You are on page 1of 13

Energy Conversion and Management xxx (2014) xxx–xxx

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Exergy analysis of zeotropic mixtures as working fluids in Organic


Rankine Cycles
S. Lecompte a,b,⇑, B. Ameel a, D. Ziviani a,b, M. van den Broek b,a, M. De Paepe a
a
Department of Flow, Heat and Combustion Mechanics, Ghent University – UGent, Sint-Pietersnieuwstraat 41, 9000 Gent, Belgium
b
Department of Industrial System and Product Design, Ghent University – UGent, Graaf Karel de Goedelaan 5, 8500 Kortrijk, Belgium

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

Article history: The thermodynamic performance of non-superheated subcritical Organic Rankine Cycles (ORCs) with
Available online xxxx zeotropic mixtures as working fluids is examined based on a second law analysis. In a previous study,
a mixture selection method based on a first law analysis was proposed. However, to assess the perfor-
Keywords: mance potential of zeotropic mixtures as working fluids the irreversibility distributions under different
Organic Rankine Cycle mixtures compositions are calculated. The zeotropic mixtures under study are: R245fa–pentane,
Zeotropic mixtures R245fa–R365mfc, isopentane–isohexane, isopentane–cyclohexane, isopentane–isohexane, isobutane–
Second law
isopentane and pentane–hexane. The second law efficiency, defined as the ratio of shaft power output
Exergy
Working fluids
and input heat carrier exergy, is used as optimization criterion. The results show that the evaporator
accounts for the highest exergy loss. Still, the best performance is achieved when the condenser heat pro-
files are matched. An increase in second law efficiency in the range of 7.1% and 14.2% is obtained com-
pared to pure working fluids. For a heat source of 150 °C, the second law efficiency of the pure fluids
is in the range of 26.7% and 29.1%. The second law efficiency in function of the heat carrier temperature
between 120 °C and 160 °C shows an almost linear behavior for all investigated mixtures. Furthermore,
between optimized ORCs with zeotropic mixtures as working fluid the difference in second law efficiency
varies less than 3 percentage points.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction Because of the simplicity of the technology and the availability


of components, the ORC has the potential to become an established
The worldwide electricity demand is still growing. Future technology in the near future. The basic ORC is already well known
projections assume an increase with 33% by 2020 and 84% by in literature and in industry. However, in order to improve the
2035 [1]. Besides growing demand, electricity prices have risen cycle performance, several modifications to the basic ORC are
in the last decade with 12%, both in Europe [2] and in the USA proposed [5]. One of these modifications is the use of zeotropic
[3]. Furthermore, environmental aspects, such as global warming, mixtures as working fluid. Zeotropic mixtures have a non-isother-
ozone layer depletion and high-levels of pollution represent a large mal phase shift. As such, they have the ability to decrease the irre-
concern and heavily influence the global energy policy. Therefore, versibility associated with heat transfer over a finite temperature
in order to reduce the dependency on fossil fuels, it is necessary difference across the working fluid and the thermal heat source
to emphasize the use of emerging and well-known renewable en- or heat sink reservoir. Zeotropic mixtures as working fluids have
ergy approaches. ORCs have the ability to convert low temperature been investigated for heat pumps [6–8] and vapor compression
heat (<300 °C) to electricity. In particular, studies show that low cycles [8–10]. The use of zeotropic mixtures as working fluids for
grade waste heat accounts for more than 50% of the total heat gen- the ORC has gained interest recently. Still, only a few studies on
erated in the industry [4]. Besides waste heat, geothermal and solar this subject have been published. Several studies focus solely on
sources are available to provide a clean alternative to fossil fuel the first law efficiency with a limited selection of mixtures: Wang
combustion. et al. [11] (R245fa–R152a), Garg et al. [12] (R245fa–isopentane)
and Borsukiewiczgozdur and Nowak [13] (propane–ethane). A
⇑ Corresponding author at: Department of Flow, Heat, and Combustion Mechan- systematic and comparative study of zeotropic mixtures as work-
ics, Ghent University – UGent, Sint-Pietersnieuwstraat 41, 9000 Gent, Belgium. ing fluids based on a first law analysis is given by the authors in
Tel.: +32 470045524; fax: +32 92643289.
a previous study [14].
E-mail address: steven.lecompte@ugent.be (S. Lecompte).

http://dx.doi.org/10.1016/j.enconman.2014.02.028
0196-8904/Ó 2014 Elsevier Ltd. All rights reserved.

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
2 S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx

Nomenclature

C mole fraction (–) WF working fluid (–)


Cp heat capacity (kJ/(kg K))
e specific exergy (kJ/kg) Greek symbols
E_ exergy flow rate (kW) g efficiency (–)
h specific enthalpy (kJ/kg)
I_ irreversibility rate (kW) Subscripts
M molar mass (g/mol)
I first law (–)
m _ mass flow rate (kg/s) II second law (–)
P pressure (bar) cond condenser (–)
PP pinch point temperature difference (°C) evap evaporator (–)
Q_ heat transfer rate (kW)
ext external (–)
s specific entropy (kJ/kg K) hc heat carrier (–)
T temperature (°C) in input stream (–)
DT hs temperature glide cooling fluid (°C) int internal (–)
T pp pinch point temperature heat carrier (°C)
N number of discretizations (–)
W_ work (kW) o dead state (–)
wf working fluid (–)
Abbreviations
HTC heat transfer coefficient (–)
ORC Organic Rankine Cycle (–)

