You are on page 1of 16

Energy Conversion and Management 136 (2017) 11–26

Contents lists available at ScienceDirect

Energy Conversion and Management


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

Thermo-economic analysis of zeotropic mixtures based on siloxanes for


engine waste heat recovery using a dual-loop organic Rankine cycle
(DORC)
Hua Tian, Liwen Chang, Yuanyuan Gao, Gequn Shu ⇑, Mingru Zhao, Nanhua Yan
State Key Laboratory of Engines, Tianjin University, 92 Weijin Road, Nankai District, Tianjin 300072, China

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

Article history: Siloxanes are usually used in the high temperature organic Rankine cycle (ORC) for engine waste heat
Received 9 November 2016 recovery, but their flammability limits the practical application. Besides, blending siloxanes with retar-
Received in revised form 22 December 2016 dants often brings a great temperature glide, causing the large condensation heat and the reduction in
Accepted 23 December 2016
net output power. In view of this, the zeotropic mixtures based on siloxanes used in a dual-loop organic
Available online 11 January 2017
Rankine cycle (DORC) system are proposed in this paper. Three kinds of binary zeotropic mixtures con-
sisting of R123 and various siloxanes (octamethylcyclotetrasiloxane ‘D4’, octamethyltrisiloxane ‘MDM’,
Keywords:
decamethyltetrasiloxane ‘MD2M’), represented by D4/R123, MDM/R123 and MD2M/R123, are selected
Zeotropic mixtures
Dual-loop organic Rankine cycle (DORC)
as the working fluid of the high temperature (HT) cycle. Meanwhile, R123 is always used in the low tem-
Waste heat recovery perature (LT) cycle. The net output power and utilization of heat source are considered as the evaluation
Thermodynamic analysis indexes to select the optimal mixture ratios for further analysis. Based on the thermodynamic and eco-
Economic performance nomic model, net output power, thermal efficiency, exergy efficiency, exergy destruction and electricity
production cost (EPC) of the DORC system using the selected mixtures have been investigated under dif-
ferent operating parameters. According to the results, the DORC based on D4/R123 (0.3/0.7) shows the
best thermodynamic performance with the largest net power of 21.66 kW and the highest thermal effi-
ciency of 22.84%. It also has the largest exergy efficiency of 48.6% and the smallest total exergy destruc-
tion of 19.64 kW. The DORC using MD2M/R123 (0.35/0.65) represents the most economic system with
the smallest EPC of 0.603 $/kW h. Besides, the irreversibility in the internal heat exchanger, turbine
and evaporator of HT cycle contributes most to the total exergy destruction which can serve as the
parameter to be optimized in the further study.
Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction as one of the most promising technologies for the engine waste
heat recovery.
The shortage of oil resources created an unprecedented energy As the exhaust gas of the engine has a high temperature with
crisis around the world. Internal combustion engines (ICEs), are the over 500 °C, there are two methods to fit this issue: the suitable
main oil-consuming equipment. The research shows in ICEs about working fluid with good thermostability and high system effi-
one third of the combustion heat is wasted through the exhaust gas ciency; the dual-loop organic Rankine cycle (DORC) system. For
[1]. If such waste heat is reused, the engine fuel economy and envi- the suitable working fluid, siloxanes [4–8], with excellent environ-
ronment pollution can be effectively relieved, bringing tremendous ment and cycle characteristics, are commonly used as the working
economic and environmental benefit. fluid, but their flammability limits the application. Besides, some
As a mature and effective heat recovery method, organic Rank- refrigerants are also widely used in ORC, such as R245fa [8–11],
ine cycle (ORC) has received more and more attention [2]. It fea- and R123 [11–14]. However, due to higher global warming poten-
tures simple structure, good applicability, and user friendliness tial (GWP) and toxicity, refrigerants usually have poorer environ-
for small-scale power generation systems. Besides, it consumes mental performance in comparison with siloxanes. To solve this
no additional fuel for extra power [3]. Therefore, it is considered problem, a combination of siloxanes and non-inflammable refrig-
erants is proposed. The mixture has the merit to overcome the
drawbacks of each fluid and realize the complementary advan-
⇑ Corresponding author. Tel./fax: +86 22 27409558. tages. The refrigerants can be seen as the retardants to suppress
E-mail address: shugqtju@163.com (G. Shu).

http://dx.doi.org/10.1016/j.enconman.2016.12.066
0196-8904/Ó 2016 Elsevier Ltd. All rights reserved.
12 H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26

Nomenclature

Abbreviations Nu Nusselt number


D4 octamethylcyclotetrasiloxane Pr Prandtl number
D5 decamethylcyclopentasiloxane Q heat absorbed or released (kW)
DORC dual-loop organic Rankine cycle Re Reynolds number
EPC electricity production cost s specific entropy (kJ/kg K)
GWP global warming potential W work input or output (kW)
HEX heat exchanger x vapor quality
HT high temperature DHvap heat of vaporization (kJ/kg)
ICEs internal combustion engines DTglide temperature glide (K)
IHE internal heat exchanger m velocity (m/s)
LT low temperature
MD2M decamethyltetrasiloxane Greek letters
MDM octamethyltrisiloxane a heat transfer coefficient (W/m2 K)
MM hexamethyldisiloxane e effectiveness of IHE (%)
ODP ozone depletion potential g efficiency (%)
Pcond1 condensation pressure in HT cycle k thermal conductivity (W/m K)
Pmax1 evaporation pressure in HT cycle l viscosity (Pa s)
RD relative deviation q density (kg/m3)
Tevap2 evaporation temperature in LT cycle
TIC total investment cost Subscripts
Tmax1 turbine inlet temperature in HT cycle b boiling point
UHS utilization of heat source
c condenser
XR123 mass fraction of R123 cri critical point
e evaporator
Symbols exh exhaust gas
A heat transfer area (m2) f working fluid
Cp specific heat capacity (kJ/kg K) fq steam in HT cycle
E exergy (kW) i IHE
G mass velocity (kg/m2 s) net net output
h specific enthalpy (kJ/kg) p pump
I exergy loss (kW) t turbine
K overall heat transfer coefficient (W/m2 K) total total heat or exergy
M molar mass (g/mol) w1 water inlet
m mass flow (kg/s) w2 water outlet

