You are on page 1of 14

Performance optimization of low-temperature power generation by

supercritical ORCs (organic Rankine cycles) using low GWP (global


warming potential) working uids
Van Long Le
a,
*
, Michel Feidt
a
, Abdelhamid Kheiri
a
, Sandrine Pelloux-Prayer
b
a
University of Lorraine, Laboratory of Energetics & Theoretical & Applied Mechanics, 2, Avenue de la Fort de Haye, 54518 Vanduvre-ls-Nancy, France
b
EDF-R&D, France, Eco-Efciency and Industrial Processes Department, Sites des Renardires, 77818 Moret-sur-Loing Cedex, France
a r t i c l e i n f o
Article history:
Received 14 September 2013
Received in revised form
19 November 2013
Accepted 13 December 2013
Available online 11 January 2014
Keywords:
Supercritical ORCs (organic Rankine cycles)
Waste heat recovery
Low-GWP (global warming potential)
refrigerants
Performance optimization
Exergy analysis
a b s t r a c t
This paper presents the system efciency optimization scenarios of basic and regenerative supercritical
ORCs (organic Rankine cycles) using low-GWP (global warming potential) organic compounds as
working uid. A more common refrigerant, i.e. R134a, was also employed to make the comparison. A
150-

C, 5-bar-pressurized hot water is used to simulate the heat source medium. Power optimization was
equally performed for the basic conguration of supercritical ORC. Thermodynamic performance com-
parison of supercritical ORCs using different working uids was achieved by ranking method and exergy
analysis method. The highest optimal efciency of the system (h
sys
) is always obtained with R152a in
both basic (11.6%) and regenerative (13.1%) congurations. The highest value of optimum electrical power
output (4.1 kW) is found with R1234ze. By using ranking method and considering low-GWP criterion, the
best working uids for system efciency optimization of basic and regenerative cycles are R32 and
R152a, respectively. The best working uid for net electrical power optimization of basic cycle is R1234ze.
Although CO
2
has many desirable environmental and safety properties (e.g. zero ODP (Ozone Depletion
Potential), ultra low-GWP, non toxicity, non ammability, etc.), the worst thermodynamic performance is
always found with the cycle using this compound as working uid.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
Nowadays, the world population is constantly growing, and so is
the world energy consumption. However fossil energy resources,
principal energy sources for human activities, are increasingly
running out. One of the most energy-consuming sectors is industry.
By the year 2010, the energy consumption of industrial sector was
accounted for about 27.9% of the world total nal consumption [1].
Furthermore, the power generation from fossil fuels causes many
environmental problems (e.g. global warming, air pollution, acid
rain, etc.). Therefore, capturing and converting industrial waste
heat into electricity is currently arousing much attention. This
improves not only energy efciency of the industrial processes, but
also reduces the thermal pollution caused by the direct release of
this heat into the environment. In industrial processes, a large
amount of energy input is always rejected as low-temperature
waste heat (exhaust gas, hot product stream, hot surfaces, etc.).
Regarding power generation from low-grade heat source,
because of the incompatibility of steam power cycle (normal
boiling point temperature of water is too high), many other kinds of
thermodynamic cycles have been studied for low-temperature
power production over the last decades, i.e. Kalina cycle [2e11],
Goswami cycle [12], TLC (Trilateral cycle) [13,14] and ORCs (organic
Rankine cycles) [15e21]. Bombarda et al. [22] found that although
the obtained net powers are similar for the Kalina cycle and ORC,
the Kalina cycle requires a very high maximum pressure in order to
reach high thermodynamic performances (100 bars compared to
about 10 bars for the ORC). These authors also concluded that the
adoption of Kalina cycle does not seem to be justied because the
gain in performance with regard to a properly optimized ORC is
very small and must be reached with a complicated plant scheme,
large surface heat exchangers and particular high-pressure resis-
tant and non-corrosive materials. In another work, Fischer [23]
found that exergy efciency for power production is higher by
14%e29% for the TLC with the two-phase expander utilization in
comparison to the ORC. But the two-phase expander is currently
commercially not available and R&D effort would be necessary for
development [24]. Indeed, among these thermodynamic cycles, the
organic Rankine cycle, which is similar to steam Rankine cycle but
* Corresponding author. Tel.: 33 383 595 592; fax: 33 383 595 551.
E-mail addresses: van-long.le@univ-lorraine.fr, long.le-van@outlook.com (V.
L. Le).
Contents lists available at ScienceDirect
Energy
j ournal homepage: www. el sevi er. com/ l ocat e/ energy
0360-5442/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.energy.2013.12.027
Energy 67 (2014) 513e526
uses an organic compound as working uid, is less complex and
requires less maintenance than other cycles [25]. Furthermore, the
technological maturity of ORC components due to their extensive
use in refrigeration applications explains why these cycles have
aroused much interest for low-temperature power generation.
In many state-of-the-art applications, the subcritical ORC, in
which the saturated or slightly superheated vapor is expanded
across a turbine, is often used for waste heat recovery. However, the
investigation of the supercritical (also called transcritical in the
literature) ORC is of high importance, since it may lead to higher
efciency making these equipments even more attractive for waste
heat recovery applications [26]. The main advantage of the super-
critical ORC over the subcritical one is a better match between the
heat sources cooling curve and the working uids heating curve
[24]. The absence of isothermal evaporation in supercritical ORC
enables the heat source to be cooled down to a lower temperature
despite an identical pinch point as in a comparable subcritical cycle.
This leads to a greater utilization of the waste heat resource.
The present work introduces the current challenges of low-
temperature waste heat recovery and the potential of the low-
temperature power generation from industrial heat losses. The
importance of the choice of the working uid and some criteria for
its selection are equally mentioned. The main objective of this pa-
per focuses on the performance optimization of the supercritical
ORCs (basic and regenerative ones) using low-GWP (global warm-
ing potential) organic compounds as working media. The ranking
method and the exergy analysis are employed to assess the ther-
modynamic performance of the optimized supercritical ORCs.
2. Low-temperature waste heat recovery potential and
challenge
Industrial waste heat refers to the energy that is generated in
industrial processes without being of practical use. The sources of
this waste include hot combustion gases discharged to the atmo-
sphere, heated products exiting industrial processes and heat
transfer from hot equipment surfaces [27]. In other words, in every
industrial process, waste heat is always generated as a byproduct
and released into the atmosphere. Various studies [28e31] have
estimated that as much as 20e50% of industrial energy inputs is
lost as waste heat. For example, according to the study of Niget et al.
[32] in 2008 the quantity of waste heat represents 40% of the en-
ergy consumed by the industrial sectors studied in Quebec. These
authors also estimated that the energy potential of industrial waste
heat (industries and thermal power plants) represents 15.3% of
consumed energy. The value of this energy deposit is estimated at
$ 2.9 milliard corresponding to 1.1% Quebec GDB (Gross Domestic
Product) by the year 2008. As the industrial sector is pursuing its
efforts to improve its energy efciency, recovering waste heat los-
ses provides an attractive opportunity for an emission-free and
less-costly energy resource [27]. In some cases, waste heat recovery
can improve energy efciency of the industrial processes by 10% to
as much as 50% [29].
Even though low-temperature waste heat possesses a lower
enthalpy and economic value than high-temperature waste, this
heat source is much abundant. Indeed, approximately 60% of waste
heat is below 230

C, and nearly 90% of waste heat is below 316

C
as reported by BCS Inc. [27]. Three major challenges for low-
temperature waste heat recovery was also mentioned in this
report. The rst one is the corrosion of heat exchanger surfaces, as
the ue gases contain varying concentrations of carbon dioxide,
water vapor, NO
x
, SO
x
, unoxidized organics and minerals. Indeed,
when the exhaust gases are cooled down below their dew point
temperature, the water vapor will condense and drop off the cor-
rosive substances on heat exchanger surfaces. This generally
implies using advanced materials, or frequently replacing heat
exchanger components, which is often uneconomical. Secondly,
small temperature gradient between uid streams induces a
requirement of large heat exchange surfaces. Finally, recovering
heat at low-temperature makes sense if there is an end-use for the
project. Nowadays, available technologies have enabled ue gases
to be cooled down below their dew point temperatures for the
waste heat recovery. On the other hand, beside of the potential
direct end uses of low-temperature waste heat recovery such as
low-temperature process heating, domestic water heating, and
space heating; heat pumps and low-temperature power generation
are also the potential solutions for the problemof end-use shortage.
The most frequently used system for power generation from
waste heat involves using heat to generate steamwhich then drives
a steam turbine. However, steam power cycles become less prof-
itable at low temperature (below 340