Systematic studies which investigate the thermodynamic influ- based on a carefully selected set of chosen input and output
ence of varying mixture composition on the Organic Rankine Cycle streams. Therefore, a clear understanding of the purpose of the sys-
components are valuable but scarce. Second law analysis provides tem is needed before defining the second law efficiency [19].
the tools for this research. The use of second law analysis on Or- The objective of this study is to analyze the performance of zeo-
ganic Rankine Cycles is well-known in open literature. Mago tropic working fluids based on a second law analysis. The work
et al. [15] examined the exergy destruction in Organic Rankine Cy- focusses on non-superheated subcritical cycles. Comparisons are
cles. Visual representations using an exergy wheel clearly show the made among a selection of zeotropic working fluids. Special care
exergy accounting for each thermodynamic process. The results is taken to quantify the distribution of irreversibilities within the
show that the evaporator has by far the highest exergy destruction system. The ORC is optimized and the cause of the optimization
rate, followed by the turbine. Therefore, it seems that cycle modi- potential is analyzed.
fications, of which the aim is to reduce exergy destruction in the
evaporator, have a major potential to increase the power output
2. Cycle
of the ORC. Roy et al. [16] studied the output power, the second
and first law efficiency and irreversibilities of an ORC using R12,
2.1. Description of the ORC
R123 and R134a as working fluids. The ORC was driven by flue
gas waste heat at 140 °C. Their results show that the point of max-
The schematic diagram of the ORC analyzed in this study is gi-
imum thermal efficiency and maximum power output do not coin-
ven in Fig. 1. Fig. 1a shows the basic cycle while Fig. 1b represents
cide. Furthermore the second law efficiency is strongly affected by
the cycle with recuperator. The basic ORC consists of a pump which
the pinch point temperature difference in the evaporator. The opti-
pressurizes the working fluid and transports it to the evaporator. In
mal evaporation temperature for a subcritical ORC based on max-
the evaporator the working fluid is heated to the point of saturated
imization of the second law efficiency was investigated by Liu et al.
vapor. Next, the working fluid expands through the turbine and
[17].
produces work. This shaft power is then converted to electric
Only a few studies investigate the ORC with zeotropic mixtures
power by the generator. The superheated working fluid at the out-
from a second law perspective. Furthermore, only a limited selec-
let of the turbine is condensed to saturated liquid in the condenser.
tion of working fluids are considered. Heberle et al. [18] investi-
The liquid working fluid is again pressurized by the pump, com-
gated the second law efficiency of an ORC with zeotropic
pleting the cycle. A recuperator (Fig. 1b) cools down the super-
mixtures of isobutane–isopentane and R227ea–R245fa as working
heated vapor at the exit of the expander and heats up the
fluids. The results show that for temperatures below 120 °C the
working fluid after the pump. As a result, the load on the condenser
second law efficiencies increased in the range of 4.3–15%. The opti-
is decreased.
mal second law efficiency was achieved when the temperature
glide of condensation and cooling water matched. Ho et al. [5]
compared the Organic Flash Cycle (OFC) to an optimized basic 2.2. Working fluid selection
ORC cycle, a zeotropic rankine cycle with a binary ammonia–water
mixture and a transcritical CO2 cycle. A distinction is made be- The selected pure working fluids are all hydrocarbons and can be
tween utilization efficiency and second law internal efficiency. found in Table 1. The selection corresponds with that of a previous
The former definition assumes that the exergy which is left in study by the authors [14]. Guidelines for mixture component selec-
the waste heat stream is discarded or unused, while the latter dis- tion in cryogenic applications are adopted [20,21]. The first compo-
cards exergy destruction due to heat transfer in the evaporator. The nent is volatile at 1–1.5 bar and therefore, low temperatures after
definition of second law efficiency is therefore not unique; it is expansion can be obtained without the need to reach vacuum. The

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx 3

(a) (b)

Fig. 1. Schematic diagram of the basic ORC (a) and ORC with recuperator (b).

Table 1 [22] provides the thermophysical data of the studied mixtures and
Thermophysical properties of the selected pure working fluids. pure fluids.
Fluid M (g/mol) T crit (°C) P crit (bar) T boiling (°C)
 A discretized model is developed for the evaporator.
Cyclohexane 84.2 280.5 40.8 80.3
Hexane 86.2 234.7 30.3 68.3 ev ap ev ap ev ap ev ap
_ wf ¼ cp;hc  m
ðhwf ;x  hwf ;ðx1Þ Þ  m _ hc  ðT hc;x  T hc;ðx1Þ Þ ð1aÞ
Isobutane 58.1 134.7 36.3 12.1
ev ap ev ap ev ap ev ap
Isohexane 86.2 224.6 30.4 59.8 hwf ;1 ¼ hwf ;in ; T hc;1 ¼ T hc;out ð1bÞ
Isopentane 72.1 187.2 33.8 27.4 ev ap ev ap ev ap ev ap
Pentane 72.1 196.6 33.7 35.7 hwf ;N ¼ hwf ;out ; T hc;N ¼ T hc;in ð1cÞ
R245fa 134.0 154.0 36.5 14.8 ev ap ev ap
PP ev ap ¼ minðT hc;x  T wf ;x Þ ð1dÞ
R365mfc 148.1 186.9 32.7 39.8

Numerical simulations show that discretization with N = 15 is ade-


quate to have an accuracy of one Watt of power output.
second component exceeds both the desired average condensation  The model of the condenser is analogous to that of the
temperature and the boiling point of the first component. The final evaporator.
selection of mixtures is: isopentane–hexane, isopentane–cyclohexane,  The model of the pump and turbine are characterized by their
isopentane–isohexane, pentane–hexane, isobutane–isopentane, isentropic efficiency:
R245fa–pentane, R245fa–isopentane and R245fa–R365mfc. The isen
most volatile component is always the first component in the
hpump;out  hpump;in
_ pump ¼ m
W _ wf  ð2aÞ
mixtures name. gpump
W _ wf  gturbine ðhturbine;in  hisen
_ turbine ¼ m
turbine;out Þ ð2bÞ
3. Model and assumptions
With these equations the model is fully defined and the thermal
The ORC is evaluated assuming steady state conditions of all efficiency is calculated as:
components. Furthermore, heat loss to ambient or pressure drops
are neglected. An upper pressure limit of 90% of the critical pres- _
W
gI ¼ _ net ð3Þ
sure of the fluid is imposed in order to ensure a stable and safe Q ev ap
operation. As each of the working fluids discussed are dry fluids X
N
ev ap ev ap
and superheating is omitted. The ORC, heat source and heat sink Q_ ev ap ¼ _ wf
ðhwf ;x  hwf ;ðx1Þ Þ  m ð4Þ
parameters are summarized in Table 2. The REFPROP 9.0 database x¼1

3.1. Second law analysis


Table 2
ORC, heat source and heat sink parameters. In this work an exergy approach was chosen to investigate the
potential of zeotropix mixtures for ORCs. As such, a direct compar-
ORC
Isentropic pump efficiency gpump 60%
ison is provided to the maximum attainable power output for a
Isentropic turbine efficiency gturbine 75% reversible cycle. However, an analysis on an energy basis [23]
DT pinch evaporator PP ev ap 10 °C should give the same optimization results.
DT pinch condenser PP cond 10 °C Exergy defines the work potential of a system or component. In
Heat source order to calculate the exergy, a dead state (P 0 ; T 0 ) needs to be de-
Inlet temperature heat carrier T hc;in 150 °C fined. This is typically the ambient condition. When the system
Mass flow rate heat carrier m_ hc 20 kg/s reaches thermal and mechanical equilibrium with the given dead
Heat carrier medium hc Water
state, there is no potential of doing work. Kinetic and potential con-
Heat carrier pressure P hc 5 bar
tributions on the flow exergies were assumed to be negligible.
Heat sink
Therefore, the specific exergy e for a steady state stream is given by:
Inlet temperature cooling water T hs;in 20 °C
Outlet temperature cooling water T hs;out 30 °C
e ¼ h  h0  T 0 ðs  s0 Þ ð5Þ

