As observed, higher evaporator temperature or lower
0.56
Model 3 condenser temperature ensure better exergy efficiency. As noted
Model 4 in Eq. (10), the exergy efficiency is determined by the net power out-
0.54
put and exergy destruction. The increase in evaporator temperature
Exergy efficiency

ensures more specific enthalpy difference in expander, and then the


0.52
mass flow reduces. The effect of increasing specific enthalpy differ-
ence is obvious than the decreasing mass flow, resulting in an
0.50
increase in net power output at first. At the same time, the exergy
destruction goes down with evaporator temperature. The compre-
0.48
hensive effect of the increasing net power output and decreasing
exergy destruction makes sure an increase in exergy efficiency. With
0.46
the further increase in evaporator temperature, the net power
output decreases later which is contributed to the predominance
0.44
0.0 0.2 0.4 0.6 0.8 1.0 of the decreasing mass flow. The effect of the decreasing exergy
Mass fraction of R245fa destruction is greater than that of net power output, resulting in
the increase in exergy efficiency. At the same time, increasing
Fig. 7. Variations of exergy efficiencies for four models with mass fraction of R245fa condenser temperature causes less specific enthalpy difference in
(TE = 363 K, TC = 303 K). expander and less mass flow, which leads to the decrease in net
power output. Less exergy destruction is yielded with condenser
temperature. The decreasing net power output outperforms that of
0.28 exergy destruction, ensuring the decrease in exergy efficiency. It
Model 1
Model 2
Model 3
0.26 Model 4 0.60 Model 3
0.55
LEC ($/kWh)

0.50 Model 1
ncy

0.45
Exergy efficie

0.24
0.40
0.35
0.30
0.22
0.25
0.20 Model 4

0.15
0.20 0.100
0.0 0.2 0.4 0.6 0.8 1.0 Co 30 302 370
n 304 Model 2 360
Mass fraction of R245fa de
ns 306 350 (K
)
er 308 tur
e
Fig. 8. Variations of LEC for four models with mass fraction of R245fa (TE = 363 K,
tem 310 340 e ra
pe p
TC = 303 K). ar 312 33 r te
m
tu
re 314 0 or ato
ap
(K
) Ev
49.68% for model 4. At the same time, it can be found that for cer-
tain mass fraction, the mixture working fluids of models 1 and 3 Fig. 9. Variations of exergy efficiency for four models with evaporator temperature
present better exergy efficiency than pure working fluid of and condenser temperature using 0.7R245fa/0.3R227ea (DT pp;E ¼ 10 K, DT sup ¼ 5 K).
Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872 867

can also be found that model 3 presents the highest exergy


efficiency; followed by model 1 and model 4; model 2 yields the 0.60
lowest exergy efficiency. For example, when the evaporator temper- 0.56
Model 2
ature and condenser temperature are 350 K and 308 K, respectively, 0.52
0.48

LEC ($/kWh)
the exergy efficiency is 44.24% for model 1, 33.69% for model 2,
Model 4
49.68% for model 3, and 39.12% for model 4. 0.44
Model 1
Fig. 10 depicts the variations of exergy efficiency for four 0.40
models with pinch point temperature difference and degree of 0.3 6
superheat using 0.7R245fa/0.3R227ea with evaporator temperature 0.32
of 363 K and condenser temperature of 303 K. The exergy efficiencies 0.28
of four models have a same behavior of decreasing trend with the 0.24
pinch point temperature difference and degree of superheat. 0.20
0.16 4
When the pinch point temperature difference increases, the Co 30 6 370
mass flow decreases, thus the net power output decreases. nd 30 360
en 308 )
Meanwhile, the exergy destruction decreases with the pinch point se 350 (K
r Model 3 re
tem 310 atu
temperature difference. The net power output decreases faster pe
340
m pe r
ar 312 33 te
than that of exergy destruction, resulting in the decrease in exergy or
tu
re
0
p orat
efficiency. At the same time, an increase in degree of superheat a
(K
) Ev
enables more specific enthalpy difference in expander and less
mass flow. While the effect of decreasing mass flow is greater than
Fig. 11. Variations of LEC for four models with evaporator temperature and
increasing specific enthalpy difference in expander, the net power condenser temperature using 0.7R245fa/0.3R227ea (DT pp;E ¼ 10 K, DT sup ¼ 5 K).
output decreases. Increasing the degree of superheat ensures less
exergy destruction. The exergy destruction decreases faster than
that of net power output, resulting in an increase in exergy thus LEC increases afterward. In addition, the increase of condenser
efficiency. When the pinch point temperature difference and temperature ensures the mean heat transfer temperature differ-
degree of superheat is 15 K and 10 K, respectively, the exergy ence in condensing section of condenser to increase, resulting in
efficiencies for models 1, 2, 3 and 4 are 49.61%, 40.53%, 43.83%, the decrease in the heat transfer area of condenser, which leads
and 43.76%, respectively. to a reduction of system investment cost. Nevertheless, the effect
The variations of LEC for four models with evaporator tempera- of the decreasing net power output would outweigh that of the
ture and condenser temperature using 0.7R245fa/0.3R227ea are decreasing system investment cost. As a result, LEC exhibits an
displayed in Fig. 11. For four models, LEC decreases first and then increasing trend with the condenser temperature. Model 2 yields
increases with evaporator temperature, while LEC keeps rising the highest LEC, followed by models 4 and 1, and the lowest LEC
with condense temperature. With the increase of evaporator is obtained by model 3. For example, when the evaporator temper-
temperature, the mean heat transfer temperature difference in ature and condenser temperature are 350 K and 308 K, respec-
superheating section of evaporator increases, resulting in a tively, LEC is 0.216 for model 1, 0.252 for model 2, 0.199 for
decrease in the heat transfer area, which leads to a reduction of model 3, and 0.227 for model 4 in unit of $/kW h.
the system investment cost. The comprehensive effect of the The variations of LEC for four models with pinch point
decreasing system investment cost and the increasing net power temperature difference and degree of superheat using
output results in the decline in LEC with the evaporator tempera- 0.7R245fa/0.3R227ea are demonstrated in Fig. 12. LEC of four mod-
ture at first. With the further increase of evaporator temperature, els present the same behavior of increasing with the pinch point
the decreasing net power output becomes gradually obvious, and temperature difference and degree of superheat. When the pinch
point temperature difference increases, the mean heat transfer
temperature difference in evaporating section and heating section