C) since low pressure steam
requires more voluminous equipments. Furthermore, the shortage
of energy to superheat the steamcauses steam condensation which
erodes turbine blades during expansion. The organic Rankine cycle
operates in a similar way to the steam Rankine cycle, but uses an
organic compound instead of water as working uid. These organic
uids have a lower boiling point temperature and a higher vapor
pressure than water. This enables low-temperature waste heat re-
covery by the ORC. The ORC technology is often used in power
generation from solar, geothermal and waste heat sources. Among
these applications, ORCs are particularly relevant for waste heat
recovery.
3. ORC working uids considering the next refrigerant
generation
For power generation from low-temperature waste heat by
Rankine cycle, water is not a suitable working uid due to its
relatively high boiling point temperature. For this purpose, organic
compounds (e.g. refrigerants, hydrocarbons, etc.) with low tem-
perature of normal boiling point present a real potential. In com-
parison with water vapor, organic uids have a higher molecular
mass enabling compact designs, higher mass ow, and higher
turbine isentropic efciency (as high as 80e85%) [27].
As the ORC performance is tremendously inuenced by the
thermo-physical properties of its working uid, many researchers
focus their works on the selection of working media for ORC ap-
plications [33e37]. The determination of an appropriate working
uid can be started with a literature review of existing organic
uids. Then, available temperature of heat source, ambient tem-
perature or coolant liquid temperature would be the most impor-
tant factors in the selection process. The next step would consist in
considering the environment (ODP (ozone depletion potential),
GWP), safety (toxicity, ammability), stability (thermal, chemical)
and compatibility criteria. Afterward, the comparisons between
thermodynamic behaviors and transport properties of the working
uids as well as the determination of cycle performances would be
carried out. The economy criteria of the working uid in particular
and of the system in general contribute to the feasibility of the
project. Depending on the objective, the order of priority of the
aforementioned criteria can be changed.
Currently, the CO
2
-transcritical Rankine cycle has been paid
much attention due to the desirable characteristics of CO
2
such as
stability, low environmental impact, low cost and abundance in
nature. However, the 60e160 bars operating pressure of the CO
2
transcritical cycle raises safety concerns and may be considered as
a disadvantage. Therefore, other uids have been studied for su-
percritical ORC application. Baik et al. [38] compared the perfor-
mance of a transcritical power cycle with two working uids: CO
2
and R125. The authors found that for an optimized power output
V.L. Le et al. / Energy 67 (2014) 513e526 514
the R125 transcritical cycle produced 14% more power than the
carbon dioxide transcritical one with the same conditions of heat
source, heat sink and a common value of total heat transfer surface
area. In another work, Chen et al. [39] found that the thermal
efciency (rst-law efciency) of the R32-based transcritical
Rankine cycle is higher than a CO
2
-based cycle at high tempera-
ture (120e180

C) and R32 operates at much lower pressures.
Cayer et al. [40] carried out a parametric study and optimization
with six performance indicators (i.e. rst-law efciency, net power
output, exergy efciency, total UA (the product of the overall heat
transfer coefcient and the heat transfer surface area) and surface
area of the heat exchangers as well as the relative cost of the
system) for transcritical power cycle. The results of this study
show that it is impossible to simultaneously optimize all six per-
formance indicators. A comparison of optimized results for three
working uids (CO
2
, ethane and R125) shows that R125 has the
best rst-law efciency, the lowest UA, heat transfer surface area
and cost while the highest specic net power output is obtained
with ethane. CO
2
has a higher total UA and a lower exchange
surface area than ethane. Some other working uids were used for
the performance comparison and parametric optimization of
subcritical and transcritical ORC in the work of Shengjun et al. [41].
For the transcritical cycle, the results showed that R125 has
excellent economical and environmental performance for a low-
temperature geothermal system. Guo et al. [42] used twelve nat-
ural and conventional working uids for their transcritical ORC
system with an internal heat exchanger and compared the per-
formance with the CO
2
-based cycle. In this study, R125 presented
the best performance for a given geothermal source temperature
of 90

C. R218 presents the highest net power output when the
geothermal source temperature is above 100

C but with an
expansion ratio 3.2 times higher on average and total heat transfer
areas 25% higher than baselines (CO
2
-based transcritical ORC).
Considering the net power output, total heat transfer surface
areas, volume ow rate at turbine inlet and expansion ratio, R32
and R143a are potential working uids for geothermal source
temperatures above 100

C. Zeotropic mixture working uids also
present a real potential for the conversion of low-grade heat into
electrical power. Chen et al. [43] proposed a supercritical Rankine
cycle using a zeotropic mixture as working uid for a comparative
study of a supercritical and subcritical ORC working under the
same thermal conditions. This study showed that the supercritical
power cycle with a zeotropic mixture of 0.7R134a/0.3R32 (mass
fraction) can achieve rst-law efciency of 10.77e13.35% with the
turbine inlet temperature of 120e180

C as compared to 9.70e
10.13% from a subcritical ORC using pure R134a. However, some
studies realized by GE (General Electric) group showed that the
ORC technology using uid mixtures is not cost effective, due to
heat exchanger design considerations to avoid de-mixing of the
uid mixture and lower heat transfer coefcients [24].
Nowadays, due to the environmental impact, especially the
depletion of the ozone layer, many synthetic chemicals were and
will be phased out in the future (Table 1). In addition, growing
awareness of climate change as a more signicant and more chal-
lenging environmental issue now announces a new generation of
refrigerants (fourth generation), to address global warming [44].
The historic progression of refrigerants, described by Calm [45],
encompasses four distinct intervals divided by changes in selection
criteria. The four generations include slightly overlapping periods
based on whatever worked (1830se1930s), improved safety and
durability (1931e1990s), stratospheric ozone protection (1990se
2010s) and attention to global warming mitigation (2012e?). Ac-
cording to the Calm and Hourahan [46] the fourth generation of
refrigerants is still to be determined, but likely to include re-
frigerants with very low or no ODP, low GWP and high efciency.
The candidates include low-GWP Hydrouorocarbons (HFCs), un-
saturated hydrouorochemicals, ammonia, hydrocarbons and wa-
ter. Currently the European regulations for mobile air conditioning
specify that as for 1st January 2017 air-conditioning systems
designed to contain uorinated greenhouse gases with a GWP
higher than 150 shall not be retrotted to vehicles type-approved
from that date [47].
In this study, some low-GWP refrigerants such as hydrocarbons,
HFOs (Hydrouoroolens); CO
2
and low-GWP HFCs, are used as
working uids for power generation by supercritical ORCs. The
thermo-physical properties and the environmental and safety data
of these working uids are respectively shown in Tables 2 and 3.
4. Supercritical (or transcritical) ORC modeling
In the literature, this kind of cycle could be called supercritical
(like supercritical steam cycle), or transcritical (like CO
2
Table 1
Summary of Montreal Protocol control measures [67].
Ozone depleting substances Developed countries Developing countries
Chlorouorocarbons (CFCs) Phased out end
of 1995
Total phase out by 2010
Halons Phased out end
of 1993
Total phase out by 2010
Carbon tetrachloride (CCl
4
) Phased out end
of 1995
Total phase out by 2010
Methyl chloroform (CH
3
CCl
3
) Phased out end
of 1995
Total phase out by 2015
Hydrochlorouorocarbons
(HCFCs)
Total phase out
by 2020
Total phase out by 2030
Hydrobromouorocarbons
(HBFCs)
Phased out end
of 1995
Phased out of 1995
Methyl bromide (CH
3
Br) Total phase out
by 2005
Total phase out by 2015
Bromochloromethane
(CH
2
BrCl)
Phase out by 2002 Phase out by 2002
Table 2
Thermo-physical properties of working uids [66].
Fluid Chemical formula MM (kg kmol
1
) T
bp
(

C) T
crit
(

C) P
crit
(bar) D
crit
(kg m
3
) DH
vap
b
(kJ kg
1
)
R134a CH
2
FCF
3
102.03 26.07 101.1 40.59 511.9 182.28
R152a CH
3
CHF
2
66.05 24.02 113.3 45.17 368.0 285.32
R32 CH
2
F
2
52.02 51.65 78.1 57.82 424.0 280.78
R744 CO
2
44.01 78.46
a
31.0 73.77 467.6 152.00
R1270 CH
3
CH]CH
2
42.08 47.62 91.1 45.55 229.6 344.28
R290 CH
3
CH
2
CH
3
44.10 42.11 96.7 42.51 220.5 344.31
R1234yf CH
2
=CFCF
3
114.04 29.45 94.7 33.82 475.6 149.29
R1234ze(E) CHF=CHCF
3
114.04 18.97 109.4 36.35 489.2 170.63
a
Normal sublimation point.
b
Heat of vaporization at 20