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
4 S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx

By multiplying the specific exergy with the mass flow rate the  Condenser
exergy flow is obtained:
I_cond ¼ ðm
_ cf econd _ cond _ cond _ cond
cf ;in þ mwf ewf ;in Þ  ðmcf ecf ;out þ mwf ewf ;out Þ ð9Þ
E_ ¼ me
_ ð6Þ
 Recuperator
A qualitative exergy flow diagram for the ORC with recuperator
ev ap
is given in Fig. 2. The external exergy flow E_ ehc;in v ap
that enters the _ wf ½ðepump
I_recup ¼ m turbine cond
wf ;out þ ewf ;out Þ  ðewf ;in þ ewf ;in Þ ð10Þ
system amounts to the maximum attainable power production.
 Pump
Furthermore, there is an internal exergy flow E_ ewfv ap ;in
to the evapora-
tor. The irreversibilities in the heat exchangers related to finite I_pump ¼ W _ wf ðepump
_ pump  m pump
wf ;out  ewf ;in Þ ð11Þ
temperature heat transfer are given as I_ev ap ; I_cond and I_recup . There
are no internal irreversibilities associated to pressure losses in  Turbine
the heat exchangers. The exergy flow E_ ewfv ap enters the turbine,
;out
_ turbine is produced and an exergy flow E_ turbine leaves. Part I_turbine ¼ m
_ wf ðeturbine turbine _
wf ;in  ewf ;out Þ  W turbine ð12Þ
work W wf ;out
of this exergy goes to the recuperator, while the remaining part
E_ cond
wf ;in goes to the condenser. The cooling water that enters the con- In order to make a fair comparison between cycle designs, a
denser has an exergy flow E_ cond _ cond
cf ;in and an exergy flow Ecf ;out is rejected consistent thermodynamic performance criterion is necessary. This
to the ambient. Irreversibilities in the pump and turbine are criterion should in essence be an indicator of the availability of the
respectively indicated as I_pump and I_turbine . studied thermodynamic cycle to convert waste heat to work. This
The irreversibility rates of the components for a steady state is accomplished by defining efficiencies based on exergy. DiPippo
flow are calculated as follows: [19] states that the efficiency based on the second law of thermo-
X 
T0 _ X X dynamics is the best comparison criteria. The framework of exergy
0¼ 1 _ þ
Qj  W _ in ein 
m _ out eout  I_
m ð7Þ
Tj analysis is extensively discussed in the work of among others Bejan
j in out
et al. [24], Moran and Shapiro [25] and Winterbone [26]. The
This gives for the components: second law efficiency gII is given as:
_
W
 Evaporator gII ¼ _ evnet ð13Þ
E ap hc;in
ev ap ev ap
I_ev ap ¼ ðm
_ hc ehc;in _ wf ewf
þm _ ev ap _ ev ap
;in Þ  ðmhc ehc;out þ mwf ewf ;out Þ ð8Þ
ev ap ev ap ev ap
E_ hc;in _ hc  ðhhc;in
¼m  h0  T 0 ðshc;in  s0 ÞÞ ð14Þ
v ap
With E_ ehc;in
the exergy flow to the system. The second law effi-
ciency is further decomposed in a second law external efficiency
(gII;ext ) and a second law internal efficiency (gII;int ). The former is
related to the heat input to the cycle and the latter is related to
the irreversibilities in the ORC.

gII ¼ gII;ext  gII;int ð15Þ


ev ap ev ap
E_ hc;in  E_ hc;out
gII;ext ¼ ð16Þ
E_ ev ap
hc;in
_
W
gII;int ¼ _ ev ap net_ ev ap ð17Þ
E E
hc;in hc;out

The exergy destruction ratio yD is introduced to compare rela-


tive component irreversibilities in case of different exergy inputs:

I_component
yD;component ¼ ð18Þ
_Eev ap  E_ ev ap
hc;in hc;out

This factor gives the percentage of exergy input to the cycle that
is destroyed in a component due to irreversibilities.
A comparative analysis is made by evaluating the exergy
destruction ratio to reference values of a pure working fluid. The
most volatile component in the mixture is chosen as working fluid
for the reference case.

DyD;component ¼ yD;component  yref


D;component ð19Þ

The sum of DyD;component over the different components is a mea-


sure of the change in irreversibilities between a thermodynamic
cycle with a zeotropic mixture as working fluid and a pure working
fluid.

3.2. Optimization

Throughout this work the second law efficiency gII is


Fig. 2. Exergy flow diagram of the ORC cycle with recuperator. maximized, unless otherwise stated. The optimization variables