0.52
Model 1
0.5 Model 2
0.50
ncy

0.48 Model 3
Exergy efficie

LEC ($/kWh)

Model 1
0.46 0.4
Model 4

0.44

0.42 0.3

0.40
Model 2 14 Model 3
0.38 12 0.2
Pi n 12 K) 6
ch 14 t( Model 4
De

poi 10 a 8
nt e
rh
gr

tem 16 8 20
pe 10
e

per 18
eo

atu 18
f su 16 K )
6 12 nce (
fs

re
dif 20 e eo 14 fere
up

fe gr 14 tur e dif
ren De
er

12
ce pera
he

(K) t tem
at

oi n
hp
(K

Pinc
)

Fig. 10. Variations of exergy efficiency for four models with degree of superheat
and pinch point temperature difference using 0.7R245fa/0.3R227ea (TE = 363 K, Fig. 12. Variations of LEC for four models with degree of superheat and pinch point
TC = 303 K). temperature difference using 0.7R245fa/0.3R227ea (TE = 363 K, TC = 303 K).
868 Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872

of evaporator to increase, resulting in the decrease in the heat for maximum exergy efficiency and LEC are in range of 0.4–0.8.
transfer area of evaporator, which leads to the system investment Therefore, 0.7R245fa/0.3R227ea is selected to give a thermoeco-
cost decreases. Still, the decreasing net power output is greater nomic comparison of four models. The genetic algorithm parame-
than that of system investment cost. As a result, LEC increases with ters are specified according to the following values: population size
the pinch point temperature difference. At the same time, the 40, generation size 100, crossover fraction 0.8 and migration frac-
increasing degree of superheat ensures the mean heat transfer tion 0.2.
temperature difference in superheating section of evaporator to The Pareto frontiers for four models using 0.7R245fa/0.3R227ea
decrease, leading to a decline in the total heat transfer area, which are displayed in Fig. 13, respectively. LEC of four models have the
causes a decreasing in system investment cost. The decreasing net same behavior of increasing trend with exergy efficiency. There
power output exhibits more obvious than that of system invest- is a clear trade-off between thermodynamic performance and eco-
ment cost, making sure that LEC increases with the degree of nomic performance, as an improvement in one criterion can only
superheat. When the pinch point temperature difference and be achieved at the expense of worsening the other. In case of model
degree of superheat is 15 K and 10 K, respectively, LEC for models 1 (as shown in Fig. 13), the maximum exergy efficiency exists at
1, 2, 3 and 4 are 0.326, 0.324, 0.320, and 0.305 in unit of $/kW h, design point C (55.52%), while LEC is the highest value at this point
respectively. (0.225 $/kW h). At the same time, the minimum LEC occurs at
design point A (0.188 $/kW h) with the lowest exergy efficiency
(49.42%). The design point C owns the optimal solution with exergy
4.2. Thermoeconomic comparison of four models using 0.7R245fa/ efficiency as single-objective function, while design point A pre-
0.3R227ea sents the optimal solution with LEC as single-objective function.
To provide a good relation between exergy efficiency and LEC,
In order to ascertain which models present better thermoeco- the curves are fitted to the optimized points and the expressions
nomic performance, the Pareto-optimal solutions for four models of four models using 0.7R245fa/0.3R227ea are given as follows:
are obtained and compared. The four system parameters, as
discussed in Section 4.1.2, are selected as the decision variables LEC ¼ 470:1g3ex  723:9g2ex þ 371:7gex  63:4
with their respective ranges in Table 1. Note that the mass fraction
R2 ¼ 0:989 ð49:42% 6 gex 6 55:22% for model 1Þ ð50Þ
of R245 should be constant for the thermoeconomic comparison of
four models. As shown in Figs. 7 and 8, the mass fractions of R245
LEC ¼ 494:9g3ex  761:5g2ex þ 390:6gex  66:6
R2 ¼ 0:984 ð49:40% 6 gex 6 55:56% for model 2Þ ð51Þ
0.245
C'''
0.240 Working fluid: 0.7R245fa/0.3R227ea C'
LEC ¼ 408:9g3ex  621:7g2ex þ 315:2gex  53:1
0.235 Model 1
0.230 Model 2 R ¼ 0:988 ð48:86% 6 gex 6 54:91% for model 3Þ ð52Þ
Model 3 C''
0.225 Model 4 C
LEC ($/kWh)