the flammability of siloxanes and the usage quantity of the refrig- heat source. Liu et al. [22] studied some zeotropic mixtures includ-
erants can be reduced in this way. The thought is supported by the ing MDM/MD2M in ORC, and found there existed optimal mole
work of Oellrich [15] who researched the flammability suppression fractions to maximize the cycle efficiency and the net power. Zhou
of propane with perfluoropropane for application in refrigeration. et al. [23] discussed various zeotropic mixtures used in a dual-loop
Moreover, Yang et al. [16] studied the flame-resistant effect of non- system for waste heat recovery from the internal combustion
flammable refrigerants on flammable refrigerants by the experi- engine. The simulation results revealed that with RC318/R1234yf
ment, and found with the number of nonflammable components as the working fluid in the low temperature loop, the engine out-
increased, the area of the nonflammable zone expanded. Li et al. put can be improved by nearly 14.4%. Chys et al. [24] explored
[17] researched the flame-resistant effect of the refrigerant R134 the potential of zeotropic mixtures used in ORC. Results showed
on R290 and R152a, and found adding R134 could change the that the ORC using mixtures always had an improvement in the
lower flammable limit of the flammable refrigerant. Kondo et al. cycle efficiency for both heat source types compared with using
[18] measured the dilution effect of CO2 on the flammability limit, pure fluids.
and the obtained values were analyzed using the extended Le However, when siloxanes are blended with nonflammable
Chatelier’s formula. Results showed that the addition of CO2 was refrigerants, the condensation temperature glides of the mixtures
a typical way of ensuring safety in the use of flammable gases. are usually very large (>80 K). If the mixtures are cooled by the
The research in our group [19,20] also indicated that the retardant water directly, the cycle effects are not ideal. Besides, the exhaust
CO2 could reduce the flammable zones of the hydrocarbon. through a single heat exchanger still has a rather high temperature
In addition, the composed mixtures have another advantage in and the direct rejection would result in a great waste. Hence, it’s
ORC application. For zeotropic mixtures, the temperature is vari- feasible to consider using these two parts of heat simultaneously
able during the phase change process, and the dew point temper- to form a dual loop cycle.
ature is higher than the bubble point temperature at a fixed phase Currently, in order to recover the engine waste heat effectively,
transition pressure. The temperature difference of the two is researches on the DORC are carried out. Song and Gu [25] pre-
known as the temperature glide, which ensures a better match sented a DORC system for engine waste heat recovery to generate
with the heat source or heat sink, improving the cycle perfor- additional power. The simulation results revealed that the maxi-
mance. Dong et al. [21] investigated zeotropic mixtures MM/ mum net power output of the DORC system could reach
MDM as the working fluid for the high temperature ORC, and found 111.2 kW, increasing the engine power by 11.2%. Shu et al.
the mixtures led to a cycle efficiency increase compared to pure [26,27] proposed a DORC to recover the waste heat of the exhaust,
fluids, and MM/MDM (0.4/0.6) showed the best match with the engine coolant and residual heat of the HT loop. Results showed
H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26 13

that the system performed better at the high operating load, and
R1234yf was found to be the best working fluid. Using water as
the working fluid of the HT loop, the system obtained maximum
net output power (39.67 kW) and exergy efficiency (42.98%). Using
siloxanes, the dual loop cycle with double regenerators performed
better. Sciubba et al. [28] raised a DORC based on a single steam
Rankine cycle loop to recover waste energy from marine engines
of different power range. The paper showed adding a second recov-
ery loop improved the system performance both in terms of recov-
ered electric power (up to 8.11% and 2.67% respectively in small
and large application size) and heat source utilization rate.
It is observed that the researches on the DORC mainly focus on
using the pure working fluids as the cycle medium. Therefore, it’s
meaningful and thoughtful to conduct an in-depth analysis and
research on applying the mixtures to the DORC. So this paper pro-
posed an application of mixtures based on siloxanes blending with
refrigerant retardants used in the engine waste heat recovery,
which can extend the application range on the basis of improving
cycle efficiency and reducing exergy loss. A set of DORC system is
designed to recover the exhaust energy and the condensation heat
generated by the temperature glide. Thermodynamic and eco-
nomic theoretical models of the system are built by MATLAB soft-
Fig. 1. Schematic diagram of the DORC system.
ware. Net output power, thermal efficiency, and electricity
production cost (EPC) of the DORC system using the selected mix-
tures have been investigated under different operating parameters. perature (LT) cycle (the blue lines). The HT cycle consists of an
Moreover, the exergy analysis on exergy efficiency and exergy evaporator (Eva1), a turbine (Tur1), an internal heat exchanger
destruction is another emphasis to be evaluated. (IHE), a condenser (Con1), and a pump (Pum1). The LT cycle
includes an evaporator (Eva2), a turbine (Tur2), a condenser
2. System description and working fluid selection (Con2), and a pump (Pum2). Among them, the condenser in the
HT cycle acts as the pre-evaporator in the LT cycle. Thus, the two
2.1. Topping ICE system separate cycles are coupled.
The thermodynamic processes of the DORC system are shown in
In this paper, an in-line 6-cylinder 4-stroke turbocharged diesel Fig. 2. The working fluids in Fig. 2(a) and (b) are the mixture D4/
engine is used as the topping system. The main testing parameters R123 and pure R123, respectively. In the HT cycle, the mixture
of the engine are listed in Table 1. Based on the fuel composition as which is pressurized into high pressure state by Pum1 (1–2), is pre-
shown in Table 1, the average chemical formula for common diesel heated in IHE (2–3). Then it flows into Eva1 (3–4), to be heated by
fuel is considered as C12H23. Supposing that the fuel is in complete the high temperature exhaust. Later, it enters Tur1 to generate
combustion, the composition of the exhaust on mass basis can be electricity (4–5). The superheated vapor goes through IHE to trans-
figured out which is also listed in Table 1. The composition is used fer heat to the working fluid out of Pum1 (5–6). Finally, the work-
for evaluating the gas properties. The outlet temperature of the ing fluid flows into Con1 to be cooled (6–1). Thus, the HT cycle is
exhaust should be at least 393.15 K [29] to avoid acid dew point finished. In the LT cycle, the working fluid pressurized in Pum2
for fuels containing sulphur. (7–8), flows into the Con1 to be heated by the released heat of
the condensation process in HT cycle (8–9). Afterwards, it is sent
2.2. Bottoming DORC system to Eva2 and heated up into a superheated vapor state by the low
temperature exhaust (9–11). The superheated vapor enters Tur2,
Fig. 1 provides the configuration of the DORC system, including to deliver work and generate electricity (11–12). Finally, the work-
a high temperature (HT) cycle (the magenta lines) and a low tem- ing fluid enters Con2 and is condensed into a saturated liquid state
by the supplied cooling water (12–7). The whole process then
Table 1 completes.
Main engine testing parameters, exhaust composition and fuel
composition.

Parameter Value 2.3. Working fluid selection


Main engine testing parameters
Rated speed (r/min) 2200 The working fluid is very critical to the cycle performance of the
Rated power (kW) 243 DORC system. It should be environment friendly, safe, low toxic,
Torque (N m) 1055
and make the system more efficient and stable [24]. Siloxanes
Exhaust temperature (K) 718.15
Exhaust pressure (MPa) 0.15 are a class of fluids composed of molecules containing alternate sil-
Exhaust mass flow rate (kg/h) 1020.74 icon and oxygen atoms in either a linear or cyclic arrangement usu-
Exhaust composition ally with two or three organic groups attached to silicon atom [30].
CO2 (%) 15.94 They have excellent thermal stability [31] with zero ozone deple-
H2O (%) 5.66 tion potential (ODP) and small global warming potential (GWP).
N2 (%) 72.92 Due to the long scientific names of siloxanes, in the academic field,
O2 (%) 5.48
they have a simplified representation. The bifunctional repetitive
Fuel composition unit, (CH3)2SiO is termed D-unit, and the monofunctional terminal
Saturated hydrocarbons (%) 75
Aromatic hydrocarbons (%) 25
group, (CH3)3SiO1/2 is termed M-unit. Therefore, cyclic molecules
are written as Dx, and linear molecules are written as MDx-2M,
14 H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26