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx 5

are the evaporating pressure and the mixture concentration. If the temperature glide in the condenser which lead to a decrease in
mixture concentration is fixed, only the evaporation pressure is working fluid outlet temperature at the condenser. Also, the pinch
optimized. Optimization is done with a Generalized Reduced Gra- point temperature is reduced, lowering the heat input limitation
ev ap
dient (GRG) nonlinear multistart algorithm [27]. Since E_ hc;in is con- induced by the pinch point temperature difference. This is the fur-
stant for a given heat source, optimization of the second law ther is discussed in Section 4.1.2. Furthermore, an increase in heat
efficiency is equal to optimization of the net power output. input follows when the latent heat capacity rate of the mixture is
reduced [28]. This shifts the pinch point to the right in a QT dia-
gram allowing a larger cooling of the heat carrier.
4. Results
4.1.2. Analysis of the optimal pinch point temperature
4.1. Effect of zeotropic mixtures
In this section, the difference between the optimal pinch point
temperature for pure fluids and zeotropic mixtures is discussed.
In this section R245fa–isopentane is used as a reference case to
The pinch point temperature T pp in this work, is defined as the tem-
illustrate the different effects of zeotropic mixtures. Deviations and par-
perature of the heat carrier at the point of minimum temperature
ticularities for other mixtures (R245fa–pentane, R245fa–R365mfc,
difference between working fluid and heat carrier. The evaporation
isopentane–isohexane, isopentane–cyclohexane, isopentane–
pressure unambiguously determines the pinch point temperature
isohexane, isobutane–isopentane, pentane–hexane) are discussed
and vice versa. A cycle with R245–isopentane is constructed with
in Section 4.2.
as reference T pp the optimal T pp of an ORC with pure R245fa
In Fig. 3c, gII is plotted in function of the mixture concentration.
(103.8 °C). Only the second optimization variable, the concentra-
An optimal mixture concentration of 0.28/0.72 gives an increase of
tion, is optimized in this cycle.
8.45% in second law efficiency compared to the pure working fluid
Fig. 8 gives the QT diagram for this suboptimal reference cycle.
R245fa. Both are evaluated at their optimal evaporation pressure.
The cycle with pure R245fa is shown as a full line. Again, the heat
The source of this increase and the underlying optimization strat-
input to the cycle is increased due to matching temperature pro-
egy are analyzed further.
files in the condenser and the location of the pinch point. However,
the heat input is lower than with an optimal pinch point temper-
4.1.1. Internal and external second law efficiency ature (see Fig. 7). The discussion below explains what happens if
In a first step the absorbed heat and heat conversion to power the pinch point temperature is optimized.
are decoupled by plotting gII;ext and gII;int (see Fig. 4c) for R245fa– From Fig. 9 it is clear that, between mixture concentrations 0 to
isopentane. By using a mixture as working fluid more heat is 0.7 the optimal pinch point temperature is larger than that of the
transferred to the ORC (higher gII;ext ) and heat is converted more reference case. The results for external and internal second law
effectively to power (higher gII;int ). Furthermore, for R245fa– efficiency for the reference case and the optimal case is given in
isopentane, the optimal concentration which maximizes gII;ext and Fig. 10. Between mixture concentrations 0–0.7 the internal second
gII;int is equal and coincides with that of maximum second law law efficiency gII;int is larger for the reference cycle while the exter-
efficiency gII . nal second law efficiency gII;ext is lower. For the cycle with optimal
An increase in gII;int is the result of a lower overall exergy pinch point temperature, internal second law efficiency is thus sac-
destruction in the cycle. The source of the reduced exergy destruc- rificed for increased heat input. Mixture concentrations from 0.7 to
tion is investigated by calculating DyD for all components. For 1 are optimized by increasing the pinch point temperature. Inter-
R245fa–isopentane DyD;cond and DyD;ev ap are given in Fig. 5c while nal second law efficiency is increased in detriment of a decreased
DyD;turbine and DyD;pump are given in Fig. 6c. Note that the values of heat input. By optimizing the evaporating pressure and thus the
DyD;cond are generally larger than that of DyD;ev ap ; DyD;turbine or pinch point temperature an optimal balance is found between
DyD;pump . Especially the pump has a low effect on the cycle internal gII;int and gII;ext maximizing the internal efficiency gII .
performance when varying mixture concentration. Furthermore it
is important to notice that the minimum value of DyD;cond is found 4.1.3. Glide slopes
at the same concentration that maximizes gII;int . For the mixture R245fa–isopentane the optimal concentration
Fig. 7 gives the QT diagram for the optimized ORC with pure (0.28/0.72) corresponds to the maximum glide slope possible in
working fluid (R245fa, full line) and with mixture (R245fa–isopen- the condenser and evaporator (see Fig. 11). As discussed earlier,
tane, dashed line). The heat carrier is cooled down to a lower tem- matching the glide slope of the working fluid in the condenser to
perature when using a mixture. Firstly, this is due to the matching the glide slope of cooling fluid increases the heat input (gII;ext )

(a) (b) (c)

Fig. 3. Second law efficiency of zeotropic mixtures in function of mixture concentration (optimal evaporation pressure), (a) isopentane-isohexane, isopentane-cyclohexane,
isopentane-hexane, (b) isobutane-isopentane, pentane-hexane, (c) R245fa-pentane, R245fa-isopentane, R245fa-R365mfc.

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
6 S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx

(a) (b) (c)

Fig. 4. Internal gII;int and external gII;ext second law efficiency of zeotropic mixtures in function of mixture concentration (optimal evaporation pressure), (a) isopentane-
isohexane, isopentane-cyclohexane, isopentane-hexane, (b) isobutane-isopentane, pentane-hexane, (c) R245fa-pentane, R245fa-isopentane, R245fa-R365mfc.

(a) (b) (c)

Fig. 5. DyD;cond and DyD;ev ap for zeotropic mixtures in function of mixture concentration (optimal evaporation pressure), (a) isopentane-isohexane, isopentane-cyclohexane,
isopentane-hexane, (b) isobutane-isopentane, pentane-hexane, (c) R245fa-pentane, R245fa-isopentane, R245fa-R365mfc.

(a) (b) (c)

Fig. 6. DyD;turbine and DyD;pump for zeotropic mixtures in function of mixture concentration (optimal evaporation pressure), (a) isopentane-isohexane, isopentane-cyclohexane,
isopentane-hexane, (b) isobutane-isopentane, pentane-hexane, (c) R245fa-pentane, R245fa-isopentane, R245fa-R365mfc.

but also contributes considerably to the increase of conversion effi- C without recuperator ¼ 0:28=0:72). However, compared to the ORC with-
ciency gII;int . As a result matching the glide slopes in the condenser out recuperator, the internal second law efficiency (gII;int ) is higher
is crucial when optimizing the power output. while the external second law efficiency (gII;ext ) and second law
efficiency (gII ) are lower. This is illustrated in Fig. 12 where the rel-
4.1.4. Effect of the recuperator ative change of gII;int ; gII;ext , and gII is shown in function of the mix-
In this section, the influence of adding a recuperator to the basic ture concentration. An increase in gII;int implies that the conversion
ORC is analyzed. The optimal pressure and concentration to maxi- from exergy to electric power occurs more efficiently. In contrast, a
mize the net power output are approximately equal for an ORC decrease in gII;ext implies that the heat input to the ORC is reduced.
with and without recuperator (P recuperator
ev ap ¼ 8:47 bar, The second law efficiency gII is reduced, since the decrease in heat
recuperator without recuperator
C ¼ 0:29=0:71 versus P ev ap ¼ 8:43 bar, input is slightly larger than the increase in conversion efficiency.

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx 7

Fig. 7. QT diagram R245fa–isopentane (optimal concentration and evaporation pressure) and pure working fluid R245fa (optimal pressure).

Fig. 8. QT diagram R245fa–isopentane (optimal concentration, fixed pinch point temperature 103.8 °C) and pure working fluid R245fa (optimal pressure).