0.220 LEC ¼ 457:6g3ex  697:4g2ex þ 354:4gex  59:9


0.215 R2 ¼ 0:994 ð49:39% 6 gex 6 55:45% for model 4Þ ð53Þ
0.210
B' B'' B Pareto-optimal solutions obtained from TOPSIS decision-
0.205 making for four models are marked with B in Fig. 13. Table 4 lists
0.200 the detailed optimal values for four models of objective and
B'''
0.195 decision variables with points A–C on the Pareto frontier.
A'' A' A A'''
With respect to the single-objective of maximum exergy effi-
0.190
0.185 ciency, model 2 presents the highest exergy efficiency of 55.56%,
0.48 0.49 0.50 0.51 0.52 0.53 0.54 0.55 0.56 followed by models 4 and 1, and model 3 yields the lowest exergy
Exergy efficiency efficiency of 54.91%. As to the single-objective of minimum LEC, the
lowest LEC of 0.187 $/kW h is obtained for models 2 and 4, while
Fig. 13. Pareto frontiers of LEC with exergy efficiency for models 1, 2, 3 and 4 using models 1 and 3 own the highest LEC of 0.188 $/kW h. However,
0.7R245fa/0.3R227ea.
the difference between the lowest and the highest LEC is so small

Table 4
Optimized values of objective and decision variables for points A–C on the Pareto frontier using 0.7R245fa/0.3R227ea for four models.

Models Optimal solution Decision makings Design variables Objectives


T E (K) T C (K) DT sup (K) DT pp;E (K) g_ ex (%) LEC ($/kW h)

Model 1 NAGA-II TOPSIS (point B) 364.70 305.17 4.46 3.00 54.08 0.204
Max g_ ex Point C 370.00 304.90 4.06 3.00 55.22 0.225
Min LEC Point A 351.15 304.60 7.44 3.08 49.42 0.188
Model 2 NAGA-II TOPSIS (point B0 ) 362.90 304.25 5.15 3.00 54.02 0.201
Max g_ ex Point C0 370.00 303.00 5.02 3.71 55.56 0.241
Min LEC Point A0 351.50 305.12 5.43 3.00 49.40 0.187
Model 3 NAGA-II TOPSIS (point B00 ) 364.32 305.29 4.01 3.35 53.79 0.204
Max g_ ex Point C00 370.00 305.29 4.36 3.33 54.91 0.227
Min LEC Point A00 350.63 304.95 3.00 3.80 48.86 0.188
Model 4 NAGA-II TOPSIS (point B000 ) 364.66 306.24 3.00 3.00 53.65 0.203
Max g_ ex Point C000 369.99 303.01 7.53 3.74 55.45 0.243
Min LEC Point A000 351.39 305.07 5.13 3.01 49.39 0.187
Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872 869

(0.001 $/kW h), indicating that the location of evaporator temper- temperature, 305.17 K, 304.25 K, 305.29 K and 306.24 K of con-
ature and condenser temperature presents slight effect on the denser temperature, 4.46 K, 5.15 K, 4.01 K and 3.00 K of pinch point
single-objective of minimum LEC. From the viewpoint of single- temperature difference, and 3.00 K, 3.00 K, 3.35 K and 3.00 K of
objective optimization of maximum exergy efficiency and degree of superheat, respectively.
minimum LEC, model 2 is the favorable operation condition for
its highest thermodynamic performance and relatively low
4.3. Thermoeconomic comparison between pure and mixture working
economic factor. Considering the bi-objective optimization, Pareto-
fluids
optimal solutions for models 1, 2, 3 and 4 in pairs of (exergy effi-
ciency (%), LEC ($/kW h)) are (54.08, 0.204), (54.02, 0.201), (53.79,
To compare the thermoeconomic performance between pure
0.204), and (53.65, 0.203), respectively. The optimum operation
and mixture working fluids, the mass fraction of R245fa is selected
parameters with Pareto-optimal solution for models 1, 2, 3 and 4
as decision variables for mixture working fluids with the range of
are 364.70 K, 362.90 K, 364.32 K and 364.66 K of evaporator
0.1–0.9. Using the same method as discussed in Section 4.2, the
Pareto frontiers of pure working fluids (R245fa and R227ea) and
mixture working fluids for four models can be obtained and dis-
0.23 played in Figs. 14 and 15, respectively. The corresponding detailed
R245fa optimal values of decision variables and objective functions for
R227ea
0.22 C pure working fluids and mixture working fluids are shown in
Tables 5 and 6, respectively. The curves for the pure and mixture
working fluids can be expressed as follows:
0.21
LEC ($/kWh)