of R123 is also relatively high (around 600 K) which suits the oper-
ating condition well. Hence, coordinating their performance char-
acteristics, the mixtures based on siloxanes and refrigerants
(R123) are selected as the working fluid in the HT cycle to recover
the waste heat from the high temperature exhaust. In the LT cycle,
R123 is used alone to recover the waste heat from the low temper-
ature exhaust and the condensation heat generated by the temper-
ature glide of the mixtures in the HT cycle. Table 2 lists the
properties of some pure siloxanes and R123. The thermodynamic
properties are calculated by the commercial software REFPROP
9.0 [35].
For the mixture proportion, Zabetakis [36] investigated flamma-
bility envelopes of some hydrocarbons mixed with inert compo-
nents, and found generally 0.3 vol fraction was enough to take
most hydrocarbons outside the flammability zone. Garg [37]
deduced that the heavier the molecular weight of the inert additive
was, the smaller the mole fraction of the inert additive was ade-
quate to suppress the flammability, and only 0.18 mol fraction
R245fa can achieve the same suppression effect as 0.3 mol fraction
CO2. Due to the lack of experimental data, 0.3 mass fraction of R123
is preliminarily set as the minimum limit in this study. Fig. 3 shows
the temperature glides of the mixtures with the mass fraction of
R123 changing from 0.3 to 0.7 at the condensation pressure of
0.1 MPa. Combining Fig. 3 with Table 2, it can be observed that
the temperature glide of zeotropic mixtures is related to the nor-
mal boiling point (Tb) of each composition. The mixture with larger
Tb difference between the two components has the higher temper-
ature glide under the same mass fraction of R123 [38]. The temper-
ature glides of D5/R123 are above 120 K under most mixing ratios,
which may result in a high concentration shift and fractionation of
the mixture components [39]. Therefore, D5/R123 will not be con-
sidered in the following analysis. On the other hand, if the compo-
sition of R123 is too high, meaning lower temperature glide, the
heat transfer of the LT cycle would get worse, and the net power

Fig. 2. T–s diagram of the DORC system, (a) for HT cycle and (b) for LT cycle.

x being the total number of silicon atoms [32]. For the high tem-
perature ORC, R123 proves to have good performance in terms of
theories and experiments [11–14]. Although R123 is classified as
Class II Ozone-depleting Substances, the GWP (77) and ODP
(0.012) of R123 are actually relatively low compared to the other
substances in the same class [33]. Especially, the GWP of R123
has met the requirement of the most restrictive environmental
regulations worldwide (GWP < 150). Also R123 has a short atmo-
spheric lifetime which highly decreases its damage to the environ-
ment. Besides, R123 system presents a minor negative pressure at
halt state, which also helps avoid leakage. Therefore it can main-
tain a low annual leakage rate (<0.5%) and reduce the possibility
of poisoning accident. Moreover, it is nonflammable [34] which
can serve as the flame retardant. The decomposition temperature Fig. 3. Condensation temperature glide of different mixtures.

Table 2
Properties of the pure working fluids.

Fluids M (g/mol) Tb (K) Tcri (K) Pcri (MPa) Combustibility ODP GWP
MDM 236.53 425.66 564.09 1.415 Flammable 0 Small
D4 296.62 448.5 586.5 1.332 Flammable 0 Small
MD2M 310.69 467.51 599.4 1.227 Flammable 0 Small
D5 370.77 484.05 619.15 1.16 Flammable 0 Small
R123 152.93 300.97 456.83 3.6618 Non-flammable 0.012 77
H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26 15

of the LT cycle is very small, so the dual loop cycle proposed in the
paper would not come into effect. Besides, the requirements for the
pinch temperature difference in IHE may not be met. In conclusion,
the mass fraction of R123 is set as 0.3–0.7 for both D4/R123 and
MDM/R123. And for MD2M/R123, the mass fraction of R123 is
0.41–0.65 to guarantee the convergence of calculation.

3. Thermodynamic model

In order to simplify the computing, some assumptions are given


before establishing the mathematic model:

(1) The whole system operates under equilibrium state and


steady condition. The pressure drop and heat loss in the heat
exchangers are ignored.
(2) The compositions of the mixtures don’t change during the
whole cycle.
(3) The isentropic efficiencies of the turbine and the pump are
set to 0.7 and 0.8, respectively. The effectiveness of IHE is
set to 0.7.
(4) The ambient condition and cooling water are set to 0.1 MPa
and 298.15 K. The degree of superheat (temperature differ-
ence between point 11 and point 10 g in Fig. 2) and the con-
densation temperature in the LT cycle are 5 K and 308 K,
respectively. The pinch point temperature of gas–liquid heat
exchangers is 30 K, and for liquid–liquid heat exchangers is
10 K.

According to the first and second laws of thermodynamics, the


thermodynamic model of the DORC system is established.
First, the exergy at each steady point i is defined as:

E_ i ¼ m
_ ½ðhi  h0 Þ  T 0 ðsi  s0 Þ ð1Þ
where the subscript ‘‘0” means the reference state, referring to the
ambient condition in this paper.
Then, specific analysis for each component of the system is as
follows. Fig. 4. Q–T diagram of the DORC system, (a) for HT cycle and (b) for LT cycle.

3.1. HT cycle process


Heat absorbed in Eva1:
For pump Pum1:
Input power of Pum1: Q_ e1 ¼ m
_ f 1 ðh4  h3 Þ ¼ m
_ exh ðhexh;in  hexh;out Þ ð7Þ
_ p1 ¼ m
W _ f 1 ðh2s  h1 Þ=gp
_ f 1 ðh2  h1 Þ ¼ m ð2Þ Exergy destruction of the Eva1:
 
Exergy destruction of Pum1: I_e1 ¼ m
_ exh ðhexh;in  hexh;out Þ  T 0 ðsexh;in  sexh;out Þ
I_p1 ¼ m
_ f 1 T 0 ðs2  s1 Þ ð3Þ _ f 1 ½ðh4  h3 Þ  T 0 ðs4  s3 Þ
m ð8Þ
For IHE:
For turbine Tur1:
Effectiveness of IHE:
Output power of Tur1:
e ¼ T 5  T 6 =T 5  T 2 ð4Þ
_ t1 ¼ m
W _ f 1 ðh4  h5s Þgt
_ f 1 ðh4  h5 Þ ¼ m ð9Þ
Heat transfer amount in the IHE:
Exergy destruction of Tur1:
Q_ i ¼ m
_ f 1 ðh3  h2 Þ ¼ m
_ f 1 ðh5  h6 Þ ð5Þ
I_t1 ¼ m
_ f 1 T 0 ðs5  s4 Þ ð10Þ
Exergy destruction of IHE:
For condenser Con1: The condenser in HT cycle is also the pre-
I_i ¼ m
_ f 1 T 0 ðs3  s2 þ s6  s5 Þ ð6Þ
evaporator in LT cycle.
For evaporator Eva1: A transcritical heating process happens in Heat exchanged in Con1:
Eva1. The heat absorbed and irreversibility are generated and cal-
culated by the pinch point temperature difference (PPTD) method Q_ c1 ¼ m
_ f 1 ðhfq;in  hfq;out Þ ¼ m
_ f 2 ðh9  h8 Þ ð11Þ
which is commonly used for mass flow rate calculation and ther-
modynamic states’ computation. The PPTD method was illustrated Exergy destruction of Con1:
 
in detail for ORC in the previous work completed by our group _ f 1 ðhfq;in  hfq;out Þ  T 0 ðsfq;in  sfq;out Þ
Ic1 ¼ m
[2,12]. Fig. 4 gives the Q-T diagram for the heat exchangers of the
_ f 2 ½ðh9  h8 Þ  T 0 ðs9  s8 Þ
m ð12Þ
cycle.
16 H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26