Fig. 10. gII;ext and gII;int in function of mixture concentration for optimal and
Fig. 9. Pinch point temperature and pressure in function of mixture concentration
reference pinch point temperature (=103.8 °C) (R245fa–isopentane).
with optimal and reference pinch point temperature (=103.8 °C) (R245fa–
isopentane).

and outlet of the condenser increases. As a result, the logarithmic


In Fig. 12, the higher temperature glide of the working fluid (the mean temperature difference increases, leading to a higher heat
glide slope peaks at a concentration of 0.2/0.8) result in a larger de- recuperation and a decreased heat input at the evaporator.
crease of gII . This is due to the imposed pinch point temperature The cause of the improved gII;int is further explained by plotting
difference at the recuperator. When the temperature glide gets ywithout recuperator
D;component  ywith recuperator
D;component in function of mixture concentration,
higher, the temperature difference between the exit of the turbine as is done in Fig. 13. This shows that the irreversibilities in the

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
8 S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx

case in geothermal, combined heat and power, or in waste heat


applications when condensation of the flue gas should be avoided.
Furthermore, the recuperator adds an extra pressure drop, which is
not taken into account in this study, but will further decrease the
second law efficiency. In addition, the investment cost of the ORC
is increased by adding a recuperator. Therefore, no recuperator is
added to the ORC in the following discussions.

4.2. Comparison of mixture compositions

Besides R245fa–isopentane other mixtures are investigated:


R245fa–pentane, R245fa–R365mfc, isopentane–isohexane, isopen-
tane–cyclohexane, isopentane–isohexane, isobutane–isopentane,
pentane–hexane. With the above discussion in mind the specifics
Fig. 11. Glide slope of R245fa–isopentane in evaporator and condenser in function of these mixtures are examined.
of concentration. For the mixtures isopentane–cyclohexane, isopentane–hexane
and isobutane–isopentane two local maxima are observed for gII
in function of the mixture concentration (Fig. 3). These maxima oc-
cur when the temperature glide of the mixture condensation ex-
ceeds the temperature glide of the cooling water. This leads to a
decrease in gII;int between these two maxima (Fig. 4) due to an in-
crease in irreversibilities in the condenser (DyD;cond , Fig. 5). The
external second law efficiency is constant or increases and thus
does not contribute to the creation of the local maxima. There is
however only one global maximum. This is primarily accounted
to the variation in pump irreversibilities, as can be noticed in
Fig. 6. High pump irreversibilities are found for large pumping
powers. The large pumping power is a result of a high mass flow
rate or a high pressure difference over the pump.
Especially for isobutane–isopentane the irreversibilities in the
pump have a noticeable effect when changing mixture concentra-
tion (Fig. 6b). The cause is the large decrease of pressure difference
Fig. 12. Relative difference in % between ORC with and without recuperator in over the pump: from 14.6 bar for pure isobutane to 4.62 bar for
function of mixture concentration for: gII ; gII;int ; gII;ext (optimal evaporation pressure,
pure isopentane. For the other mixtures, the pressure difference
R245fa–isopentane).
over the pump is only around 5 bar or lower. Furthermore, for
the mixture isobutane–isopentane, there is a large increase in
DyD;ev ap when going to higher concentrations of isopentane
(Fig. 5b). This is explained by the optimization strategy. In order
to maximize the second law efficiency, the pinch point tempera-
ture is strongly reduced (12 °C in comparison to pure isobutane).
This in turn increases the irreversibilities in the evaporator.

4.3. Optimal mixture composition

In Table 3 the results of the optimization of pressure and mix-


ture concentrations are summarized. It is clear that the use of zeo-
tropic mixtures increases gII . For working fluids with optimal
mixture concentration and optimal pressure the variation in sec-
ond law efficiency is low (between 30.94% and 32.05%). The best
Fig. 13. ywithoutrecuperator
D;component  ywithrecuperator
D;component in function of mixture concentration for: result (gII = 32.05%) is found for a mixture of isobutane–isopentane
evaporator, condenser, recuperator (optimal evaporation pressure, R245fa– with a concentration of 0.81/0.19. Both the internal and external
isopentane). second law efficiencies are higher for the ORC with mixtures as
working fluid. The optimal condensing glide slope is slightly smal-
ler than the glide slope of the cooling fluid. This is a result of the
evaporator and condenser are lower for an ORC with a recuperator. change in DyD for the evaporator, pump and turbine when chang-
This agrees with the general interpretation that the average heat ing mixture composition (Figs. 5 and 6). The optimal evaporation
rejection temperature is lowered and the average heat addition pressures for the mixtures are lower than those of the most volatile
temperature is increased by the incorporation of a recuperator. component in the mixture. This is favorable in view of construction
As a result, the working fluid better matches the temperature pro- and operation.
files of the cooling and heating medium, reducing the irreversibil- The relative improvement in second law efficiency is within
ities at the condenser and evaporator. 7.1% to 14.2% and is presented in Fig. 14. The increase of perfor-
For other mixtures as working fluid, the effect is similar. The mance between pure working fluid and mixture is the largest for
second law efficiency of an ORC with recuperator closely approxi- isopentane when mixing with isobutane (an increase of 14.2%).
mates the second law efficiency of the ORC without recuperator. R245fa is best mixed with pentane to maximize the power output.
Only when the cooling of the heat carrier is constrained or the heat The working fluid which is the least volatile in the mixture benefits
after the ORC is used, a recuperator can be beneficial. This is the most from adding a more volatile component.

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx 9

Table 3
Characteristic cycle parameters for heat carrier at 150 °C (optimal evaporation pressure and concentration).

Working fluid C [mole frac.] gI (%) gII gII;ext gII;int P ev ap P cond Slope evap Slope cond m_ wf W_ turbine W_ pump
(%) (%) (%) (bar) (bar) (°C) (°C) (kg/s) (kW) (kW)
Isopentane 1 9.05 28.07 75.92 36.98 6.08 1.46 0.0 0.0 46.5 1810 59.6
Hexane 1 8.94 27.65 74.57 37.08 1.95 0.36 0.0 0.0 42.5 1741 17.6
Cyclohexane 1 9.14 26.71 71.39 37.42 1.35 0.24 0.0 0.0 39.1 1675 9.5
Pentane 1 9.05 27.84 75.11 37.07 4.91 1.12 0.0 0.0 43.5 1781 45.4
Isohexane 1 8.91 27.87 75.36 36.98 2.48 0.48 0.0 0.0 44.9 1761 23.6
R245fa 1 9.39 28.67 76.31 37.58 11.29 2.43 0.0 0.0 90.8 1888 99.7
R365mfc 1 8.95 28.39 76.67 37.03 4.85 0.96 0.0 0.0 82.8 1814 43.7
Isobutane 1 10.04 29.14 77.20 37.74 19.78 5.19 0.0 0.0 46.5 1810 239.4
Isopentane–hexane 0.19–0.81 9.09 30.94 79.70 38.82 2.41 0.42 7.0 8.8 44.8 1952 23.2
Isopentane–cyclohexane 0.08–0.92 9.27 30.06 76.88 39.11 1.56 0.26 7.4 9.3 41.7 1886 11.9
Isopentane–isohexane 0.44–0.56 9.22 31.63 80.52 39.29 3.75 0.72 7.1 8.7 47.5 2011 38.3
Pentane–hexane 0.54–0.46 9.32 31.39 79.30 39.59 3.19 0.57 7.0 8.3 44.3 1988 30.6
Isobutane–isopentane 0.81–0.19 9.61 32.05 81.91 39.13 13.48 3.36 6.9 9.3 52.0 2154 156.0
R245fa–pentane 0.33–0.67 9.46 31.76 80.27 39.57 7.35 1.50 8.0 7.5 60.9 2053 72.8
R245fa–isopentane 0.28–0.72 9.38 31.11 80.01 38.88 8.43 1.85 6.4 5.7 61.9 2027 87.6
R245fa–R365mfc 0.41–0.59 9.29 31.39 80.46 39.01 7.20 1.37 5.7 6.4 88.3 2025 68.0