LEC ¼ 385:2g3ex  576:4g2ex þ 287:4gex  47:6


0.20
R2 ¼ 0:992 ð47:19% 6 gex 6 54:68%for R245faÞ ð54Þ

0.19 B
LEC ¼ 535:7g3ex  853:8g2ex þ 453:6gex  80:2
0.18 A
A' B' C'
R2 ¼ 0:997 ð53:53% 6 gex 6 54:36% for R227eaÞ ð55Þ

0.17 LEC ¼ 482:8g3ex  792:8g2ex þ 434:1gex  79:1


0.46 0.48 0.50 0.52 0.54 0.56
Exergy efficiency R2 ¼ 0:979 ð53:02% 6 gex 6 57:60% for model 1Þ ð56Þ

Fig. 14. Pareto frontiers of LEC with exergy efficiency for R245fa and R227ea.
LEC ¼ 336:1g3ex  540:5g2ex þ 289:5gex  51:5
R2 ¼ 0:931 ð49:53% 6 gex 6 59:35% for model 2Þ ð57Þ
0.26

0.25 Model 1
C'' LEC ¼ 280:8g3ex  449:2g2ex þ 239:5gex  42:4
Model 2 C'
0.24 Model 3 R2 ¼ 0:987 ð50:26% 6 gex 6 59:23% for model 3Þ ð58Þ
Model 4
0.23
LEC ¼ 183:0g3ex  291:9g2ex þ 155:3gex  27:4
LEC ($/kWh)

0.22 C'''
0.21
R2 ¼ 0:981 ð49:70% 6 gex 6 58:49% for model 4Þ ð59Þ

0.20 B''
C In case of the single-objective of maximizing exergy efficiency
B'''
for two pure working fluids, the higher exergy efficiency is
0.19
B' obtained for R245fa (54.68%), while R227ea owns the lower exergy
0.18 A' A'' B efficiency (54.36%). Compared to the maximum exergy efficiency
0.17 of R245fa, the improvement in exergy efficiency for models 1, 2,
A
A''' 3 and 4 are 5.96%, 9.18%, 8.96% and 7.50%, respectively. As for
0.16
0.48 0.50 0.52 0.54 0.56 0.58 0.60 the minimum LEC, R245fa presents lowest LEC of 0.174 $/kW h,
while the highest LEC is obtained for R227ea of 0.189 $/kW h.
Exergy efficiency
The same value of minimum LEC with R245fa is obtained for mod-
Fig. 15. Pareto frontiers of LEC with exergy efficiency for models 1, 2, 3 and 4. els 2, 3 and 4, while model 1 presents a slight increment in LEC. It
indicates that when mass fraction is considered as decision

Table 5
Optimized values of objective and decision variables for points A–C on the Pareto frontier for R245fa and R227ea.

Working fluids Optimal solution Decision makings Design variables Objectives


T E (K) T C (K) DT sup (K) DT pp;E (K) g_ ex (%) LEC ($/kW h)

R245fa NAGA-II TOPSIS (point B) 362.07 303.00 3.12 3.00 53.18 0.188
Max g_ ex Point C 370.00 303.00 3.00 3.02 54.68 0.219
Min LEC Point A 348.86 303.00 5.61 5.78 47.19 0.174
R227ea NAGA-II TOPSIS (point B0 ) 369.99 303.00 3.00 3.30 54.03 0.190
Max g_ ex Point C0 370.00 303.00 3.00 3.00 54.36 0.191
Min LEC Point A0 369.86 303.00 3.01 3.76 53.53 0.189
870 Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872

Table 6
Optimized values of objective and decision variables for points A–C on the Pareto frontier for four models.

Models Optimal solution Decision makings Design variables Objectives


T E (K) T C (K) DT sup (K) DT pp;E (K) x g_ ex (%) LEC ($/kW h)

Model 1 NAGA-II TOPSIS (point B) 354.29 303.24 4.30 3.21 0.41 56.71 0.188
Max g_ ex Point C 354.30 303.20 4.28 3.01 0.55 57.60 0.120
Min LEC Point A 353.90 303.00 4.34 4.06 0.90 53.02 0.177
Model 2 NAGA-II TOPSIS (point B0 ) 356.78 303.18 4.09 3.05 0.44 57.67 0.192
Max g_ ex Point C0 369.65 303.13 4.53 3.00 0.58 59.35 0.246
Min LEC Point A0 346.99 303.41 3.77 5.48 0.90 49.53 0.174
Model 3 NAGA-II TOPSIS (point B00 ) 358.31 304.24 3.93 4.44 0.52 57.11 0.194
Max g_ ex Point C00 368.86 303.02 7.01 3.00 0.61 59.23 0.249
Min LEC Point A00 348.56 303.44 3.02 5.37 0.90 50.26 0.174
Model 4 NAGA-II TOPSIS (point B000 ) 360.05 304.22 3.00 4.50 0.43 56.91 0.192
Max g_ ex Point C000 366.36 304.81 7.16 3.01 0.57 58.49 0.218
Min LEC Point A000 347.58 303.00 4.65 6.18 0.90 49.70 0.174

variables, the mixtures present better thermodynamic perfor- kW h for R245fa, and 54.03% and 0.190 $/kW h for R227ea, respec-
mance and equivalent economic performance than the pure work- tively. Taking the mass fraction into account, Pareto-optimal solu-
ing fluids. tions for models 1, 2, 3 and 4 in pairs of (exergy efficiency (%), LEC
In case of Pareto-optimal solutions for two pure working fluids, ($/kW h)) are (56.71, 0.188), (57.67, 0.192), (57.11, 0.194), and
the optimum exergy efficiency and LEC are 53.18% and 0.188 $/ (56.91, 0.192), respectively. Compared with the pure working flu-