3.2. LT cycle process 4.1. Heat exchanger surface area calculation

For pump Pum2: Shell-and-tube heat exchanger is chosen for the heat exchang-
ers in this work. The tube banks consist of stainless steel tubes
_ p2 ¼ m
W _ f 2 ðh8s  h7 Þ=gp
_ f 2 ðh8  h7 Þ ¼ m ð13Þ with a hexagonal arrangement and the shell side sets up baffle
plates. While several heat exchanger parameters are given in
I_p2 ¼ m
_ f 2 T 0 ðs8  s7 Þ ð14Þ Table 3, several other parameters are determined by the number
of tubes which is considered as the only manipulating variable to
For evaporator Eva2: be modified later to satisfy the pressure constraints. The working
fluid always flows in the tube side for Eva1, Eva2 and Con2 while
Q_ e2 ¼ m
_ f 2 ðh11  h9 Þ ¼ m
_ exh ðhexh;out  hexh;final Þ ð15Þ
the two-phase fluid flows in the tube side for Con1 and IHE.
  The heat transfer surface area of shell-and-tube heat exchanger
I_e2 ¼ m
_ exh ðhexh;out  hexh;final Þ  T 0 ðsexh;out  sexh;final Þ is calculated as:
_ f 2 ½ðh11  h9 Þ  T 0 ðs11  s9 Þ
m ð16Þ
Q_
A¼ ð28Þ
For turbine Tur2: K DT lm
_ t2 ¼ m
W _ f 2 ðh11  h12s Þgt
_ f 2 ðh11  h12 Þ ¼ m ð17Þ where Q_ is the heat transfer rate; K is the overall heat transfer coef-
ficient; DT lm is the logarithmic mean temperature difference
I_t2 ¼ m
_ f 2 T 0 ðs12  s11 Þ ð18Þ (LMTD).
The overall heat transfer coefficient is obtained by:
For condenser Con2:  1
do rs;i do do lnðdo =di Þ 1
K¼ þ þ þ r s;o þ ð29Þ
Q_ c2 ¼ m
_ f 2 ðh12  h7 Þ ¼ m
_ w ðhw2  hw1 Þ ð19Þ ai di di 2k ao
where ai and ao are heat transfer coefficient inside and outside
I_c2 ¼ m
_ f 2 ½ðh12  h7 Þ  T 0 ðs12  s7 Þ
tubes, respectively; k is the thermal conductivity of the tube mate-
_ w ½ðhw2  hw1 Þ  T 0 ðsw2  sw1 Þ
m ð20Þ rial and set as 16.3 W/m K; do and di are outer and inner diameters
of tubes; r s;i and rs;o are fouling resistance inside and outside tubes,
respectively.
3.3. Thermodynamic performance parameters of DORC system The logarithmic mean temperature difference is calculated as
follow:
Some parameters are chosen to evaluate the thermodynamic
performance of the DORC system. Dt0  Dt00
DT lm ¼ ð30Þ
The net output power of the HT cycle: ln ðDt 0 =Dt 00 Þ

_ net;HT ¼ W
W _ t1  W
_ p1 ð21Þ where Dt 0 and Dt00 are the maximum and minimum temperature
differences at the ends of the heat exchangers, respectively.
The net output power of the LT cycle:

_ net;LT ¼ W
_ t2  W
_ p2 4.1.1. Heat transfer and pressure drop in shell side
W ð22Þ
The shell side heat transfer coefficient is calculated by the Bell-
The total net output power of the DORC: Delaware method [41], which adopts some correction factors
based on the heat transfer factor for the ideal tube bank consider-
_ net ¼ W
W _ net;HT þ W
_ net;LT ð23Þ ing effects of baffle window flow, baffle leakage and bundle bypass.
The utilization of the heat source (UHS) is used to evaluate use The baffle plate is equidistant and the effect of adverse tempera-
degree of the exhaust, which can be calculated as: ture gradient in laminar flow is negligible in this paper.
The shell side heat transfer coefficient is given as:
UHS ¼ ðhexh;in  hexh;final Þ=ðhexh;in  hexh;min Þ ð24Þ
ashell ¼ aid jc jl jb ð31Þ
where hexh; min represents the enthalpy of exhaust’s acid dew-point.
where aid is the heat transfer coefficient for pure cross-flow in an
The thermal efficiency of the DORC:
ideal tube bank; jc ; jl ; jb are correction factors for baffle window
gth ¼ W_ net =Q_ total ¼ W_ net =ðQ_ e1 þ Q_ e2 Þ ð25Þ
The exergy efficiency of the DORC: Table 3
Some parameters of shell-and-tube heat exchanger.
. 
gex ¼ W_ net =E_ total _ net E_ exh;in  E_ exh;final þ E_ w1  E_ w2
¼W ð26Þ Parameter Value
Geometric dimensions (mm)
The total exergy destruction of the DORC system: Tube outer diameter 6
Tube inner diameter 5
I_ ¼ I_p1 þ I_i þ I_e1 þ I_t1 þ I_c1 þ I_p2 þ I_e2 þ I_t2 þ I_c2 ð27Þ Tube pitch 15
Pressure drop allowance (bar)
Exhaust side 0.1
4. Economic model Cooling water side 0.2
Working fluid side 0.3

The electricity production cost (EPC) is selected to evaluate the Fouling resistance (m2 K/W)
economic performance of the DORC system in this paper. The work Cooling water 1.7  104
Exhaust gas 1.7  104
in this part is mainly referred from the book [40], which is widely Working fluid 8.6  105
used for analysis, synthesis and design of chemical processes.
H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26 17