The irreversibility distributions are given for zeotropic mixtures dominant, this results in an increased pinch point temperature dif-
as working fluid with optimal mixture concentration and pressure ference for a fixed heat exchanger area and vice versa. Therefore, in
(Fig. 15) and for pure working fluids with optimal pressure a cost benefit analysis a thermo-economic approach is crucial. To
(Fig. 16). The evaporator contributes the most to the losses in the have a first assessment, the sensitivity of the pinch point temper-
ORC (around 50%) followed by the condenser (around 30%), turbine ature difference on the second law efficiency is studied. The pinch
(around 19%) and pump (around 1%). For pure isobutane, but also point temperature difference in both the evaporator and condenser
for the mixture isobutane–isopentane, the pump takes a relatively are incrementally increased from 10 °C tot 13 °C.
large share of the total irreversibilities compared to other working In Fig. 18 the results of this analysis is given for the different
fluids. The reason and effect of this was discussed in Section 4.2. mixtures under consideration. The decrease in gII in the given
Important to notice is the shift in irreversibility distribution be- range is approximately linear. For a decrease of 1 °C in both the
tween the pure working fluid and a mixture as working fluid. As condenser and evaporator, an absolute reduction of around 1 per-
discussed earlier, the improved performance is the result of a centage point in second law efficiency is observed. It is interesting
reduction in irreversibilities in the condenser. The result of the to note, that in general a pinch point temperature increase of 3 °C
optimization therefore reduces the share of condenser losses rela- in the condenser and evaporator, already outweighs the observed
tive to the total losses in the cycle. For the current working fluids performance benefit associated to zeotropic mixtures. Further-
the reduction lies between 3 and 6 percentage points. more, the ranking in terms of gII stays the same for each calculated
pinch point temperature difference.
4.4. Sensitivity studies

4.4.1. Sensitivity on inlet heat carrier temperature


In a next step, the inlet temperature of the heat carrier is varied 4.4.3. Sensitivity on cooling fluid glide slope
between 120 °C and 160 °C, while optimizing the pressure and The sensitivity with respect to the temperature glide of the
concentration in order to maximize the net output power. In cooling fluid DT hs is evaluated by performing calculations for a
Fig. 17 the resulting second law efficiency is shown for the various variety of temperature glides, i.e. 5, 10 and 15 °C. These correspond
mixtures investigated. The second law efficiency in function of the with a cooling fluid outlet temperature of 25, 30, 35 °C. In Table 4
inlet temperature shows an almost linear behavior for all the the results of the sensitivity analysis on the cooling fluid tempera-
investigated mixtures. The difference between the best and worst ture glide are given. The working fluid with the highest second law
performing working fluid is between 1 to 3 percentage points. efficiency is R245fa–R365mfc for DT hs ¼ 5 °C, isobutane–isopen-
Therefore in order to further differentiate the best working fluid tane for DT hs ¼ 10 °C and isopentane–hexane for DT hs ¼ 15 °C.
thermo-economic considerations [29] should be taken into ac- Mixtures with R245fa have the lowest glide slope for the work-
count. A mixture of isopentane–isobutane has the highest second ing fluid. They provide the highest second law efficiency for a
law efficiency in the range 130 °C to 160 °C. At 120 °C a mixture DT hs ¼ 5 °C. For larger DT hs these mixtures are not able to achieve
of isopentane–isohexane has the highest second law efficiency. a good match with the cooling fluid. Mixtures with R245fa are
For increasing heat carrier inlet temperatures the difference be- therefore preferred for low DT hs . Isopentane–cyclohexane has as
tween second law efficiency of isopentane–isohexane and isopen- only working fluid a good match for DT hs ¼ 15 °C. Still, isopen-
tane–isohexane increases. In contrast to the optimal evaporation tane–hexane has a larger second law efficiency due to a better tem-
pressure, the optimal concentration only changes slightly from perature glide match in the evaporator associated to a lower latent
the values given in Table 3. Again, this indicates that primarily heat of the mixture.
the condenser operating conditions affect the use of zeotropic In contrast to a change in pinch point temperature difference or
mixtures. heat carrier inlet temperature, the glide slope of the cooling fluid
has a decisive effect on the selection of the optimal working fluid.
4.4.2. Sensitivity on pinch point temperature difference This supports the importance of matching the cooling fluid glide
The convective heat transfer coefficient (HTC) of zeotropic mix- slope in the condenser. For DT hs ¼ 5 °C, all the mixtures provide
tures is usually lower compared to pure fluids. An extensive over- a good temperature glide match. However, for larger DT hs not all
view of the degradation effects on the HTC is found in the work of the mixtures achieve a sufficient large glide slope in the condenser,
Rajapaksha [8]. If the convective HTC on the working fluid side is which is reflected in a lower second law efficiency.

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
10 S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx

4.5. Comparison to other work

Some noteworthy remarks from literature related to the pre-


sented results are listed in this section.
First of all, Heberle et al. [18] investigated mixtures of R245fa–
R227ea and isobutane–isopentane as working fluids for geother-
mal ORCs. Their results showed that for temperatures below
120 °C the second law efficiency increased in the range of 4.3–
15% by using zeotropic mixtures. In this work, more fluids were
investigated and a comparable improvement of second law effi-
ciency is found for a heat carrier of 150 °C. In addition, the impor-
tance of matching the temperature glide at condensation with the
temperature difference of the cooling water is supported.
Garg et al. [12] investigated mixtures of isopentane and R245fa. A
mixture concentration of 30% R245fa and 70% isopentane was specifi-
cally chosen to suppress flammability of isopentane. In this study,
Fig. 14. Relative improvement when using mixtures compared to the pure working approximately the same mixture concentration leads to a maximiza-
fluids (optimal evaporation pressure and concentration). tion of the second law efficiency under the conditions given in Table 2.