380 12
R245fa
R227ea
370 10 Model 1
Evaporator temperature (K)

Model 2
Degree of superheat (K)

Model 3
360 Model 4
8

350
6
R245fa
340 R227ea
Model 1
Model 2 4
330
Model 3
Model 4
320 2
0 10 20 30 40 0 10 20 30 40
Population Population

Fig. 16. Scattered distribution of evaporator temperature for R245fa, R227ea and Fig. 18. Scattered distribution of degree of superheat for R245fa, R227ea and four
four models with population in Pareto frontier. models with population in Pareto frontier.

314 12
R245fa
Pinch point temperature difference (K)

R227ea
312 Model 1
10
Condenser temperature (K)

R245fa Model 2
R227ea Model 3
310 Model 1 Model 4
Model 2 8
Model 3
308 Model 4
6
306

4
304

302 2
0 10 20 30 40 0 10 20 30 40
Population Population

Fig. 17. Scattered distribution of condenser temperature for R245fa, R227ea and Fig. 19. Scattered distribution of pinch point temperature difference for R245fa,
four models with population in Pareto frontier. R227ea and four models with population in Pareto frontier.
Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872 871

1.0 pure working fluids. Whether the mixtures exhibit better


thermodynamic and economic performance than the pure
working fluids depend on the operation parameters and
0.8 mass fraction of mixtures.
Mass fraction of R245fa