flow, baffle leakage effects and bundle bypass effects, respectively The pressure drop for tube-side fluid is calculated as follow:
[41]. And the corresponding values of the latter two correction fac-
DPtube ¼ DPi þ DPN ð42Þ
tors are 0.8 and 0.9.
The heat transfer coefficient for pure cross-flow in an ideal tube DP i represents the frictional pressure drop, for single-phase
bank is calculated as follow: fluid which is given by [44]:
 0:14
GCp l f p G2 L
aid ¼ jH ð32Þ DP i ¼ ð43Þ
Pr 2=3 lw 2qD
where jH is the Colburn j-factor for an ideal tube bank, the subscript where the friction factor f p is calculated by Eq. (44) for fully devel-
‘‘w” stands for the property of the shell side flow at wall- oped flow.
temperature.
The shell side pressure drop is calculated by the following f p ¼ 0:316Re1=4 ; Re < 2  104
ð44Þ
equations: f p ¼ 0:184Re1=5 ; Re P 2  104
 0:14
_ 2 Nc
m l DP N is the pressure drop of inlet and outlet connecting tube
DPbk ¼ 4f k ð33Þ
2Ac q
2 l w which is given by:
qv 2
_2
m DPN ¼ 1:5 ð45Þ
DPwk ¼ ð2 þ 0:6Ncw Þ ð34Þ 2
2Ab Ac q
 4.1.2.2. Heat transfer and pressure drop for two-phase fluid. Accord-
Ncw
DPshell ¼ ½ðN b  1ÞDPbk Rb þ N b DPwk Rl þ 2DPbk Rb 1 þ ð35Þ ing to the Chen correlation [45], the heat transfer coefficient of
Nc
the two-phase fluid during the phase change process can be calcu-
where DPbk and DPwk represent the ideal friction pressure drops in lated by:
cross-flow and window flow, respectively. f k is the friction factor
aTP ¼ aLS F ð46Þ
for ideal tube banks; N cw and N c are the effective number of rows
crossed in window flow and cross-flow, respectively. N b is the num- The heat transfer coefficient assuming liquid phase flowing
ber of baffles; Rl and Rb are the correction factors for baffle leakage alone in the tube aLS is calculated from the Dittus-Boelter correla-
and bundle bypass [41] with corresponding values of 0.6 and 1, tion [45]:
respectively.  0:8
D kl
aLS ¼ 0:023 Gð1  xÞ ðPrl Þ0:4 ð47Þ
4.1.2. Heat transfer and pressure drop in tube side ll D
4.1.2.1. Heat transfer and pressure drop for single-phase fluid. The where G is the mass velocity; x is the vapor quality.
Petukhov correlation [42] is used to predict the heat transfer coef- The Reynolds number factor F is computed from [46]:
ficient for single-phase fluid in smooth tube
F ¼ 1:0; 1=X tt 6 0:1
ðf =8ÞRePr ð48Þ
Nu ¼ h i ð36Þ F ¼ 2:35ð1=X tt þ 0:213Þ0:736 ; 1=X tt > 0:1
0:5 2=3
12:7ðf =8Þ ðPr  1Þ þ 1:07
The Lockhart–Martinelli parameter Xtt is calculated as:
 0:9  0:5  0:1
k 1x qv ll
a ¼ Nu ð37Þ X tt ¼ ð49Þ
D x ql lv
where a is the heat transfer coefficient; Nu is the Nusselt number; k where the subscript ‘‘v” and ‘‘l” represent the property of the vapor
is the fluid thermal conductivity. and liquid, respectively.
Re, Pr and the friction factor f in Eq. (36) is obtained by: The condensation heat transfer coefficient of zeotropic mixtures
qmD amix is obtained by [47]:
Re ¼ ð38Þ
l 1
¼
1
þ
YV
ð50Þ
amix amono aVS
lCp
Pr ¼ ð39Þ where Y V ¼ xCpv
DT glide
, amono is the condensing heat transfer coeffi-
k DHv ap
cient calculated with mixture properties using the Chen correlation
2
f ¼ ð1:82lgRe  1:5Þ ð40Þ for pure fluids.
aVS is the heat transfer coefficient assuming vapor phase to be
For supercritical fluid, the Krasnoshchekov-Protopopov correla- flowing alone in the tube, given by:
tion [43] is applied for the Nu calculation:
 0:8
!0:35  GxD ðPr v Þ0:4 kv
0:33  0:11 aVS ¼ 0:023 ð51Þ
ðf =8ÞRePr Cp kb lb lv D
Nu ¼ h i
0:5
12:7ðf =8Þ ðPr2=3  1Þ þ 1:07 Cpw kw lw The frictional pressure drop of two-phase fluid inside the tubes
is obtained by [46]:
ð41Þ
DPi;TP ¼ DPi;LS /2LS ð52Þ
where a, Re, Pr and f calculations use Eqs. (37)–(40) and the
subscript ‘‘b” and ‘‘w” represent the property of the working where DPi;LS is the frictional pressure drop in the case of liquid
fluid at bulk-temperature and wall-temperature, respectively. phase flowing alone in the tube, calculated by Eqs. (43) and (44).
Cp is the averaged over cross-section specific heat under con- The two-phase friction multiplier /2LS is calculated by the fol-
stant pressure. lowing equations [46]:
18 H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26

12 1 The capital recovery factor (CRF) is estimated based on the fol-


/2LS ¼ 1 þ þ 2 ð53Þ
X X lowing equation:
 0:5  ið1 þ iÞ
time
q 1  x Re0:1 CRF ¼ ð61Þ
X ¼ 18:65 v v ð54Þ ð1 þ iÞ
time
1
ql x Re0:5
l
where i is interest rate, whose value is 5%; time is economic life
4.2. Investment calculation time, whose value is 15 years [2].
The annuity of the investment Ank can be expressed as:
The total investment cost of the DORC system depends mainly Ank ¼ Cost2014 CRF ð62Þ
on the cost of each component. The system total investment cost
(TIC) based on economic situation in year 2014 in the present study The electricity production cost (EPC) can be obtained by
is obtained below: equation:

Cost 2001 ¼ C BM;HEX þ C BM;Turbine þ C BM;Pump ð55Þ Ank þ f k Cost2014


EPC ¼ ð63Þ
W_ net hfullload
Cost 2001 CEPCI2014
Cost 2014 ¼ ð56Þ wherein, fk is operation, maintenance and insurance cost factor,
CEPCI2001
whose value is 1.65%; hfull-load is full-load operation hours, whose
wherein, CEPCI2001 = 382, CEPCI2014 = 586.77 (CEPCI means Chemical value is 7500 h [2].
Engineering Plant Cost Index).
The bare module cost is calculated by the following equations to 5. Model validation
estimate the investment cost of each component referring to [40].
The coefficients for the calculation of bare module cost of compo- The RD (relative deviation) is determined as
nents are listed in Table 4, which are chosen based on the compo-
nents’ types. ðref  estÞ
RD ¼  100% ð64Þ
ref
C BM ¼ C P F BM ð57Þ
where ref symbol means the reference value and est symbol means
lg C P ¼ K 1 þ K 2 lg X þ K 3 ðlg XÞ
2
ð58Þ the estimated value.
According to the equations above, a program is written by
F BM ¼ B1 þ B2 F M F P ð59Þ MATLAB R2014a. The fluids’ state parameters are calculated by
linking MATLAB and REFPROP 9.0. Then the energy conversion
2 and exergy distribution in each component are also computed in
lg F P ¼ C 1 þ C 2 lg P þ C 3 ðlg PÞ ð60Þ
this program to further acquire the thermodynamic and economic
where CBM is bare module cost: sum of direct and indirect costs for performance indicators. The transcritical model of HT cycle is ver-
each component; CP is purchased equipment cost in base condi- ified with the same operating parameters in Ref. [48] to validate
tions: component made of the most common material and opera- the results. R134a/R32 (0.7/0.3) is selected as the working fluid.
tion at ambient pressure; FBM is bare module factor; X is the Results are shown in Table 5. The relative deviations are acceptable
capacity parameter, which refers to the expansion power, the pump with the biggest one of 4.04% corresponding to the thermal effi-
power consumptions and the heat transfer area for the turbine, the ciency. The average condensation temperature is selected as the
pump and the heat exchanger, respectively; FM is the material fac- operating parameter for the condensation process in Ref. [48].
tor; FP is the pressure factor; P is the design pressure. Since the software ChemCAD is used in Ref. [48] to evaluate the

Table 4
Coefficients for the calculation of bare module cost of components.