C.
V.L. Le et al. / Energy 67 (2014) 513e526 515
transcritical Rankine cycle) or even hypercritical cycle. Kim et al.
[48] distinguish that the working uid is, in the case of transcritical
cycle, compressed to pressure above its critical pressure and ex-
pands to subcritical pressure, while the working medium is used
entirely above its critical pressure with supercritical conguration.
In the present work, supercritical cycle is described as the cycle for
which the working uid goes through both under and above its
critical pressure. Over recent years, the supercritical ORC has
attracted more and more the attention for power generation from
low-temperature heat source [38e40,48e52] due to its advantages,
such as higher performance and less irreversibility. However, this
conguration of ORC also presents several challenges to overcome,
i.e. high pressure required, high pressure ratio in turbine, super-
critical heat transfer and the uncertainty of working uid proper-
ties in the supercritical region.
In the basic supercritical ORC (Fig. 1), the working uid is rst
compressed to the pressure above its critical one from slightly sub-
cooled liquid state. The working uid is then heated, without
isothermal phase change process, by hot water in high-
temperature heat exchanger. In next step, the working medium
releases its energy to drive the turbine blades during expansion
process. Finally, the working uid is cooled down and condensed by
the cooling water before being pumped again to high pressure. The
assumptions for supercritical ORC modeling are described below:
v Each process of the cycle is considered as a steady-state and
adiabatic process
v Heat and friction losses as well as the potential and kinetic
energy are neglected
The given parameters for the cycle modeling and the optimi-
zation are described in Table 4. The equations for the successive
thermodynamic transformations in the basic supercritical ORC (cf.
Fig. 1) are described as follows.
v Pumping process
h
p;is
=
h
2;is
h
1
h
2
h
1
(1)
_
W
p
= _ m
wf
(h
2
h
1
) (2)
P
p
=
_
W
p
=h
motor
(3)
v High-temperature heat transfer process
_
Q
HTHEX
= _ m
wf
(h
3
h
2
) = _ m
h
(h
hsi
h
hso
) (4)
v Expansion process
h
t;is
=
h
3
h
4;is
h
3
h
4
(5)
_
W
t
= _ m
wf
(h
3
h
4
) (6)
P
t
= h
gen
_
W
t
(7)
v Condensation process
_
Q
cond
= _ m
wf
(h
4
h
1
) = _ m
c
(h
cso
h
csi
) (8)
The pinch position of high-temperature heat exchanger in the
case of supercritical ORC is determined by an iteration like in the
work of Pan et al. [53]. The scheme of the iteration is described as in
Fig. 2. Firstly, an initial mass ow rate of working uid is given.
Then different temperatures of working uids heating curve are
calculated like in Eq. (9) with the temperature interval (dT) deter-
mined by Eq. (10). In this study, the high-temperature heat
exchanger is divided into 200 sections for temperature calculation
of the different points. The temperatures at different points of hot
waters cooling curve are calculated after the determination of their
enthalpy by Eq. (11). After the calculation of corresponding tem-
peratures on the heating and cooling curves, the temperature dif-
ferences between two streams can be determined by Eq. (12).
Afterward, the minimum temperature difference could be found. If
this value is equal to the given pinch value (Pinch
H
), then the mass
ow rate of working uid is the real one, otherwise it would be
increased or decreased if minimum temperature difference is
respectively superior or inferior to Pinch
H
.
T
wf;i1
= T
wf;i
dT (9)
dT = (T
3
T
2
)=200 (10)
h
h;i1
= h
h;i
_ m
wf

h
wf ;i1
h
wf;i
.
_ m
h
(11)
DT
i
= T
h;i
T
wf;i
(12)
To improve the cycle performance, some modied congura-
tions of ORC have been studied (e.g. regenerative and reheat cycle)
over recent years. Mago et al. [54,55] recognized that the regen-
erative ORC has higher rst- and second-law efciencies than the
basic cycle. These authors also afrmed that the performance of
reheat ORC is very similar to the performance of basic ORC [54]. In
another work, Xu et al. [56] used a vapor injector as a regenerator to
improve the cycle performance. The vapor injector is a simple and
compact device to pump cold liquid to produce an outlet liquid
with a higher pressure than the vapor inlet pressure. These authors
found that the rst-law efciency of the regenerative conguration
is 1.18% higher than basic ORC with optimal extraction pressure of
about 220 kPa. By using the regenerative ORC for solar thermal
electricity generation, Pei et al. [57] discovered that the maximum
efciency of regenerative ORC is 9.2% higher than without-
regenerative conguration. The maximum overall system ef-
ciency is also 4.6e5.4% greater with the regenerative ORC.
Table 3
Safety and environmental data of working uids [46].
Fluid OEL
(ppmv)
LFL (%) HOC
(MJ kg
1
)
Safety
group
Atmospheric
life (yr)
ODP GWP
100 yr.
R134a 1000 None 4.2 A1 13.4 0.000 1370
R152a 1000 4.8 17.4 A2 1.5 0.000 133
R32 1000 14.4 9.4 A2L r 5.2 0.000 716
R744 5000 None e A1 >50 0.000 1
R1270 500 2.7 e A3 0.001 0.000 <20
R290 1000 2.1 50.4 A3 0.041 0.000 w20
R1234yf 500 6.2 10.7 A2L r 0.029 0.000 <4.4
R1234ze(E) 1000 7.6 e e 0.045 0.000 6
OEL: occupational exposure limit; LFL: lower ammability limit (% volume in air);
HOC: heat of combustion.
A: lower toxicity; B: higher toxicity; 1: no ame propagation; 2: lower ammability;
3: higher ammability; A2L: lower ammability with a maximum burning velocity
of _10 cm/s.
Sufxes to safety classication indicate recommended changes that are not nal yet
(r for revision or addition).
V.L. Le et al. / Energy 67 (2014) 513e526 516
From the aforementioned references, there is no research of
supercritical ORC using heat regeneration. Indeed, the regenerative
conguration is often used with supercritical steam cycle but there
are not many studies of regenerative supercritical ORC in the
literature. In this study, the regenerative supercritical ORC (Fig. 3) is
used to improve the system efciency of the cycle. In this cong-
uration, a feed-uid heater (also called heater henceforth) is
incorporated into the basic cycle. The vapor exits the turbine at
point 6 (cf. Fig. 3) is mixed with the liquid fromthe rst pump in the
heater. The mixtures exit the feed-uid heater at slightly sub-
cooled liquid state to avoid pump cavitation phenomenon before
being pumped to high-temperature heat exchanger by the second
pump. The vapor exiting the turbine at point 7 is cooled down and
condensed by cooling water before being mixed with the vapor
exiting the turbine at point 6. The equations of regenerative cycle
components are similar to the ones of the basic supercritical ORC
described above. Two mass ow rates of working uid entering the
heater are calculated by Eqs. (13) and (14).
_ m
wf
= _ m
1
_ m
2
(13)
_ m
wf
h
3
= _ m
1
h
6
_ m
2
h
2
(14)
The supercritical ORC performance is determined by the equa-
tions as follows:
Net mechanical power output
_
W
net
=
_
W
t

_
W
p
(15)
For regenerative cycle
_
W
p
=
_
W
p1

_
W
p2
Net electrical power output
P
net
= P
t
P
p
P
wp
(16)
For regenerative cycle P
p
= P
p1
P
p2
First-law efciency
h
I
=
_
W
net
=
_
Q
HTHEX
(17)
System efciency
h
sys
= P
net
=
_
Q
HTHEX
(18)
5. Exergy analysis
According to the rst lawof thermodynamics, the energy cannot
be destroyed or created. It is conserved in any device or process
except the nuclear processes. However, for depicting some
important aspect of resource utilization, the rst lawanalysis alone
is not adequate. The useful (noble) energy loss of the system or
device cannot be justied by the rst law of thermodynamics,
because it does not distinguish the quality and quantity of energy.
Over last decades, the exergyanalysis (availabilityanalysis) basedon
the second lawof thermodynamics is found to be a useful method in
the design, evaluation, optimization and improvement of energy
systems. The term availability was made popular in the United
States by the M.I.Ts School of Engineering in the 1940s [58]. Indeed,
as mentionedinthe workof Feidt [59], the termusable energy was
inventedby Gouy (1889) and Stodola (1898) but until 1956, the term
exergy, which is preferred today, was introduced by Rant [60]. As
dened in a thermodynamic textbook [61], the exergy is referred to
the maximum theoretical work obtainable from an overall system
consisting of a system and the environment as the system comes
into equilibriumwith the environment (passes to the dead state). By
the exergy analysis method, the location, cause and true magnitude
of energy resource waste and loss can be determined [62].
Inthe exergyanalysis method, the temperature (T
0
) andpressure
(p
0
) of the environment (dead state) are often taken as standard-
state values (i.e. 25

C and 1 atm, respectively) for computational
ease. However, these properties may be specied differently
depending on the application [62]. In the present work, the inlet
temperature of cold water (15

C) and atmospheric pressure (1 atm)


are selected as the temperature and pressure of the dead state.
For a control volume, exergy can be transferred by three means:
exergy transfer associated with work, heat transfer, and with the
matter entering and exiting. Nevertheless exergy can be destroyed
by the irreversibility within the system or the control volume. At
steady state, the control volume exergy rate balance takes the form
[62] as follows
Fig. 1. Conguration and Tes diagram of basic supercritical ORC.
Table 4
Given parameters of supercritical ORCs.
Pump and turbine isentropic efciency (%) 80
Pump motor and electrical generator efciency (%) 90
Pinch
H
(

C) 10
Pinch
C
(

C) 3
Heat source/sink medium Water
Inlet pressure of the heat source medium (bar) 5
Inlet temperature of the heat source medium (

C) 150
Mass ow rate of the heat source medium (kg s
1
) 0.1
Inlet pressure of the heat sink medium (bar) 5
Inlet temperature of the heat sink medium (