(3) From the viewpoint of single-objective optimization of max-


imum exergy efficiency and minimum LEC using
0.6
0.7R245fa/0.3R227ea, model 2 (T 7 ¼ T E ; T 3 ¼ T C ) is the
favorable operation condition for its highest thermodynamic
performance and lowest economic factor.
0.4
(4) Taking the mass fraction as decision variable and compared
with the pure working fluids from the single-objective of
0.2 Model 1
maximum exergy efficiency and minimum LEC, the mixtures
Model 2 present better thermodynamic performance and approach-
Model 3 ing equivalent economic performance than the pure working
Model 4 fluids.
0.0
0 10 20 30 40 (5) Taking the mass fraction into account, Pareto-optimal solu-
Population tions for models 1, 2, 3 and 4 in pairs of (exergy efficiency
(%), LEC ($/kW h)) are (56.71, 0.188), (57.67, 0.192), (57.11,
Fig. 20. Scattered distribution of mass fraction of R245fa for four models with
0.194), and (56.91, 0.192), respectively. Compared with the
population in Pareto frontier.
pure working fluids, the mixture working fluids present bet-
ter exergy efficiency but worse LEC except model 1 which
exhibits the same value of LEC as pure R245fa. The optimum
ids, the mixture working fluids present better exergy efficiency but
mass fraction of R245fa for models 1, 2, 3, and 4 are in range of
worse LEC except model 1, which exhibits the same value of LEC as
0.32–0.90, 0.29–0.90, 0.47–0.90, and 0.34–0.90, respectively.
pure R245fa.
The distributions of decision variables corresponding to Pareto
frontier for pure and mixture working fluids are displayed in Acknowledgment
Figs. 16–20. The upper bound and lower bound (the dash line in
Figs. 16–20) are marked to distinguish the location of decision vari- The authors gratefully acknowledge financial support from
ables in their respective ranges. For pure and mixture working flu- China Scholarship Council (CSC).
ids, the optimum evaporator temperature in the upper boundary,
while the condenser temperature tends to become as low as possi-
ble. Furthermore, the optimum values of pinch point temperature References
difference and degree of superheat are varied from the lower
[1] Jawahara CP, Saravananb R, Brunoc Joan Carles, Coronasc Alberto. Simulation
bound to the upper bound. The optimum mass fraction of R245fa studies on gax based Kalina cycle for both power and cooling applications.
for models 1, 2, 3, and 4 are in range of 0.32–0.90, 0.29–0.90, Appl Therm Eng 2013;50(2):1522–9.
0.47–0.90, and 0.34–0.90, respectively. [2] Peng S, Hong H, Jin HG, Wang ZF. An integrated solar thermal power system
using intercooled gas turbine and Kalina cycle. Energy 2012;44(1):732–40.
[3] Goswami DY. Solar thermal power technology: present status and ideas for the
future. Energy Sources 1998;20(2):137–45.
5. Conclusion [4] Zamfirescu C, Dincer I. Thermodynamic analysis of a novel ammonia-water
trilateral Rankine cycle. Thermochim Acta 2008;477(1):7–15.
Based on the thermoeconomic multi-objective optimization, [5] Hung TC, Shai TY, Wang SK. A review of organic Rankine cycles (ORCs) for the
recovery of low-grade waste heat. Energy 1997;22:661–7.
simultaneously considering exergy efficiency and levelized energy [6] Tchanche BF, Lambrinos G, Frangoudakis A, Papadakis G. Low-grade heat
cost (LEC), the thermoeconomic comparisons between pure and conversion into power using organic Rankine cycles – a review of various
mixture working fluids have been investigated. Four models are applications. Renew Sust Energy Rev 2011;15(8):3963–79.
[7] Vélez F, Segovia JJ, Martín MC, Antolín G, Chejne F, Quijano A. A technical,
proposed based on the different location of evaporating bubble
economical and market review of organic Rankine cycles for the conversion of
point temperature or condensing dew point temperature for the low-grade heat for power generation. Renew Sust Energy Rev 2012;16
mixture working fluids. The effects of mass fraction and four key (6):4175–89.
parameters (evaporator temperature, condenser temperature, [8] Roy JP, Mishra MK, Ashok Misra. Performance analysis of an Organic Rankine
Cycle with superheating under different heat source temperature conditions.
pinch point temperature difference and degree of superheat) on Appl Energy 2011;88:2995–3004.
exergy efficiency and levelized energy cost (LEC) are examined. [9] Rayegan R, Tao YX. A procedure to select working fluids for solar organic
Pareto-optimal solutions of four models using 0.7R245fa/0.3R227ea Rankine cycles (ORCs). Renew Energy 2011;36(2):659–70.
[10] Guo T, Wang HX, Zhang SJ. Fluids and parameters optimization for a novel
are obtained and compared. The mass fraction of R245fa is used as cogeneration system driven by low-temperature geothermal sources. Energy
decision variable in the thermoeconomic optimization for mixture 2011;36(5):2639–49.
working fluids. Further investigation of the thermoeconomic com- [11] Al-Sulaiman FA, Hamdullahpur F, Dincer I. Greenhouse gas emission and
exergy assessments of an integrated organic Rankine cycle with a biomass
parisons between pure and mixture working fluids is discussed. combustor for combined cooling, heating and power production. Appl Therm
The main conclusions drawn from the investigation are summa- Eng 2011;31(4):439–46.
rized as follows: [12] Sun F, Ikegami Y, Jia B, Arima H. Optimization design and exergy analysis of
organic Rankine cycle in ocean thermal energy conversion. Appl Ocean Res
2012;35:38–46.
(1) With the increase in mass fraction of R245fa, the exergy effi- [13] Katsanos CO, Hountalas DT, Pariotis EG. Thermodynamic analysis of a Rankine
ciencies of models 1 and 3 initially increase and then decrease, cycle applied on a diesel truck engine using steam and organic medium.
Energy Convers Manage 2012;60:68–76.
while that of model 2 presents a reverse trend. The exergy effi-
[14] He M, Zhang X, Zeng K, Gao K. A combined thermodynamic cycle used for
ciency of model 4 increases smoothly with mass fraction of waste heat recovery of internal combustion engine. Energy 2011;36
R245fa. The LEC of four models have a same behavior of an (12):6821–9.
increase first and then decrease with mass fraction of R245fa. [15] Yari M, Mahmoudi SMS. Utilization of waste heat from GT-MHR for power
generation in organic Rankine cycles. Appl Therm Eng 2010;30(4):366–75.
(2) The mixtures don’t always present better thermodynamic [16] Bao JJ, Zhao L. A review of working fluid and expander selections for organic
performance and worse economic performance than the Rankine cycle. Renew Sust Energy Rev 2013;24:325–42.
872 Y. Feng et al. / Energy Conversion and Management 106 (2015) 859–872