Components K1 K2 K3 C1 C2 C3 B1 B2 FM FBM
Pump 3.870 0.316 0.122 0.245 0.259 0.014 1.89 1.35 2.35 /
Turbine 2.705 1.440 0.177 / / / / / / 6.2
HEX 4.325 0.303 0.163 0.039 0.082 0.012 1.63 1.66 1.35 /

Table 5
Comparison of the present calculated results with Ref. [48]

Working fluid Thermal efficiency (%) Optimal thermal efficiency (%) Exergy efficiency of condensation process (%)
R134a/R32(0.7/0.3) Ref. 13.35 15.08 81.64
This work 13.89 14.85 79.90
RD (%) 4.04 1.53 2.13

Table 6
Comparison of the present calculated results with Ref. [29]

Working fluid P ORC (kW) gORC _ f (kg/s)


m V_ 3 (m3/s) v 4 =v 3 Dh30 4 (kJ/kg)

Benzene Ref. 349.3 0.1986 2.737 0.052 107 130.5


This work 349.05 0.2025 2.6747 0.051 107.8 130.5
RD (%) 0.0716 1.9637 2.2762 1.9231 0.7477 0
H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26 19

properties of fluids other than REFPROP, there exists a slight differ- in such a relative deviation. The subcritical model of LT cycle is
ence in the condensation pressure which is critical for the cycle compared with same status parameters in Ref. [29] using benzene
performance. And the properties of each state point also have a lit- as the working fluid. The present results are in good agreement
tle difference due to the use of different calculation softwares. All with those in Ref. [29]. Results are shown in Table 6. These small
these factors have an impact on the thermal efficiency, resulting errors are probably due to the difference of the software version.
In Ref. [29], REFPROP 8.0 is used, while REFPROP 9.0 is used in
the present study.

6. Results and discussion

6.1. Selection of mixture ratios

In this part, the net output power (W_ net ) and utilization of heat
source (UHS) are chosen as the evaluation indexes, in order to
select the optimal mixture ratios for further analysis. The variable
parameters are mass fraction of R123 (XR123) and evaporation pres-
sure in HT cycle (Pmax1).

6.1.1. The net output power (W _ net ) of the DORC system


Fig. 5 gives the variation of net output power (W _ net ) of the DORC
system over the mass fraction of R123 (XR123) and evaporation
_ net
pressure in HT cycle (Pmax1). In general, with XR123 increases, W
_ net decreases slowly after
gradually increases. As Pmax1 increases, W
reaching a peak value. Therefore, there exist optimal XR123 and
Pmax1 for three groups of mixtures to obtain maximum W _ net . The
mixture MDM/R123 always has a smaller W _ net , while the other
_
two have the similar magnitude in W net . MDM/R123 (0.3/0.7) has
_ net of 19.95 kW at Pmax1 = 11 MPa. D4/R123 (0.3/0.7)
the largest W
obtains maximum W _ net of 21.6 kW at Pmax1 = 10 MPa. MD2M/
R123 (0.35/0.65) gains maximum W _ net of 21.41 kW at
Pmax1 = 9 MPa.
The three groups of mixtures have the same change trend and
the reasons are also similar. Take the mixture D4/R123 as an exam-
ple to explain the phenomenon. Fig. 6 shows the net power in HT
cycle and LT cycle for D4/R123. Horizontally, with the increase of
XR123, the enthalpy drop in turbine (Tur1) rises, which leads to
the increasing expansion power. So the net power in HT cycle
(W_ net;HT ) increases gradually. Besides, the mass flow rate in LT cycle
_ f 2 ) decreases slightly, resulting in a drop of the net power in LT
(m

Fig. 5. Net output power (Wnet) of the DORC system. Fig. 6. Net output power (Wnet) for D4/R123.
20 H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26

cycle (W_ net;LT ). But the former increases faster, so the total net
power also increases. The changes of these parameters with
respect to XR123 for D4/R123 can be intuitively reflected in the
Appendix. Vertically, since the HT cycle is a transcritical cycle, with
the increasing Pmax1, expansion power increases with a limited
rate, while the pump power increases rapidly. At a high pressure,
the increase of the former can’t offset the loss of the latter, result-
ing in the final decrease in W _ net;HT after the increment stage at the
lower pressure. Besides, temperature after the turbine (T5)
decreases, so does the temperature after IHE (T6), which makes
the heat transfer from HT cycle to LT cycle decrease leading to a
_ f 2 . The LT cycle is a subcritical cycle, so expan-
slight reduction of m
sion power and pump power change little, and W _ net;LT decreases
slightly. The above reasons lead to the change trend of net power.

6.1.2. The utilization of heat source (UHS) of the DORC system


The UHS is to measure the utilization rate and perfection degree
of the engine exhaust waste heat. Fig. 7 shows the variation of UHS
over the mass fraction of R123 (XR123) and evaporation pressure in
HT cycle (Pmax1). By and large, with XR123 and Pmax1 increase, UHS
firstly increases, and then decreases. Thus, for three groups of mix-
tures, there exist optimal XR123 and Pmax1 corresponding to optimal
UHS. At Pmax1 = 11 MPa, MDM/R123 (0.68/0.32) has the largest UHS
of 93.08%. For D4/R123 (0.6/0.4), the optimal UHS is 94.81%. And as
for MD2M/R123 (0.55/0.45), it’s 94.79%.
Taking the mixture D4/R123 as an example, reasons are as fol-
lows. As can be seen from Eq. (24), the UHS depends on the specific
enthalpy of exhaust in different states related to corresponding
temperatures. The final outlet temperature of exhaust (Texh,final) is
the key factor. With XR123 and Pmax1 increase, Texh,final goes down
first, and then goes up, but the whole changing scale is small. So
the UHS also changes a little numerically.
To sum up, the three groups of mixtures achieve optimal values
at different Pmax1 and XR123. And in the following section, in order to
better analyze the effect of operating parameters on cycle perfor-
mance, only the optimal mixtures are discussed. They are MDM/
R123 (0.3/0.7), D4/R123 (0.3/0.7), MD2M/R123 (0.35/0.65), MDM/
R123 (0.68/0.32), D4/R123 (0.6/0.4), and MD2M/R123 (0.55/0.45)
respectively.

6.2. Effect of operating parameters

The operating parameters, such as turbine inlet temperature in


HT cycle (Tmax1), condensation pressure in HT cycle (Pcond1) and
evaporation temperature in LT cycle (Tevap2) have great effects on
the cycle performance. In this paper, the net output power, thermal
efficiency and electricity production cost (EPC) are considered as
the evaluation indexes to select the optimal cycle parameters for
different mixtures.

6.2.1. Effect of turbine inlet temperature in HT cycle (Tmax1)


Fig. 8 illustrates the variation of net power, thermal efficiency
and EPC with turbine inlet temperature in HT cycle (Tmax1).
As can be seen from Fig. 8(a), except MDM/R123 mixtures, the
net power of the other four mixtures increases gradually, and later
flattens. It is noted that the net power of MDM/R123 (0.3/0.7) sud-
denly increases when Tmax1 approaches its maximum value. And Fig. 7. Utilization of heat source (UHS) of the DORC system.
for MDM/R123 (0.68/0.32), the net power reveals a trend of grad-
ual decrease. With Tmax1 increases, the temperature difference
between exhaust gas and mixtures decreases, leading to better
change of the total net power for all mixtures. In detail, if the for-
thermal match and reduced mass flow rate in HT cycle ðm _ f 1 Þ, so
mer falls greater, the total net power also decreases, and vice versa.
W_ net;HT descends. However, the increasing Tmax1 makes more heat
In Fig. 8(b), with Tmax1 rises, thermal efficiency decreases grad-
absorbed in LT cycle, leading to the increase of m _ f 2 . Therefore, ually. Fig. 9 shows the net power and heat absorbed of D4/R123
W_ net;LT is improved. The combined effect of the two causes the (0.3/0.7) under various Tmax1. As can be seen, the total heat
H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26 21

Fig. 9. Net power and heat absorbed for D4/R123 (0.3/0.7).