C) 15
V.L. Le et al. / Energy 67 (2014) 513e526 517
X
j
_
Ex
q;j

_
W
cv

X
i
_
Ex
i

X
e
_
Ex
e

_
I = 0 (19)
Or
X
j

1
T
0
T
j
!
$
_
Q
j

_
W
cv

X
i
_ m
i
Ex
i

X
e
_ m
e
Ex
e

_
I = 0 (20)
Where
_
Ex
q;j
eis the exergy transfer rate associated with heat transfer
_
W
cv
eis the work rate of the control volume excluding the ow
work
_
Q
j
eis the heat transfer rate at the location on the boundary of
the control volume where the instantaneous temperature is T
j
Exeis the specic exergy and can be expressed in terms of four
components: physical exergy (Ex
PH
), kinetic exergy (Ex
KN
), po-
tential exergy (Ex
PT
) and chemical exergy (Ex
CH
). The subscripts
i and e denoted inlet and outlet, respectively.
Ex = Ex
PH
Ex
KN
Ex
PT
Ex
CH
(21)
Ex
PH
= h h
0
T
0
(s s
0
) (22)
Ex
KN
= V
2
=2 (23)
Ex
PT
= gz (24)
The chemical exergy for the ORC components is considered zero.
The exergy destruction rate is calculated by Eq. (25).
_
I = T
0
_
S
gen
(25)
v Exergy efciency
The exergy efciency in this study is dened as the ratio of the
useful exergy obtained to exergy consumed by system [63] by Eq.
(26).
h
Ex
=
_
Ex
useful
_
Ex
spent
(26)
v Second-law efciency
The second-law efciency is dened as the ratio of the rst-law
efciency of the system to the efciency of a reversible system
operating between the same thermodynamic states [58]
h
II
= h
I
=h
rev
(27)
For sensible heat source and sink, such as in a solar thermal
systemor waste heat fromhot exhaust gases etc., the efciency of a
reversible system is calculated by Eq. (28).
h
rev
= 1
~
T
c
~
T
h
(28)
In fact, this is the efciency of the equivalent Carnot cycle where
the entropic average temperatures of heat source (
~
T
h
) and heat sink
(
~
T
c
) media are calculated by Eqs. (29) and (30).
~
T
h
=
h
hsi
h
hso
s
hsi
s
hso
(29)
Fig. 2. Pinch positions and iterative process of the calculation of working uids mass
ow rate.
Fig. 3. Conguration and Tes diagram of regenerative supercritical ORC.
V.L. Le et al. / Energy 67 (2014) 513e526 518
~
T
c
=
h
csi
h
cso
s
csi
s
cso
(30)
As exergy is dened in relation to the environment (dead state),
the exergy efciency is affected by the choice of the references
conditions, while the second-lawefciency, as dened in Eq. (27), is
independent to the environmental state [64]. The second-law ef-
ciency is equal to the exergy efciency only if it is chosen that
T
0
=
~
T
c
. The equation of component exergy analysis in the case of
basic supercritical ORC (cf. Fig. 1) are described below.
v Pumping process
v Consumed exergy
_
W
p
= _ m
wf
(h
2
h
1
) (31)
v Useful exergy
_
Ex
p
= _ m
wf
[h
2
h
1
T
0
(s
2
s
1
)[ (32)
v Exergy destruction
_
I
p
= T
0
_ m
wf
(s
2
s
1
) (33)
v High-temperature heat transfer process
v Consumed exergy
_
Ex
h
= _ m
h
[h
hsi
h
hso
T
0
(s
hsi
s
hso
)[ (34)
v. Useful exergy
_
Ex
HTHEX
= _ m
wf
[h
3
h
2
T
0
(s
3
s
2
)[ (35)
v Exergy destruction
_
I
HTHEX
= T
0
h
_ m
wf
(s
3
s
2
) _ m
h
(s
hso
s
hsi
)
i
(36)
v Expansion process
v Consumed exergy
_
Ex
t
= _ m
wf
[h
3
h
4
T
0
(s
3
s
4
)[ (37)
v Useful exergy
_
W
t
= _ m
wf
(h
3
h
4
) (38)
v Exergy destruction
_
I
t
= T
0
_ m
wf
(s
4
s
3
) (39)
v Condensation process
v Consumed exergy
_
Ex
cond
= _ m
wf
[h
4
h
1
T
0
(s
4
s
1
)[ (40)
v Useful exergy
_
Ex
c
= _ m
c
[h
cso
h
csi
T
0
(s
cso
s
csi
)[ (41)
v Exergy destruction
_
I
cond
= T
0
h
_ m
wf
(s
1
s
4
) _ m
c
(s
cso
s
csi
)
i
(42)
For the regenerative cycle, the equations of exergy analysis of
the cycle components are described in the same manner. In addi-
tion, the exergy balance of feed-uid heater (cf. Fig. 3) are described
below.
v Consumed exergy
_
Ex
reg;h
= _ m
1
[h
6
h
3
T
0
(s
6
s
3
)[ (43)
v Useful exergy
_
Ex
reg;c
= _ m
2
[h
3
h
2
T
0
(s
3
s
2
)[ (44)
v Exergy destruction
_
I
reg
= T
0

_ m
wf
s
3
_ m
1
s
6
_ m
2
s
2

(45)
The system exergy analysis and second-law efciency of the
supercritical ORCs are represented by the following equations:
Total ow rate of irreversibility of the system
_
I
tot
=
X
_
I (46)
The exergy balance of basic supercritical ORC system
_
Ex
h

_
W
p
=
_
W
t

_
Ex
c

_
I
tot
(47)
The exergy balance of regenerative supercritical ORC
_
Ex
h
=
_
W
p1

_
W
p2
=
_
W
t

_
Ex
c

_
I
tot
(48)
The exergy efciency of the supercritical ORC
V.L. Le et al. / Energy 67 (2014) 513e526 519
h
Ex
=
_
W
t

_
W
p
_
Ex
h
(49)
In the case of regenerative cycle
_
W
p
=
_
W
p1

_
W
p2
Or the Eq. (49) can be rewritten as follows:
h
Ex
= 1
_
Ex
c

_
I
tot
_
Ex
h
(50)
The second-law efciency
h
II
=
h
I
1
~
Tc
~
T
h
(51)
Without a second utilization of heat source outlet, the ow rate
of exergy associated with the outlet hot water is considered as
exergy loss and calculated by Eq. (52).
_
Ex
h;loss
= _ m
h
[h
hso
h
0
T
0
(s
hso
s
0
)[ (52)
6. Performance optimization scenario
Many optimization methods are implemented in EES software
[65]. After some examinations, two methods are chosen to perform
the optimizations in this study
v The Direct Search method is used to perform the efciency
optimizations of supercritical ORCs.
v The Genetic Optimization method is used to perform the net
electrical power optimization
The Direct Search method is sometimes called the Conjugate
Directions method or Powells method. The basic idea of this
method is to use a series of one-dimensional searches to locate the
optimum. In its simplest form, EES will hold all but one of the
optimization variables constant, and then vary the single
remaining variable in order to locate the value at which the
objective function is maximized (or minimized) along the one-
dimensional path. This process is repeated for each independent
variable as many times as necessary to achieve the stopping
criteria. Although, there are four variables for the optimization, i.e.
outlet temperature of heat source medium (T
hso
), turbine inlet
pressure (P
h
), turbine inlet temperature (T
h
) and condensing
temperature (T
cond
), the system efciency only depends on the
turbine inlet pressure and temperature and condensing tempera-
ture of working uid. Outlet heat source temperature can be
recalculated after the determination of the three other variables
by optimization process. In the case of system efciency optimi-
zation, the same results are obtained by two used methods, but it
is very much faster with the Direct Search method.
In the case of power optimization, the optimal value of the
objective function (net electrical power output) is much harder to
be achieved. Indeed, the objective function strictly depends on all
four optimized variables. Therefore, the Genetic Optimization Al-
gorithm method is employed to carry out the power optimization.
The genetic optimization algorithm is designed to reliably locate
global optimal values even if the surface has many local peaks.
Two constraints are considered for the optimization, i.e. pinch
constraints and the constraint of vapor quality of working uid at
turbine exit. The latter must be greater or equal to 0.95 to avoid
droplet erosion of turbine blades [39]. Furthermore the value of
optimized variable must be inside the applicable range of working
uid properties.
7. Results and discussions
7.1. Efciency optimizations
7.1.1. First-law efciency (h
I
) optimization applied to the basic
supercritical ORC
The basic idea behind all the modications to increase the rst-
law efciency (h
I
) of a power cycle is to increase the average high
temperature (at which heat is transferred to the working uid from
the heat source medium) or/and decrease the average low tem-
perature (at which heat is rejected from the working uid to the
heat sink medium) of the cycle [58]. In practice, average high
temperature of a power cycle is limited by available inlet temper-
ature of heat source medium and pinch value of high-temperature
heat transfer process. Similarly, average low temperature of the
cycle is constrained by available inlet temperature of heat sink
medium and pinch value of low-temperature heat transfer process.
In this study, the rst-law efciency optimization was only per-
formed for the basic supercritical ORC to bring out the above idea.
Indeed, the upper limit of turbine inlet temperature was in this
Table 5
Optimum operating conditions of basic cycle in the case of efciency optimizations.
Fluid T
h
(