Fig. 15. Irreversibility distribution for an ORC with zeotropic mixtures as working fluid (optimal evaporation pressure and concentration).

Fig. 16. Irreversibility distribution for an ORC with pure working fluids (optimal evaporation pressure).

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx 11

Fig. 17. Second law efficiency gII in function of variable heat carrier temperature between 120 °C and 160 °C (optimal evaporation pressure and concentration).

Fig. 18. Second law efficiency gII in function of variable pinch point temperature difference between 10 °C and 13 °C) (optimal evaporation pressure and concentration).

Table 4
Results of the sensitivity analysis on the cooling fluid temperature glide DT hs : second law efficiency and working fluid temperature glide (optimal evaporation pressure and
concentration).

Working fluid Temperature glide cooling fluid condenser DT hs


5 °C 10 °C 15 °C
gII Slope WF gII Slope WF gII Slope WF
cond (° C) cond (° C) cond (° C)
Isopentane–hexane 32.08 4.4 30.94 8.8 30.43 12.8
Isopentane–cyclohexane 32.30 5.6 30.06 9.3 29.73 18.4
Isopentane–isohexane 32.65 5.4 31.63 8.7 28.70 8.7
Pentane–hexane 32.45 5.4 31.39 8.3 28.44 8.3
Isobutane–isopentane 33.71 4.8 32.05 9.3 30.37 11.8
R245fa–pentane 33.41 5.3 31.76 7.5 28.80 7.6
R245fa–isopentane 33.56 4.7 31.11 5.7 28.18 5.8
R245fa–R365mfc 33.59 4.7 31.39 6.4 28.39 6.4

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
12 S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx

4.5.1. Irreversibility distribution Acknowledgements


Several authors report a similar irreversibility distribution as in
this work. Heberle et al. [18] indicates values around 25% for the The results presented in this paper have been obtained within
share in condenser exergy losses, dependable on the mixture con- the frame of the IWT SBO-110006 project The Next Generation Or-
centration of isobutane–isopentane. Zhu et al. [30] compares ganic Rankine Cycles (www.orcnext.be), funded by the Institute for
water, ethanol, R113, R123 and R245fa as working fluids in an the Promotion and Innovation by Science and Technology in Flan-
ORC for engine exhaust heat recovery. Their values for the share ders. This financial support is gratefully acknowledged.
in condenser exergy losses range between 20% and 30%. El-Emama
and Dincer [31] provide an exergoeconomic analyses of a geother- References
mal ORC. For an optimized ORC configuration with a heat carrier
inlet temperature of 160 °C, the condenser accounts for 30.68% in [1] U.S. Energy Information Administration, International energy outlook 2011,
Washington, DC; 2011.
the total exergy losses. Shengjun et al. [32] compared subcritical [2] Eurostat, Electricity prices for household consumers; 2013. <http://
and transcritical ORCs for geothermal applications. Their results www.epp.eurostat.ec.europa.eu>.
indicate that the share in condenser exergy losses lies between [3] U.S Energy Information Administration, Average retail price of electricity to
ultimate customers. <http://www.eia.gov>.
18.6% and 23.2% for subcritical ORCs. In the works cited above, dif-
[4] Hung T, Shai T, Wang S. A review of organic rankine cycles (ORCs) for the
ferent boundary conditions are used, yet all of the values fall in a recovery of low-grade waste heat. Energy 1997;22:661–7.
range between 18.6% and 30%. [5] Ho T, Mao SS, Greif R. Comparison of the organic flash cycle (OFC) to other
However, some studies report a lower share of condenser exer- advanced vapor cycles for intermediate and high temperature waste heat
reclamation and solar thermal energy. Energy 2012;42(1):213–23. http://
gy destruction (between 2% and 5%). This is attributed to two rea- dx.doi.org/10.1016/j.energy.2012.03.067.
sons. Firstly, the exergy destruction in the condenser is low if an [6] Yilmaz M. Performance analysis of a vapor compression heat pump using and
isothermal heat sink is used whose temperature is close to the con- zeotropic refrigerant mixtures. Energy Convers Manage 2003;44:267–82.
[7] Karagoz S, Yilmaz M, Comakli O, Ozyurt O. R134a and various mixtures of R22/
densation temperature of the working fluid [33]. However, this im- R134a as an alternative to R22 in vapour compression heat pumps. Energy
plies an infinite cooling fluid mass flow rate. Secondly, some papers Convers Manage 2004;45(2):181–96. http://dx.doi.org/10.1016/S0196-
impose unrealistic temperature profiles in the condenser. In the 8904(03).
[8] Rajapaksha L. Influence of special attributes of zeotropic refrigerant mixtures
work of Tchanche et al. [34] the cooling fluid (water) exits the on design and operation of vapour compression refrigeration and heat pump
condenser at a higher temperature than the working fluid inlet systems. Energy Convers Manage 2007;48(2):539–45. http://dx.doi.org/
temperature. Also in the work of Mago et al. [35] the temperature 10.1016/j.enconman.2006.06.001.
[9] Arcaklioğlu E, Çavusßğlu A, Erisßen A. An algorithmic approach towards finding
profiles cross in the condenser. As a result, exergy is locally gener- better refrigerant substitutes of cfcs in terms of the second law of
ated in the condenser, in violation of the second law. As a conse- thermodynamics. Energy Conversion and Management
quence, the overall exergy destruction in the condenser is 2005;46(Turkey):1595–611. http://dx.doi.org/10.1016/
j.enconman.2004.07.012.
unrealistically low.
[10] Jin X, Zhang X. A new evaluation method for zeotropic refrigerant mixtures
based on the variance of the temperature difference between the refrigerant
and heat transfer fluid. Energy Convers Manage 2011;52(1):243–9. http://
5. Conclusions dx.doi.org/10.1016/j.enconman.2010.06.062. URLhttp://dx.doi.org/10.1016/
j.enconman.2010.06.062.
[11] Wang J, Zhao L, Wang X. A comparative study of pure and zeotropic mixtures
Using mixtures in subcritical ORCs as working fluids has been in low-temperature solar rankine cycle. Appl Energy 2010;87(11):3366–73.
studied in this paper by using second law analysis. This result in http://dx.doi.org/10.1016/j.apenergy.2010.05.016.
[12] Garg P, Kumar P, Srinivasan K, D.P., isopentane Evaluation of. R-245fa and their
three interesting conclusions. mixtures as working fluids for organic rankine cycles. Appl Therm Eng
Using mixtures results in an improvement of second law effi- 2013;51:292–300.
ciency between 7.1% and 14.2% compared to pure working fluids. [13] Borsukiewiczgozdur A, Nowak W. Comparative analysis of natural and
synthetic refrigerants in application to low temperature clausius rankine
The source of this improvement lies in a combination of a higher heat cycle. Energy 2007;32(4):344–52. http://dx.doi.org/10.1016/
input (gII;ext ) and a higher heat conversion efficiency (gII;int ) and is j.energy.2006.07.012.
mainly ascribed to decreased irreversibilities in the condenser. [14] Chys M, van den Broek M, Vanslambrouck B, De Paepe M. Potential of
zeotropic mixtures as working fluids in organic rankine cycles. Energy
The distribution of irreversibilities has changed and the share of
2012;44(1):623–32. http://dx.doi.org/10.1016/j.energy.2012.05.030.
condenser losses is reduced with 3 to 6 percentage points relative [15] Mago PJ, Srinivasan KK, Chamra LM, Somayaji C. An examination of exergy
to the pure working fluid. Exergy losses in the condenser are re- destruction in organic rankine cycles. Int J Energy Res 2008;32(10):926–38.
duced by matching the glide slope of the working fluid with the http://dx.doi.org/10.1002/er.1406.
[16] Roy J, Mishra M, Misra A. Parametric optimization and performance analysis of
glide slope of the cooling fluid. Consequently, an optimal mixture a waste heat recovery system using organic rankine cycle. Energy
concentration exists which maximizes the second law efficiency. 2010;35(12):5049–62. http://dx.doi.org/10.1016/j.energy.2010.08.013.
The optimal condensing glide slope is slightly smaller than the [17] Liu C, He C, Gao H, Xu X, Xu J. The optimal evaporation temperature of
subcritical ORC based on second law efficiency for waste heat recovery.
glide slope of the cooling fluid due to the exergy losses in the Entropy 2012;14(12):491–504. http://dx.doi.org/10.3390/e14030491.
pump, turbine and evaporator. [18] Heberle F, Preißinger M, Brüggemann D. Zeotropic mixtures as working fluids
The second law efficiency of an ORC using zeotropic mixtures as in organic rankine cycles for low-enthalpy geothermal resources. Renew
Energy 2012;37(1):364–70. http://dx.doi.org/10.1016/j.renene.2011.06.044.
working fluid with or without recuperator are approximately [19] DiPippo R. Second law assessment of binary plants generating power from
equal. low-temperature geothermal fluids. Geothermics 2004;33(5):565–86. http://
The second law efficiency in function of the heat carrier temper- dx.doi.org/10.1016/j.geothermics.2003.10.003.
[20] Alfeev V, Brodyanski V, Yagodin V, Nikolsky V, Ivantsov A. Refrigerant for a
ature between 120 °C and 160 °C shows an almost linear behavior cryogenic throtteling unit, Patent 1,336,892; 1973.
for all the investigated mixtures. Between 120 °C and 130 °C iso- [21] Venkatarathnam G. Cryogenic mixed refrigerant processes. International
pentane–isohexane has the highest second law efficiency. While cryogenics monographs series. New York: Springer; 2008.
[22] Lemmon E, Huber M, McLinden M. Reference fluid thermodynamic and
between 130 °C and 160 °C isobutane–isopentane has the highest
transport properties-refprop, standard reference database 23, version 9.0,
second law efficiency. However, the difference in second law effi- National Institute of Standards and Technology; 2007.
ciency between optimized ORCs with zeotropic mixtures as work- [23] Liu B, Chien K, Wang C. Effect of working fluids on organic rankine cycle for
ing fluid is small (around 3 percentage points). As such, thermo- waste heat recovery. Energy 2004;29(8):1207–17. http://dx.doi.org/10.1016/
j.energy.2004.01.004.
economic criteria should be taken into account in order to further [24] Bejan A, Tsatsaronis G, Moran M. Thermal design and optimization. New
differentiate the optimal mixture. York: John Wiley; 1996.