6.2.2. Effect of condensation pressure in HT cycle (Pcond1)


Fig. 10(a) and (b) describe that with the condensation pressure
in HT cycle (Pcond1) increases, net power and thermal efficiency of
these mixtures decrease almost linearly. This is because a higher
Pcond1 means a larger T5. So enthalpy drop in the turbine decreases,
finally leading to a great decline in W _ net;HT . Moreover, a high tem-
perature after IHE (T6) benefits heat transfer in LT cycle, and makes
a small growth of W _ net;LT . As a result, the net power falls. On the
other hand, with the increasing Pcond1, the total heat absorbed of
this cycle slightly increases. As a result, thermal efficiency is also
decreased.
It is found that the EPC increases almost linearly with the
increase of Pcond1 from Fig. 10(c). The reason is that the reduction
of net power has a greater impact on the EPC, compared to the
decline of system total investment cost.
As a whole, D4/R123 has a better thermodynamic performance
than others while MD2M/R123 possesses the best economic per-
formance. For MD2M/R123, Pcond1 can’t exceed 0.12 MPa to meet
the requirements limited by the pinch point in IHE. To sum up,
high condensation pressure in HT cycle (Pcond1) is unfavorable to
the cycle performance.

6.2.3. Effect of evaporation temperature in LT cycle (Tevap2)


In this part, the effect of evaporation temperature in LT cycle
(Tevap2) is given in Fig. 11. The operating parameters in HT cycle
are fixed at their optimal values.
As Tevap2 rises, the net power of different mixtures changes dif-
ferently. The two mixtures of MDM/R123 are always at the lower
level. When Tevap2 is less than 420 K, the top two is D4/R123
(0.3/0.7) and MD2M/R123 (0.35/0.65). And their net power gradu-
ally decreases with the change of Tevap2. In contrast, the net power
of the D4/R123 (0.6/0.4) and MD2M/R123 (0.55/0.45) increases
quickly. The largest net power is obtained by D4/R123 (0.3/0.7)
Fig. 8. Effect of Tmax1 on net power, thermal efficiency and EPC. at Tevap2 = 400 K, Tmax1 = 633.15 K and the corresponding value is
21.66 kW.
Reasons are as follows: The changes of Tevap2 mainly affect the
absorbed increases linearly, and the increment is greater than that LT cycle. With Tevap2 increases, the heat transfer between exhaust
of net power. Hence, thermal efficiency has such changes. and working fluid in LT cycle gets better. For D4/R123 (0.6/0.4)
With the increase of Tmax1, the EPCs of all the mixtures except and MD2M/R123 (0.55/0.45), the temperature of outlet exhaust
MDM/R123 (0.3/0.7) decrease more and more slowly which can in HT cycle (Texh,out) is higher and the pinch point is always at
be obtained from Fig. 8(c). When Tmax1 approaches its maximum the starting point of the heat transfer (point 9 in Fig. 2), which cor-
value, the EPC of MDM/R123 (0.3/0.7) decreases rapidly due to responds to the final outlet point of exhaust in LT cycle (Texh,final). So
the increase of net power. In addition to, MD2M/R123 (0.35/0.65) the total heat absorbed in LT cycle is large and W _ net;LT increases
provides the best economic performance with smallest EPC, fol- gradually. As a result, the total net power is increasing. For
lowed by D4/R123 (0.3/0.7). And the economic performance of MDM/R123 (0.3/0.7) and D4/R123 (0.3/0.7), Texh,out is smaller and
these two is obviously superior to the other mixtures. the pinch point locates at the evaporation bubble point (point
22 H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26

Fig. 11. Effect of Tevap2 on net power, thermal efficiency and EPC.

Fig. 10. Effect of Pcond1 on net power, thermal efficiency and EPC.

Tmax1 = 633.15 K, followed by MDM/R123 (0.3/0.7), MD2M/R123


10). With Tevap2 increases, Texh,final also increases. So the total heat (0.35/0.65), D4/R123 (0.6/0.4), MD2M/R123 (0.55/0.45) and
absorbed in LT cycle gets rare, causing smaller W _ net;LT . For MDM/ MDM/R123 (0.68/0.32). For all mixtures, the total heat absorbed
R123 (0.68/0.32) and MD2M/R123 (0.35/0.65), Texh,out is middle is decreasing. Compared with D4/R123 (0.3/0.7), D4/R123
and the pinch point lies between point 9 and 10. The total heat (0.6/0.4) has a larger heat transfer amount, so its efficiency is
absorbed decreases within a small range. Therefore, the net power smaller.
of the two has no big change. It is can be seen that the effects of Tevap2 on EPC and on net
As shown in Fig. 11(b), thermal efficiency of all mixtures gets power for each mixture are related through the comparison
larger with Tevap2 increases. D4/R123 (0.3/0.7) stays ahead with between Fig. 11(c) and (a). The increase of net power almost means
the highest thermal efficiency of 22.84% at Tevap2 = 445 K and the decrease of EPC, and vice versa. As shown in Fig. 12, the total
H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26 23

Fig. 12. Effect of Tevap2 on TIC. Fig. 14. Effect of the IHE in LT cycle on TIC.

In addition, the effect of the IHE in LT cycle on TIC is displayed


investment costs of all mixtures increase with the increment of in Fig. 14. It can be observed that the TIC of all the mixtures
Tevap2. For MDM/R123 (0.3/0.7), D4/R123 (0.3/0.7) and MD2M/ increases obviously after adding the IHE in LT cycle. Since the cost
R123 (0.35/0.65), the EPC keeps growing resulted from the increase of pumps and turbines changes a little with adding the IHE, in
of total investment cost as well as the decline of net power. As for this case the TIC mainly depends on the cost of heat exchangers.
D4/R123 (0.6/0.4) and MD2M/R123 (0.55/0.45), although the total Although the cost of Con1 and Con2 both decreases slightly, the
investment cost increases, the raise of net power has a more cost corresponding to the additional IHE in LT cycle has a more
important effect on EPC, leading to the decline of EPC. The EPC of important impact on the TIC, leading to such a change of TIC.
MDM/R123 (0.68/0.32) corresponding to higher Tevap2 remains Noted that, the DORC based on MDM/R123 (0.3/0.7) has the
nearly constant due to the small change of net power. MD2M/ greatest growth rate of TIC, in contrast to the smallest growth
R123 (0.35/0.65) also has the best economic performance with rate for MD2M/R123 (0.55/0.45). Noticeably, no matter whether
the least EPC of 0.603 $/kW h at Tevap2 = 400 K, Tmax1 = 633.15 K. there exists the IHE in LT cycle, MD2M/R123 (0.35/0.65) always
has the lowest TIC, which further validates its advantage on the
6.3. Effect of the IHE in LT cycle economic performance.
In a word, when equipped with the IHE in LT cycle, the DORC
When adding the IHE to LT cycle, the change trends of ther- system acquires an enhancement on thermal efficiency at the
mal efficiency for different mixtures are illustrated in Fig. 13. expense of more TIC. And for the DORC using MD2M/R123
The operating parameters are fixed at their optimal values. As (0.55/0.45), it is most feasible to add the IHE in LT cycle, due to
can be seen, the thermal efficiency for all the mixtures increases the maximum increase rate of thermal efficiency with the mini-
in a different degree. D4/R123 (0.3/0.7) always has the highest mum increase rate of TIC. On the contrary, the feasibility of adding
thermal efficiency. Without the IHE the efficiency is 20.53%, in the IHE for MDM/R123 (0.3/0.7) is lowest.
contrast to 20.97% with the IHE, representing an efficiency
increase of 2.12%. And MD2M/R123 (0.55/0.45) shows the high- 6.4. Exergy analysis
est increase rate of 4.04% while MDM/R123 (0.3/0.7) has the
lowest increase rate of 1.21%. When the LT cycle is equipped 6.4.1. Exergy efficiency
with the IHE, superheated vapor out of the turbine can be used In Fig. 15, exergy efficiencies of all the mixtures are compared at
to preheat the fluid out of the pump. This can recover part of the a wide range of pressure (4–13 MPa). The system exergy efficien-
heat energy and reduce heat load of the condenser in LT cycle. cies reach their respective maximum values at different Pmax1,
It’s the main reason that leads to an improvement in the total and then slowly fall down. Reasons are as follows: Exergy effi-
thermal efficiency. ciency not only depends on the total net output power, but also lies