C) T
cond
(

C)
P
h
(bar) T
hso
(

C)
_ m
wf
(kg s
1
)
_ mc
(kg s
1
)
T
cso
(

C)
x
t,out
R134a
a
139.0 20.0 55.86 70.3 0.1402 3.099 17.2 1.12
R134a
b
139.0 24.2 53.25 71.8 0.1383 0.974 22.0 1.15
R152a
a
139.0 20.0 50.08 87.6 0.0756 2.608 17.1 1.03
R152a
b
139.0 24.2 47.74 87.8 0.0755 0.820 21.6 1.06
R32
a
139.0 20.0 93.85 57.5 0.1234 4.038 17.0 0.96
R32
b
139.0 24.4 84.2 58.6 0.1196 1.226 21.6 1.02
R744
a
139.0 20.0 197.3 46.3 0.1848 3.528 17.7 1.34
R744
b
139.0 22.9 166.9 46.7 0.1762 1.241 22.6 1.59
R1270
a
139.0 20.0 66.21 64.4 0.0809 3.375 17.2 1.11
R1270
b
139.0 24.1 62.56 65.8 0.0794 1.073 21.9 1.14
R290
a
139.0 20.0 60.49 68.8 0.0744 3.107 17.3 1.14
R290
b
139.0 24.0 57.45 70.1 0.0731 0.997 22.0 1.17
R1234yf
a
131.9 20.0 49.62 52.8 0.1965 3.568 17.4 1.20
R1234yf
b
131.9 24.0 46.08 56.7 0.1885 1.112 22.4 1.24
R1234ze(E)
a
139.0 20.0 47.2 78.4 0.1297 2.683 17.3 1.16
R1234ze(E)
b
139.0 24.1 45.12 79.7 0.1280 0.848 22.3 1.19
a
First-law efciency optimization.
b
System efciency optimization.
Table 6
Result of efciency optimizations of basic supercritical ORC.
Fluid
_
Wnet
(kW)
h
I
(%) P
net
(kW)
h
sys
(%)
h
II
(%) h
Ex
(%)
_
Itot
(kW)
_
Itot
_
Wnet
_
Ex
h;loss
(kW)
R134a
a
4.7 14.0 3.2 9.6 57.7 57.0 3.4 0.73 2.0
R134a
b
4.4 13.4 3.6 10.8 56.2 54.1 3.4 0.77 2.1
R152a
a
4.0 14.9 2.7 10.3 57.2 56.6 2.9 0.75 3.3
R152a
b
3.8 14.3 3.1 11.6 55.9 54.1 2.9 0.78 3.3
R32
a
5.3 13.5 3.4 8.6 58.8 58.1 3.7 0.70 1.2
R32
b
4.9 12.8 3.9 10.0 56.8 54.7 3.7 0.75 1.3
R744
a
4.7 10.7 2.4 5.6 49.7 48.9 4.7 1.00 0.7
R744
b
4.3 9.9 2.9 6.7 47.3 45.1 4.7 1.09 0.7
R1270
a
5.0 13.7 3.3 9.1 58.0 57.3 3.6 0.72 1.6
R1270
b
4.7 13.1 3.7 10.4 56.4 54.3 3.5 0.76 1.7
R290
a
4.8 13.8 3.2 9.3 57.3 56.6 3.5 0.74 1.9
R290
b
4.5 13.2 3.6 10.5 55.8 53.6 3.5 0.78 2.0
R1234yf
a
5.3 12.9 3.6 8.7 57.5 56.7 3.9 0.73 1.0
R1234yf
b
4.8 12.3 3.9 9.8 55.4 53.0 3.8 0.79 1.2
R1234ze(E)
a
4.3 14.1 3.0 9.8 56.0 55.4 3.3 0.78 2.6
R1234ze(E)
b
4.0 13.5 3.3 11.0 54.6 52.6 3.3 0.81 2.7
a
First-law efciency optimization.
b
System efciency optimization.
V.L. Le et al. / Energy 67 (2014) 513e526 520
work set at 139

C for all working uid except R1234yf (131.9

C,
because of its low applicable maximum temperature, i.e. 136.9

C
[66]) on account of the inlet heat source temperature (150

C) and
the pinch value (10

C) of high-temperature heat transfer process.
Similarly, the lower limit of working uid condensing temperature
was set at 20

C for all working uid in taking into account the inlet
heat sink temperature (15

C) and pinch value (3

C) of low-
temperature heat transfer process. The optimization results (cf.
Table 5) showthat optimal efciency of the cycle was achieved with
the highest possible value of turbine inlet temperature and the
lowest possible value of condensing temperature for all working
uids. Turbine inlet pressure, outlet temperature of the heat source
medium and the other operating parameters of the cycle at the
optimum efciency are encapsulated in Table 5. The magnitude of
the optimal rst-law efciencies of the cycle using various working
uids (cf. Table 6) is in the order of working uid critical temper-
atures. The highest optimal rst-law efciency (14.9%) is obtained
with R152a whose critical temperature is the most important; the
lowest optimal efciency (10.7%) is found with CO
2
, which pos-
sesses the lowest critical temperature. Because the highest possible
value of turbine inlet temperature was set at 131.9

C instead of
139

C in the case of R1234yf, the rule is not correct for this uid. By
maintaining turbine inlet and condensing temperature at highest
and lowest possible values, the rst-lawefciency variation of basic
supercritical ORC in terms of turbine inlet pressure with different
working uids is illustrated in Fig. 4.
7.1.2. System efciency (h
sys
) optimizations
The main objective of the efciency optimization scenario is to
optimize the system efciency, h
sys
, as dened in Eq. (18). The
system efciency optimization was performed with both cycle
congurations (basic and regenerative ones). In both congura-
tions, the optimal system efciency of the cycle is always reached
with the highest possible value of turbine inlet temperature (cf.
Tables 5 and 8). Turbine inlet pressure; condensing and outlet heat
source temperatures and other operating parameters at the opti-
mum efciency of the basic cycle using different working uids
were determined and reported in Table 5. The mass ow rate of
heat sink mediumin the case of systemefciency optimization was
found much lower than in the case of rst-law efciency optimi-
zation. It is also important to note that all working uids exit the
turbine at superheated states (x
t,out
> 1). This could present the
possibility to install an internal heat exchanger as a recuperator to
improve the system efciency. Like in the rst optimization sce-
nario, the order of optimal system efciencies is identical with one
of the critical temperature of working uids. The highest optimal
system efciency (11.6%) is found with R152a and the lowest one
(6.7%) is obtained with CO
2
. In changing the optimization scenario
from the rst-law efciency optimization to system efciency
optimization, mechanical power output of the system decreases
whereas electrical power output increases (cf. Table 6). The latter
can be explained by the inuence of the required power of cooling
water pump. In rst optimization scenario, this quantity was not
considered. It seems that the supplied power of cooling water
pump was just determined to guarantee that the heat rejection
took place at the lowest value of average low temperature as
Fig. 4. First-law efciency variation in terms of turbine inlet pressure in the case of
basic cycle using different working uids.
Table 7
Ranking method applied to basic cycle at maximum system efciency.
Fluid
_
Wnet h
I
P
net
h
sys
h
II
h
Ex
Total
R134a 4 2 3 2 1 1 13
R152a 8 1 6 1 1 1 18
R32 1 3 1 3 1 1 10
R744 4 8 8 8 8 8 44
R1270 2 2 2 2 1 1 10
R290 3 2 3 2 1 1 12
R1234yf 1 4 1 3 2 2 13
R1234ze(E) 6 2 5 1 2 2 18
Table 8
Operating conditions of regenerative cycle using different working uid at
maximum system efciency.
Fluid T
h
(