[17] Hung TC, Wang SK, Kuo CH, Pei BS, Tsai KF. A study of organic working fluids [45] Braimakis K, Preißinger M, Brüggemann D, Karellas S, Panopoulos K. Low grade
on system efficiency of an ORC using low-grade energy sources. Energy waste heat recovery with subcritical and supercritical Organic Rankine Cycle
2010;35(3):1403–11. based on natural refrigerants and their binary mixtures. Energy 2015. http://
[18] Hung TC. Waste heat recovery of organic Rankine cycle using dry fluids. Energy dx.doi.org/10.1016/j.energy.2015.03.092.
Convers Manage 2001;42:539–53. [46] Mavrou P, Papadopoulos AI, Stijepovic MZ, Seferlis P, Linke P, Voutetakis S.
[19] Roy JP, Mishra MK, Misra A. Performance analysis of an Organic Rankine Cycle Novel and conventional working fluid mixtures for solar Rankine cycles:
with superheating under different heat source temperature conditions. Appl performance assessment and multi-criteria selection. Appl Therm Eng
Energy 2011;88(9):2995–3004. 2015;75:384–96.
[20] Saleh B, Koglbauer G, Wendland M, Fischer J. Working fluids for low- [47] Le VL, Kheiri A, Feidt M, Pelloux-Prayer S. Thermodynamic and economic
temperature organic Rankine cycles. Energy 2007;32:1210–21. optimizations of a waste heat to power plant driven by a subcritical ORC
[21] Fernandez FJ, Prieto MM, Suorez I. Thermodynamic analysis of high- (Organic Rankine Cycle) using pure or zeotropic working fluid. Energy
temperature regenerative organic Rankine cycles using siloxanes as working 2014;78:622–38.
fluids. Energy 2011;36:5239–49. [48] Feng YQ, Zhang YN, Li BX, Yang JF, Shi Y. Sensitivity analysis and
[22] Lai NA, Wendland M, Fischer J. Working fluids for high-temperature organic thermoeconomic comparison of ORCs (organic Rankine cycles) for low
Rankine cycles. Energy 2011;36:199–211. temperature waste heat recovery. Energy 2015;82:664–77.
[23] Wang J, Sun Z, Dai Y, Ma SL. Parametric optimization design for supercritical [49] Feng YQ, Zhang YN, Li BX, Yang JF, Shi Y. Comparison between regenerative
CO2 power cycle using genetic algorithm and artificial neural network. Appl organic Rankine cycle (RORC) and basic organic Rankine cycle (BORC) based on
Energy 2010;87:1317–24. thermoeconomic multi-objective optimization considering exergy efficiency
[24] Vetter C, Wiemer H, Kuhn D. Comparison of sub- and supercritical Organic and levelized energy cost (LEC). Energy Convers Manage 2015;96:58–71.
Rankine Cycles for power generation from low-temperature low-enthalpy [50] Al-Sulaiman FA. Exergy analysis of parabolic trough solar collectors integrated
geothermal wells, considering specific net power output and efficiency. Appl with combined steam and organic Rankine cycles. Energy Convers Manage
Therm Eng 2013;51:871–9. 2014;77:441–9.
[25] Shu GQ, Gao YU, Tian H, Wei HQ, Liang XY. Study of mixtures based on [51] Incropera FP, Dewitt DP, Bergman TL, Lavine AS. In: Xinshi Ge, Hong Ye,
hydrocarbons used in ORC (Organic Rankine Cycle) for engine waste heat editors. Fundamentals of heat and mass transfer, 6th ed., vol. 318. Chemistry
recovery. Energy 2014;74:428–38. Industry Press; 2007. Trans. [in Chinese].
[26] Heberle F, Preißinger M, Brüggemann D. Zeotropic mixtures as working fluids [52] García-Cascales JR, Vera-García F, Corberán-Salvador JM, Gonzálvez-Maciá J.
in Organic Rankine Cycles for low-enthalpy geothermal resources. Renew Assessment of boiling and condensation heat transfer correlations in the
Energy 2012;37(1):364–70. modelling of plate heat exchangers. Int J Refrig 2007;30(6):1029–41.
[27] Zhang J, Zhang HG, Yang K, Yang FB, Wang Z, Zhao GY, et al. Performance [53] Kandlikar SG. A general correlation for saturated two-phase flow boiling heat
analysis of regenerative organic Rankine cycle (RORC) using the pure working transfer inside horizontal and vertical tubes. J Heat Transfer 1990;112
fluid and the zeotropic mixture over the whole operating range of a diesel (1):219–28.
engine. Energy Convers Manage 2014;84:282–94. [54] Yan YY, Lio HC, Lin TF. Condensation heat transfer and pressure drop of
[28] Dong BS, Xu GQ, Cai Y, Li HW. Analysis of zeotropic mixtures used in high- refrigerant R-134a in a plate heat exchanger. Int J Heat Mass Transfer 1999;42
temperature Organic Rankine cycle. Energy Convers Manage 2014;84:253–60. (6):993–1006.
[29] Li W, Feng X, Yu LJ, Xu J. Effects of evaporating temperature and internal heat [55] Cayer E, Galanis N, Nesreddine H. Parametric study and optimization of a
exchanger on organic Rankine cycle. Appl Therm Eng 2011;31(17):4014–23. transcritical power cycle using a low temperature source. Appl Energy
[30] Garg P, Kumar P, Srinivasan K, Dutta P. Evaluation of isopentane, R-245fa and 2010;87:1349–57.
their mixtures as working fluids for organic Rankine cycles. Appl Therm Eng [56] Nafey AS, Sharaf MA. Combined solar organic Rankine cycle with reverse
2013;51(1):292–300. osmosis desalination process: energy, exergy, and cost evaluations. Renew