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028
S. Lecompte et al. / Energy Conversion and Management xxx (2014) xxx–xxx 13

[25] Moran M, Shapiro H. Fundamentals of engineering thermodynamics. 5th [31] El-Emam R, Dincer I. Exergy and exergoeconomic analyses and optimization of
ed. New York: John Wiley; 2006. geothermal organic rankine cycle. Appl Therm Eng 2013;58:435–44.
[26] Winterbone D. Advanced thermodynamics for engineers. New York: John [32] Shengjun Z, Huaixin W, Tao G. Performance comparison and parametric
Wiley; 1997. optimization of subcritical organic rankine cycle (ORc) and transcritical power
[27] Deb K. Optimization for engineering design. Algorithms and cycle system for low-temperature geothermal power generation. Appl Energy
examples. Prentice-Hall of India; 1995. 2011;88(8):2740–54. http://dx.doi.org/10.1016/j.apenergy.2011.02.034.
[28] Larjola J. Electricity from industrial waste heat using high-speed organic [33] Wang E, Zhang H, Fan B, Ouyang M, Zhao Y, Mu Q. Study of working fluid
rankine cycle (ORC). Int J Prod Econ 1995;41:227–35. selection of organic rankine cycle (ORc) for engine waste heat recovery. Energy
[29] Lecompte S, Huisseune H, van den Broek M, De Schampheleire S, De Paepe M. 2011;36(5):3406–18. http://dx.doi.org/10.1016/j.energy.2011.03.041.
Part load based thermo-economic optimization of the organic rankine cycle [34] Tchanche B, Lambrinos G, Frangoudakis A, Papadakis G. Exergy analysis of
(ORC) applied to a combined heat and power (CHP) system. Appl Energy micro-organic rankine power cycles for a small scale solar driven reverse
2013;111:871–81. http://dx.doi.org/10.1016/j.apenergy.2013.06.043. osmosis desalination system. Appl Energy 2010;87(4):1295–306. http://
[30] Zhu S, Deng K, Qu S. Energy and exergy analyses of a bottoming rankine cycle dx.doi.org/10.1016/j.apenergy.2009.07.011.
for engine exhaust heat recovery. Energy 2013;58:448–57. http://dx.doi.org/ [35] Mago PJ, Chamra LM, Srinivasan K, Somayaji C. An examination of regenerative
10.1016/j.energy.2013.06.031. organic rankine cycles using dry fluids. Appl Therm Eng 2008;28(8-
9):998–1007. http://dx.doi.org/10.1016/j.applthermaleng.2007.06.025.

Please cite this article in press as: Lecompte S et al. Exergy analysis of zeotropic mixtures as working fluids in Organic Rankine Cycles. Energy Convers
Manage (2014), http://dx.doi.org/10.1016/j.enconman.2014.02.028

You might also like