C) T
cond
(

C)
P
h
(bar) P
ext
(bar)
T
hso
(

C)
_ m
wf
(kg s
1
)
_ mc
(kg s
1
)
T
cso
(

C)
x
t,out
R134a 139.0 23.7 54.76 18.94 92.4 0.1393 0.775 21.4 1.13
R152a 139.0 23.9 50.26 17.86 103.7 0.0756 0.649 21.1 1.03
R32 139.0 24.0 85.19 36.90 77.9 0.1199 1.039 21.0 1.01
R1270 139.0 23.6 63.74 25.63 85.6 0.0799 0.882 21.3 1.13
R290 139.0 23.6 58.36 22.41 89.8 0.0735 0.806 21.5 1.16
R1234yf 131.9 23.6 47.75 17.22 81.7 0.1909 0.880 21.8 1.22
R1234ze(E) 139.0 23.7 46.37 15.34 99.4 0.1289 0.658 21.6 1.17
Fig. 5. System efciency variation by considering turbine inlet pressure and
condensing temperature in the case of basic cycle.
V.L. Le et al. / Energy 67 (2014) 513e526 521
mentioned in the basic idea to improve the rst-law efciency in
Section 7.1.1. In the second optimization scenario, the required
power of cooling water pump was included in the objective func-
tion. Therefore, an optimal value of condensing temperature was
found instead of the lowest possible value of this variable. The
variation of system efciency by considering the turbine inlet
pressure and condensing temperature in the case of basic cycle
using R152a is shown in Fig. 5. As shown in Table 6, the lowest total
exergy destruction is found with the cycle using R152a, whereas
the lowest ratio of total exergy destruction to mechanical power
output is obtained with the cycle using R32. The latter lowest ratio
corresponds to the highest second-law and exergy efciencies of
the cycle. The order of exergy ow rate associated with the heat
source outlet is the same as critical temperature order of working
uid.
Regarding the regenerative conguration, the optimal pressure
of bleeding point was determined and reported in Table 8. In
comparison to the basic conguration, the rst-law efciency and
optimal system efciencies of the regenerative cycle are improved
(as shown in Fig. 7) while the net power outputs (mechanical and
electrical ones) decrease (cf. Fig. 8). The second-law and exergy
efciencies of the regenerative cycle are greater than the basic
cycle (cf. Fig. 9). The other operating parameters at the optimum
system efciency of the cycle with different working media were
identied and encapsulated in Table 8. The highest optimal sys-
tem efciency (13.1% or an increase of w12.9% in comparison to
basic cycle) is always obtained with R152a (cf. Table 9). The
regenerative conguration was not used with CO
2
as working
uid because of low critical temperature of this uid. Thus, the
lowest value of optimal system efciency (11.1% or an increase of
13.3% compared with basic cycle) was found with R1234yf. In this
optimization scenario, the lowest total exergy destruction, the
lowest ratio of the total exergy destruction to mechanical power
output and the highest second-law and exergy efciencies are
found with R152a.
7.1.3. Ranking method and exergy analysis applied to the
supercritical ORCs at maximum system efciency
In this study, simple ranking method is used to make a global
comparison of thermodynamic performance of supercritical ORCs
(basic and regenerative cycles) at the optima (systemefciency and
power output). With this method, the performance indicators, i.e.
mechanical and electrical power outputs; rst-law and system ef-
ciencies; second-law and exergy efciencies, of the cycle at the
optima will be ranged from 1 (highest value) to 8 (lowest value).
The interval value of each performance indicator is calculated as:
(highest value lowest value)/7. When the values of some in-
dicators are not so different, these indicators would be classied in
a same ranking (Table 7). The sum of the order of performance
indicators is after calculated for each uid and the best cycle per-
formance would correspond to the minimum of this sum.
In the case of the basic cycle, although the highest optimal
system efciency is found with R152a, it seems that R32 and
R1270 are the best working uids for this conguration. R1234yf,
R134a and R290 also represent as potential working uids. Even
though there are many desirable environmental characteristics,
the worst thermodynamic performance is found with the cycle
using CO
2
. The operating condition of the cycle using this working
Fig. 6. Distribution of average component exergy destructions (a) and percentage of exergy destructions in the component of cycle using CO
2
(b) for system efciency optimization
of basic cycle.
Fig. 7. First-law and optimum system efciency comparison between basic and
regenerative cycle in the case of system efciency optimization.
Fig. 8. Net mechanical and electrical power output comparison between basic and
regenerative cycle in the case of system efciency optimization.
V.L. Le et al. / Energy 67 (2014) 513e526 522
medium is also severest with an operating pressure of about
200 bars.
Always in the case of basic cycle, the most destroyed exergy is
found in high-temperature heat exchanger. The exergy destruction
in the turbine is also very high. Even, with the high critical pressure
working uids (R32 and CO
2
), the destroyed exergy in the turbine is
higher than in the high-temperature heat exchanger. The distri-
bution of average ow rates of exergy destruction in the cycle
components (henceforth referred to as component exergy
destruction), calculated for all working uids excluding CO
2
, is
shown in Fig. 6a. The distribution of component exergy destruction
of the cycle using CO
2
is shown in Fig. 6b. Indeed, the distribution of
component exergy destruction is useful information to effectively
improve system exergy efciency.
By using ranking method for the regenerative cycle at maximum
systemefciency (cf. Table 10), the best cycle performance seems to
be found with R134a and followed by R152a. Although the me-
chanical and electrical power outputs of the cycle using R152a are
lowest, the other performance indicators of the cycle were found
very high with this uid. The worst thermodynamic performance of
regenerative cycle was discovered with R1234yf.
Concerning the destroyed exergy in regenerative cycle, the
distribution of average ow rate (calculated for all working uids)
of component exergy destruction is shown in Fig. 10. Not like in the
basic conguration, the most important destroyed exergy is found
in the turbine (36% of average total exergy destruction) instead of
the high-temperature heat exchanger. The percentage of average
exergy destruction in high-temperature heat exchanger represents
27% of average total exergy destruction. Indeed, a part of exergy
destruction in this device is transferred to the feed-uid heater
(14% of average total exergy destruction).
7.2. Power optimization
7.2.1. Net electrical power output optimization
Because the cycle with heat regeneration makes no power
output increase in comparison to the basic conguration, the net
electrical power optimization was just carried out with the basic
supercritical ORC. The highest value (4.1 kW) of optimal electrical
power output is found with R1234ze and the lowest one (3.0 kW) is
obtained with CO
2
(cf. Table 12). The optimal operating conditions
of this scenario are found softer than the efciency optimization
scenario but the mass ow rate of working uids and heat sink
medium are higher (Table 11). While the order of optimal system
efciency of the cycle with different working uids is identical with
the one of their critical temperatures, the order of optimal electrical
power output seems to depend on both critical and normal boiling
point temperature of the working medium. Indeed, while a lower
critical temperature is preferred for a higher optimal power output;
the normal boiling point temperature is preferred to be higher. In
comparison to optimization scenario of system efciency, the
second-law and exergy efciency of this optimization are higher.
The total destroyed exergy in the cycle is also higher but the
quantity of exergy associated with outlet heat source medium is
lower.
7.2.2. Ranking method and exergy analysis applied to the basic
supercritical cycle at maximum electrical power output
Even though the highest optimal electrical power output was
found with the cycle using R1234ze, by using ranking method
(Table 13) the best choice of working uid for this optimization
scenario seems to be R134a followed by R1234ze, R290 and R32.
Indeed, if the environmental and safety criteria are taken into ac-
count the best working uid is R1234ze. The highest system ef-
ciency of the cycle at maximum electrical power output is always
found with the cycle using R152a but the other performance in-
dicators of the cycle at maximum electrical power output are not
very high for this medium. The worst thermodynamic performance
Fig. 9. Second-law and exergy efciency comparison between basic and regenerative
cycle in the case of system efciency optimization.
Table 9
Result of system efciency optimization of regenerative supercritical ORC.
Fluid
_
Wnet
(kW)
h
I
(%) P
net
(kW)
h
sys
(%)
h
II
(%)
h
Ex
(%)
_
Itot
(kW)
_
Itot
_
Wnet
_
Ex
h;loss
(kW)
R134a 3.7 15.3 3.0 12.2 58.3 56.6 2.6 0.71 3.7
R152a 3.2 16.2 2.6 13.1 59.2 57.6 2.2 0.68 4.8
R32 4.4 14.2 3.4 11.1 57.6 55.8 3.2 0.73 2.5
R1270 4.0 14.7 3.1 11.5 57.7 55.9 2.9 0.73 3.1
R290 3.8 14.9 3.0 11.8 57.5 55.8 2.8 0.73 3.5
R1234yf 4.1 14.0 3.2 11.1 55.9 54.0 3.2 0.78 2.8
R1234ze(E) 3.3 15.4 2.7 12.5 57.5 55.8 2.4 0.73 4.4
Table 10
Ranking method applied to the regenerative cycle at maximum system efciency.
Fluid
_
Wnet h
I
P
net
h
sys
h
II
h
Ex
Total
R134a 4 3 3 3 2 2 17
R152a 7 1 7 1 1 1 18
R32 1 6 1 7 3 4 22
R1270 2 5 3 5 3 3 21
R290 3 4 3 4 4 4 22
R1234yf 2 7 2 7 7 7 32
R1234ze(E) 6 3 6 2 4 4 25
Fig. 10. Percentage of average ow rates of component exergy destruction in the case
of system efciency optimization of regenerative supercritical ORC.
V.L. Le et al. / Energy 67 (2014) 513e526 523
of the basic cycle at maximum electrical power output is always
relative to the case of CO
2
.
Regarding the exergy destruction, the highest average compo-
nent exergy destruction found in high-temperature heat exchanger
(Fig. 11a). Also with high critical pressure working uids such as
R32 and CO
2
, the highest average ow rate of destroyed exergy is
caused by high-temperature heat exchanger (Fig. 11b).
By the above result analysis, the system efciency scenario is
recommended to use when a second solution of low-temperature
recovery exist. Furthermore, the regenerative cycle is better
matched to this mission than the basic one. If there is no solution to
recover the energy of heat source outlet and to maximize the heat
recovery; power optimization seems to be more suitable.
8. Conclusions
In this study, the performance optimizations of the super-
critical ORCs (basic and regenerative congurations) are carried
out with low-GWP organic compounds as working uid. These
compounds are currently the potential candidates of the fourth
generation of refrigerants. The present work demonstrates that
critical temperature and normal boiling point temperature are
the important factor for the working uid selection. With this
study the combination of ranking method and exergy analysis
were found to be a useful approach to assess the optimal ther-
modynamic performance of the supercritical ORCs using different
working media. The optimal system efciency is considerably
improved by the incorporation of a feed-uid heater to a basic
cycle.
For each optimization scenario, the optimal operating pa-
rameters are not identical; the best working uid is not the same
for each cycle conguration. While R32 seems to be the best
choice for basic conguration in the case of the system efciency
optimization, R152a presents the best thermodynamic perfor-
mance evaluated by ranking method in the case of regenerative
cycle. The combination between heat regeneration and super-
critical ORC presents a considerable improvement of system ef-
ciency (more than 10%). In the case of power optimization and
taking into account the environmental criteria, R1234ze, with the
highest optimum electrical power output, seems to be the best
working uid. This work also recommended that the power
optimization scenario should be carried out to maximize the rate
of heat recovery but if there is a second solution to make use the
heat source outlet, the system efciency optimization should be
performed.
Some remarks can be extracted:
v The requirement of power for auxiliaries (e.g. supplied pump
powers of heat source and sink) should be considered for the
optimization.
v The system efciency gain is always paid by a loss of power
output. Therefore, multi-objective optimization scenarios
should be executed.
Table 11
Operating condition of basic cycle using different working media at maximum
electrical power output.
Fluid T
h
(