[31] Chys M, van den Broek M, Vanslambrouck B, De Paepe M. Potential of Energy 2010;35(11):2571–80.
zeotropic mixtures as working fluids in organic Rankine cycles. Energy [57] El-Emam RS, Dincer I. Exergy and exergoeconomic analyses and optimization
2012;44(1):623–32. of geothermal organic Rankine cycle. Appl Therm Eng 2013;59(1):435–44.
[32] Yang K, Zhang HG, Wang Z, Zhang J, Yang FB, Wang EH, et al. Study of zeotropic [58] Li YR, Du MT, Wu CM, Wu SY, Liu C. Economical evaluation and optimization of
mixtures of ORC (organic Rankine cycle) under engine various operating subcritical organic Rankine cycle based on temperature matching analysis.
conditions. Energy 2013;58:494–510. Energy 2014;68:238–47.
[33] Chen HJ, Goswami DY, Rahman MM, Stefanakos EK. A supercritical Rankine [59] Wang JY, Wang JF, Dai YP, Zhao P. Thermodynamic analysis and optimization
cycle using zeotropic mixture working fluids for the conversion of low-grade of a transcritical CO2 geothermal power generation system based on the cold
heat into power. Energy 2011;36(1):549–55. energy utilization of LNG. Appl Therm Eng 2014;70:531–40.
[34] Lecompte S, Ameel B, Ziviani D, van den Broek M, De Paepe M. Exergy analysis [60] Elsayed K, Lacor C. Modeling and Pareto optimization of gas cyclone separator
of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy performance using RBF type artificial neural networks and genetic algorithms.
Convers Manage 2014;85:727–39. Powder Technol 2012;217:84–99.
[35] Zhao L, Bao JJ. The influence of composition shift on organic Rankine cycle [61] Deb K, Pratap A, Agarwal S, Meyarivan T. A fast and elitist multi objective
(ORC) with zeotropic mixtures. Energy Convers Manage 2014;83:203–11. genetic algorithm: NSGA-II. IEEE Trans Evol Comput 2002;6:182–97.
[36] Yue C, Han D, Pu WH, He WF. Thermal matching performance of a geothermal [62] Zitzler E, Deb K, Thiele L. Comparison of multi objective evolutionary
ORC system using zeotropic working fluids. Renew Energy 2015;80:746–54. algorithms: empirical results. Evol Comput 2000;8:173–95.
[37] Li SL, Dai YP. Thermo-economic analysis of waste heat recovery ORC using [63] Monghasemi S, Nikoo MR, Fasaee MAK, et al. A novel multi criteria decision
zeotropic mixtures. J Energy Eng 2014. http://dx.doi.org/10.1061/(ASCE) making model for optimizing time–cost–quality trade-off problems in
EY.1943-7897.0000245. construction projects. Expert Syst Appl 2015;42(6):3089–104.
[38] Liu Q, Duan Y, Yang Z. Effect of condensation temperature glide on the [64] Li YQ, Liao SM, Liu G. Thermo-economic multi-objective optimization for a
performance of organic Rankine cycles with zeotropic mixture working fluids. solar-dish Brayton system using NSGA-II and decision making. Elect Power
Appl Energy 2014;115:394–404. Energy Syst 2015;64:167–75.
[39] Wang J, Zhao L, Wang X. A comparative study of pure and zeotropic mixtures [65] Ahmadi MH, Hosseinzade H, Sayyaadi H, Mohammadi AH, Kimiaghalam F.
in low-temperature solar Rankine cycle. Appl Energy 2010;87(11):3366–73. Application of the multi-objective optimization method for designing a
[40] Jung HC, Taylor L, Krumdieck S. An experimental and modelling study of a powered Stirling heat engine: design with maximized power, thermal
1 kW organic Rankine cycle unit with mixture working fluid. Energy 2015. efficiency and minimized pressure loss. Renew Energy 2013;60:313–22.
http://dx.doi.org/10.1016/j.energy.2015.01.003. [66] Sadatsakkak SA, Ahmadi MH, Ahmadi MA. Thermodynamic and thermo-
[41] Andreasen JG, Larsen U, Knudsen T, Pierobon L, Haglind F. Selection and economic analysis and optimization of an irreversible regenerative closed
optimization of pure and mixed working fluids for low grade heat utilization Brayton cycle. Energy Convers Manage 2015;94:124–9.
using organic Rankine cycles. Energy 2014;73:204–13. [67] Sayyaadi H, Mehrabipour R. Efficiency enhancement of a gas turbine cycle
[42] Liu Q, Shen A, Duan Y. Parametric optimization and performance analyses of using an optimized tubular recuperative heat exchanger. Energy 2012;38
geothermal organic Rankine cycles using R600a/R601a mixtures as working (1):362–75.
fluids. Appl Energy 2015;148:410–20. [68] Ahmadi MH, Sayyaadi H, Dehghani S, Hosseinzade H. Designing a solar
[43] Xiao L, Wu SY, Yi TT, Liu C, Li YR. Multi-objective optimization of evaporation powered Stirling heat engine based on multiple criteria: maximized thermal
and condensation temperatures for subcritical organic Rankine cycle. Energy efficiency and power. Energy Convers Manage 2013;75:282–91.
2015. http://dx.doi.org/10.1016/j.energy.2015.02.081. [69] NIST Standard Reference Database 23. NIST thermodynamic and transport
[44] Li YR, Du MT, Wu CM, Wu SY, Liu C. Potential of organic Rankine cycle using properties of Refrigerants and Refrigerant Mixtures REFROP. Version 9.0. 2010.
zeotropic mixtures as working fluids for waste heat recovery. Energy
2014;77:509–19.

Вам также может понравиться