C) T
cond
(

C)
P
h
(bar) T
hso
(

C)
_ m
wf
(kg s
1
)
_ mc
(kg s
1
)
T
cso
(

C)
x
t,out
R134a 117.7 23.6 49.12 48.0 0.2133 1.616 20.6 0.98
R152a 125.8 23.7 47.02 77.7 0.0964 1.120 20.7 0.97
R32 135.2 23.5 72.91 55.1 0.1198 1.458 20.8 1.06
R744 139.0 22.7 156.00 44.9 0.1733 1.277 22.5 1.63
R1270 128.3 23.7 56.80 58.0 0.0899 1.303 21.2 1.11
R290 131.0 23.5 56.67 62.0 0.0855 1.276 21.1 1.11
R1234yf 122.5 25.3 55.06 38.6 0.2833 1.381 22.2 1.00
R1234ze(E) 115.4 24.6 39.56 44.2 0.2394 1.427 21.6 0.97
Table 12
Result of electrical power optimization of basic supercritical ORC.
Fluid
_
Wnet
(kW)
h
I
(%) P
net
(kW)
h
sys
(%)
h
II
(%) h
Ex
(%)
_
Itot
(kW)
_
Itot
_
Wnet
_
Ex
h;loss
(kW)
R134a 5.2 12.0 4.0 9.3 56.3 54.4 4.0 0.76 0.8
R152a 4.2 13.6 3.3 10.8 55.1 53.5 3.3 0.80 2.5
R32 5.0 12.4 3.9 9.7 56.0 54.1 3.9 0.77 1.1
R744 4.3 9.7 3.0 6.7 46.6 44.4 4.8 1.12 0.6
R1270 4.9 12.5 3.8 9.8 55.7 53.7 3.8 0.78 1.3
R290 4.8 12.9 3.7 10.1 56.2 54.3 3.7 0.77 1.5
R1234yf 5.1 10.9 3.9 8.3 54.4 51.9 4.2 0.82 0.4
R1234ze(E) 5.2 11.8 4.1 9.3 56.5 54.2 4.0 0.76 0.6
Table 13
Ranking method applied to the basic cycle at maximum electrical power output.
Fluid
_
Wnet h
I
P
net
h
sys
h
II
h
Ex
Total
R134a 1 3 1 3 1 1 10
R152a 8 1 6 1 1 1 18
R32 2 3 2 2 1 1 11
R744 7 8 8 8 8 8 47
R1270 3 2 2 2 1 1 11
R290 3 2 3 2 1 1 12
R1234yf 1 5 2 5 2 2 17
R1234ze(E) 1 4 1 3 1 1 11
Fig. 11. Distribution of average component exergy destruction (a) and of component exergy destruction with CO
2
(b) in case of power optimization of basic cycle.
V.L. Le et al. / Energy 67 (2014) 513e526 524
v It is recommended that the feasibility of the plant will be more
clearly brought out with thermoeconomic analysis, themoeco-
nomic optimization.
Acknowledgment
The authors thank the French National Research Agency who
has funded the work reported in this paper.
Nomenclature
Ex specic exergy (J kg
1
)
T temperature (

C)
h specic enthalpy (J kg
1
)
s specic entropy (kJ kg
1
K
1
)
MM molecular mass (kg kmol
1
)
V velocity (m s
1
)
g gravitational acceleration (m s
2
)
z altitude (m)
T
crit
critical temperature (

C)
P
crit
critical pressure (bar)
D
crit
critical density (kg m
3
)
T
h
turbine inlet temperature (

C)
P
h
turbine inlet pressure (bar)
_
I ow rate of exergy destruction (kW)
_
W mechanical power output (kW)
P electrical power output (kW)
_
Q heat transfer ow rate (kW)
_
Ex exergy ow rate (kW)
_
S
gen
ow rate of generated entropy (kW K
1
)
_ m
wf
mass ow rate of working uid (kg s
1
)
_ m
c
mass ow rate of heat sink medium (kg s
1
)
_ m
h
mass ow rate of heat source medium (kg s
1
)
x
t,out
vapor quality at turbine outlet
Greek letters
h
I
/h
II
rst-/second-law efciencies (%)
h
sys
system efciency (%)
h
Ex
exergy efciency (%)
h
motor
motor efciency (%)
h
gen
electrical generator efciency (%)
h
rev
efciency of a reversible system (%)
DH
vap
heat of vaporization
DT temperature difference
Subscripts
cond condenser/condensation
HTHEX high-temperature heat exchanger
p/p1/p2/wp pump/pump 1/pump 2/cooling water pump
t turbine
reg feed-uid heater
bp normal boiling point
in/out inlet/outlet
is isentropic
ini initial
hsi/hso heat source inlet/outlet
csi/cso heat sink inlet/outlet
tot total
h/c heat source/sink
0 dead state
References
[1] IEA. Key world energy statistics; 2012. p. 37.
[2] Ibrahim MB, Kovach RM. A Kalina cycle application for power generation.
Energy 1993;18(9):961e9.
[3] Rogdakis ED. Thermodynamic analysis, parametric study and optimum
operation of the Kalina cycle. Fuel Energy Abstr 1996;37(3):234.
[4] Nag PK, Gupta AVSSKS. Exergy analysis of the Kalina cycle. Appl Therm Eng
1998;18(6):427e39.
[5] Lolos PA, Rogdakis ED. A Kalina power cycle driven by renewable energy
sources. Energy 2009;34(4):457e64.
[6] Arslan O. Exergoeconomic evaluation of electricity generation by the medium
temperature geothermal resources, using a Kalina cycle: Simav case study. Int
J Therm Sci 2010;49(9):1866e73.
[7] Arslan O. Power generation from medium temperature geothermal resources:
ANN-basedoptimizationof Kalinacyclesystem-34. Energy2011;36(5):2528e34.
[8] Jawahar CP, Saravanan R, Bruno JC, Coronas A. Simulation studies on gax
based Kalina cycle for both power and cooling applications. Appl Therm Eng
2011;(0).
[9] Zhang X, He M, Zhang Y. A review of research on the Kalina cycle. Renew
Sustain Energy Rev 2012;16(7):5309e18.
[10] Peng S, Hong H, Jin H, Wang Z. An integrated solar thermal power system
using intercooled gas turbine and Kalina cycle. Energy 2012;(0).
[11] Sun F, Ikegami Y, Jia B. A study on Kalina solar system with an auxiliary su-
perheater. Renew Energy 2012;41(0):210e9.
[12] Yogi Goswami D. Solar thermal power technology: present status and ideas
for the future. Energy Sourc 1998;20(2):137e45.
[13] Lai NA, Fischer J. Efciencies of power ash cycles. Energy:(0).
[14] Zamrescu C, Dincer I. Thermodynamic analysis of a novel ammoniaewater
trilateral Rankine cycle. Thermochim Acta 2008;477(1e2):7e15.
[15] Quoilin S, Broek MVD, Declaye S, Dewallef P, Lemort V. Techno-economic
survey of Organic Rankine Cycle (ORC) systems. Renew Sustain Energy Rev
2013;22(0):168e86.
[16] Quoilin S, Declaye S, Tchanche BF, Lemort V. Thermo-economic optimization
of waste heat recovery Organic Rankine Cycles. Appl Therm Eng 2011;31(14e
15):2885e93.
[17] Aghahosseini S, Dincer I. Comparative performance analysis of low-
temperature Organic Rankine Cycle (ORC) using pure and zeotropic working
uids. Appl Therm Eng 2013;54(1):35e42.
[18] Tempesti D, Manfrida G, Fiaschi D. Thermodynamic analysis of two micro CHP
systems operating with geothermal and solar energy. Appl Energy
2012;97(0):609e17.
[19] Lemort V, Zoughaib A, Quoilin S. Comparison of control strategies for waste
heat recovery Organic Rankine Cycle systems. In: Sustainable thermal energy
management in the process industries international conference (Sus-
TEM2011). Newcastle 2011.
[20] Quoilin S, Lemort V, Lebrun J. Experimental study and modeling of an Organic
Rankine Cycle using scroll expander. Appl Energy 2010;87(4):1260e8.
[21] Quoilin S, Aumann R, Grill A, Schuster A, Lemort V, Spliethoff H. Dynamic
modeling and optimal control strategy of waste heat recovery organic
Rankine cycles. Appl Energy 2011;88(6):2183e90.
[22] Bombarda P, Invernizzi CM, Pietra C. Heat recovery from diesel engines: a
thermodynamic comparison between Kalina and ORC cycles. Appl Therm Eng
2010;30(2e3):212e9.
[23] Fischer J. Comparison of trilateral cycles and organic Rankine cycles. Energy
2011;36(10):6208e19.
[24] Kalra C, Becquin G, Jackson J, Laursen AL, Chen H, Myers K, et al. High-po-
tential working uids and cycle concepts for next generation binary organic
Rankine cycle for enhanced geothermal systems. In: Stanford geothermal
workshop. Stanford, Cafornia: Stanford University; 2012.
[25] Chen H, Goswami DY, Stefanakos EK. A review of thermodynamic cycles and
working uids for the conversion of low-grade heat. Renew Sustain Energy
Rev 2010;14(9):3059e67.
[26] Schuster A, Karellas S, Aumann R. Efciency optimization potential in su-
percritical organic Rankine cycles. Energy 2010;35(2):1033e9.
[27] BCS I. Waste heat recovery: technology and opportunities in U.S. industry;
2008.
[28] Viswanathan VV, Davies RW, Holbery JD. Opportunity analysis for recovering
energy from industrial waste heat and emissions; 2006. Medium: ED; Size:
PDFN.
[29] Pellegrino JL, Margolis N, Justiniano M, Miller M, Thedki A. Energy use, loss
and opportunities analysis: U.S. manufacturing & mining; 2004. p. 17.
[30] Cook E. The ow of energy in an industrial society. W.H. Freeman and Com-
pany; 1971.
[31] Blaney BL, Energy, Environmental Analysis i, Laboratory IER. Industrial waste
heat recovery and the potential for emissions reduction. National Technical
Information Service; 1984.
[32] Niget J-M. Potentiel nergtique des rejets thermiques industriels au Qubec.
Qubec: INNOVAGRO; 2011.
[33] Bao J, Zhao L. A review of working uid and expander selections for organic
Rankine cycle. Renew Sustain Energy Rev 2013;24(0):325e42.
[34] Gao H, Liu C, He C, Xu X, Wu S, Li Y. Performance analysis and working uid
selection of a supercritical organic Rankine cycle for low grade waste heat
recovery. Energies 2012;5(9):3233e47.
[35] Mikielewicz D, Mikielewicz J. A thermodynamic criterion for selection of
working uid for subcritical and supercritical domestic micro CHP. Appl
Therm Eng 2010;30(16):2357e62.
[36] Papadopoulos Athanasios I, Linke Patrick SM. On the systematic design and
selection of optimal working uids for Organic Rankine Cycles. Appl Therm
Eng 2010;30(6e7):760e9.
V.L. Le et al. / Energy 67 (2014) 513e526 525
[37] Wang EH, Zhang HG, Fan BY, Ouyang MG, Zhao Y, Mu QH. Study of working
uid selection of organic Rankine cycle (ORC) for engine waste heat recovery.
Energy 2011;36(5):3406e18.
[38] Baik Y-J, Kim M, Chang KC, Kim SJ. Power-based performance comparison
between carbon dioxide and R125 transcritical cycles for a low-grade heat
source. Appl Energy 2011;88(3):892e8.
[39] Chen H, Yogi Goswami D, Rahman MM, Stefanakos EK. Energetic and exer-
getic analysis of CO2- and R32-based transcritical Rankine cycles for low-
grade heat conversion. Appl Energy 2011;88(8):2802e8.
[40] Cayer E, Galanis N, Nesreddine H. Parametric study and optimization of a
transcritical power cycle using a low temperature source. Appl Energy
2010;87(4):1349e57.
[41] Shengjun Z, Huaixin W, Tao G. Performance comparison and parametric
optimization of subcritical Organic Rankine Cycle (ORC) and transcritical po-
wer cycle system for low-temperature geothermal power generation. Appl
Energy 2011;88(8):2740e54.
[42] Guo T, Wang H, Zhang S. Comparative analysis of natural and conventional
working uids for use in transcritical Rankine cycle using low-temperature
geothermal source. Int J Energy Res 2011;35(6):530e44.
[43] Chen H, Goswami DY, Rahman MM, Stefanakos EK. A supercritical Rankine
cycle using zeotropic mixture working uids for the conversion of low-grade
heat into power. Energy 2011;36(1):549e55.
[44] Calm JM. Refrigerant transitions...again. Conference Refrigerant Tran-
sitions...Again, Gaithersburg, MD, USA. American Society of heating refriger-
ating, and air-conditioning Engineers (ASHRAE).
[45] Calm JM. The next generation of refrigerants e historical review, consider-
ations, and outlook. Int J Refriger 2008;31(7):1123e33.
[46] CalmJM, HourahanGC. Physical, safety, andenvironmental data for current and
alternative regrigerants. In: The 23rd international congress of refrigeration
Prague. Czech Republic: International Institute of Refrigeration (IIR/IIF); 2011.
[47] Union E. Directive 2006/40/EC of the European Parliament and of the Council of
17 May 2006 relating to emissions from air-conditioning systems in motor ve-
hicles and Amending Council Directive 70/156/EEC. Off J Eur Union 2006:12e8.
[48] Kim YM, Kim CG, Favrat D. Transcritical or supercritical CO
2
cycles using both
low- and high-temperature heat sources. Energy 2012;43(1):402e15.
[49] Cayer E, Galanis N, Desilets M, Nesreddine H, Roy P. Analysis of a carbon di-
oxide transcritical power cycle using a low temperature source. Appl Energy
2009;86(7e8):1055e63.
[50] Baik Y-J, Kim M, Chang K-C, Lee Y-S, Yoon H-K. Power enhancement potential
of a mixture transcritical cycle for a low-temperature geothermal power
generation. Energy 2012;47(1):70e6.
[51] Chen Y, Pridasawas W, Lundqvist P. Dynamic simulation of a solar-driven
carbon dioxide transcritical power system for small scale combined heat
and power production. Sol Energy 2010;84(7):1103e10.
[52] Song Y, Wang J, Dai Y, Zhou E. Thermodynamic analysis of a transcritical CO
2
power cycle driven by solar energy with liquied natural gas as its heat sink.
Appl Energy 2012;92(0):194e203.
[53] Pan L, Wang H, Shi W. Performance analysis in near-critical conditions of
organic Rankine cycle. Energy 2012;37(1):281e6.
[54] Mago PJ, Chamra LM, Srinivasan K, Somayaji C. An examination of regenera-
tive organic Rankine cycles using dry uids. Appl Therm Eng 2008;28(8e9):
998e1007.
[55] Mago PJ, Srinivasan KK, Chamra LM, Somayaji C. An examination of exergy
destruction in organic Rankine cycles. Int J Energy Res 2008;32(10):926e38.
[56] Xu R-J, He Y-L. A vapor injector-based novel regenerative organic Rankine
cycle. Appl Therm Eng 2011;31(6e7):1238e43.
[57] Pei G, Li J, Ji J. Analysis of low temperature solar thermal electric generation
using regenerative organic Rankine cycle. Appl Therm Eng 2010;30(8e9):
998e1004.
[58] Yunus A, Cengel MAB. Thermodynamics an engineering approach. 5th ed.
McGraw-Hill; 2006.
[59] Feidt M, Tutica D,AB. Energy versus environment. Sci Bull 2012;74(1):117e26.
[60] Rant Z. Exergie, ein neues Wort fr technische Arbeitsfhigkeit. Forschung
auf dem Gebiete des Ingenieurswesens 1956;22(1):36e7.
[61] Micheal J, Moran HNS, Boettner Daisie D, Bailey Margaret B. Fundamentals of
engineering thermodynamics. 7th ed. Don Fowley; 2011.
[62] Kreith F, Goswami DY. The CRC handbook of mechanical engineering. 2nd ed.
CRC Press; 2005.
[63] Feidt M. Thermodynamique et optimisation nergtique des systmes et
procds. Tech. & Doc./Lavoisier; 1996.
[64] Hasan AA, Goswami DY, Vijayaraghavan S. First and second law analysis of a
new power and refrigeration thermodynamic cycle using a solar heat source.
Sol Energy 2002;73(5):385e93.
[65] Klein SA, editor. EES: engineering equation solver. Academic professional
V9.447e3D. Madison: F-Chart Software; 2013.
[66] E.W. L, M.L. H, M.O. M, editors. NIST standard reference Database 23: reference
uid thermodynamic and transport properties-REFPROP. Version 9.1. Gai-
thersburg: National Institute of Standards and Technology; 2013.
[67] Government A. Montreal protocol on substances that deplete the ozone layer.
Canberra; 2013.
V.L. Le et al. / Energy 67 (2014) 513e526 526

You might also like