Fig. 13. Effect of the IHE in LT cycle on thermal efficiency. Fig. 15. Effect of Pmax1 on exergy efficiency.
24 H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26

exergy efficiency of MDM/R123 (0.68/0.32) always keeps at the


minimum value. The optimal exergy efficiencies for the above mix-
tures are: 48.6%, 47.83%, 45.76%, 45.36%, 45.34% and 42.93%. And
the corresponding optimal Pmax1 is mainly focused on the range
of 6–9 MPa. Generally speaking, D4/R123 (0.3/0.7) performs rela-
tively well, and its best exergy efficiency is almost 13.21% higher
than that of MDM/R123 (0.68/0.32).

6.4.2. Exergy destruction


Fig. 16 shows the irreversibility of each component in the DORC
system. The operating parameters are in the optimal values. D4/
R123 (0.3/0.7) has the smallest total exergy destruction of
19.64 kW, followed by MDM/R123 (0.3/0.7), MDM/R123
(0.68/0.32), MD2M/R123 (0.35/0.65), D4/R123 (0.6/0.4) and
MD2M/R123 (0.55/0.45). The corresponding exergy destructions
are: 19.89 kW, 20.29 kW, 20.6 kW, 20.89 kW and 21.28 kW.
Fig. 16. Exergy destructions for different mixtures. Fig. 17 gives the relative contribution of components to total
exergy destruction for different mixtures. In general, for the six
mixtures, exergy destruction of the HT cycle is far more than that
on the exergy variation of the heat and cool source. The total of the LT cycle, making up about three fourths of the total exergy
exergy variation changes little essentially, while the net output destruction. Exergy destructions in the turbine Tur1, IHE and
power for each mixture increases first and then declines. Hence, evaporator Eva1 are always the main causes of exergy
exergy efficiency also has the similar change trend. destruction. Depending on different mixtures, their proportions
D4/R123 (0.3/0.7) always has the highest exergy efficiency fol- of total exergy destruction are 18.73–28.78%, 18.6–26.52% and
lowed by MD2M/R123 (0.35/0.65). The next three are D4/R123 13.55–27.37%, respectively. Exergy destructions occurred in the
(0.6/0.4), MD2M/R123 (0.55/0.45), MDM/R123 (0.3/0.7). The turbine Tur2 and evaporator Eva2 are the next two, and their

Fig. 17. Contribution of components to total exergy destruction for different mixtures.
H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26 25

respective proportions are 6.07–16.17% and 6.41–13.12%. The In most cases, the change of net power has the greatest
maximum proportions of the exergy destructions in the con- impact on EPC with the variation of operating parameters.
denser Con2, Con1 and pump Pum1 are less than 8%. The smallest MD2M/R123 (0.35/0.65) always provides the best eco-
exergy destruction is produced in the pump Pum2 with a tiny nomic performance with the smallest EPC of 0.603 $/
proportion of total exergy destruction. As a whole, the heat kW h at Tevap2 = 400 K, Tmax1 = 633.15 K, followed by D4/
exchange process in the IHE and evaporator as well as the non- R123 (0.3/0.7).
isentropic expansion process in the turbine contribute most to (3) The optimal Pmax1 is at the range of 6–9 MPa for all the mix-
the system exergy destruction. tures considering the exergy efficiency. D4/R123 (0.3/0.7)
performs relatively well, and its best exergy efficiency is
7. Conclusions 48.6%. D4/R123 (0.3/0.7) also has the smallest total exergy
destruction of 19.64 kW. The irreversibility in the IHE, tur-
In this study, the zeotropic siloxanes mixtures used in a dual- bine Tur1 and evaporator Eva1 contributes most to the total
loop organic Rankine cycle are proposed, which consists of a HT exergy destruction, which can serve as the parameter to be
cycle and a LT cycle. Based on the analysis above, some conclusions optimized in the further study.
are gained and listed below: (4) The cycle performance of D4/R123 (0.3/0.7) and MD2M/
R123 (0.35/0.65) is obviously superior to that of the other
(1) For three groups of mixtures, there exist optimal XR123 and mixtures. Considering thermodynamic and economic perfor-
Pmax1 to obtain maximum W _ net and UHS. D4/R123 (0.3/0.7) mance, D4/R123 (0.3/0.7) and MD2M/R123 (0.35/0.65) are
_ net of 21.6 kW at Pmax1 = 10 MPa. the best choices as the working fluid in HT cycle, respec-
obtains the maximum W
tively. In contrast, MDM/R123 (0.68/0.32) is not suitable
At Pmax1 = 11 MPa, D4/R123 (0.6/0.4) has the optimal UHS
due to the poorest performance both on thermodynamics
of 94.81%.
and economy.
(2) In the aspect of operating parameters, the mixtures have
different optimal Tmax1 and Tevap2. Besides, higher Pcond1 is
unfavorable to the cycle performance. D4/R123 (0.3/0.7)
performs relatively better in net power and thermal effi- Acknowledgments
ciency. Under the optimal Tmax1 of 633.15 K, it obtains
the largest net power of 21.66 kW at Tevap2 = 400 K and This work was supported by the National Natural Science
the highest thermal efficiency of 22.84% at Tevap2 = 445 K. Foundation of China (No. 51676133), the authors gratefully
acknowledge it for support of this work.

Appendix A. Some calculation results for D4/R123

Units D4/R123 D4/R123 (0.3/0.7)


XR123 Tmax1/K Pcond1/MPa Tevap2/K
0.5 0.6 0.7 598.15 618.15 633.15 0.10 0.12 0.15 400 420 445
_ f1
m kg/s 0.302 0.304 0.306 0.342 0.306 0.283 0.306 0.308 0.31 0.283 0.283 0.283
_ f2
m kg/s 0.213 0.169 0.123 0.097 0.123 0.140 0.123 0.131 0.142 0.140 0.104 0.041
h1 kJ/kg 16.1 58.9 101.1 101.1 101.1 101.1 101.1 107.5 115.7 101.1 101.1 101.1
h2 kJ/kg 27.4 69.6 111.4 111.4 111.4 111.4 111.4 117.8 126.1 111.4 111.4 111.4
h3 kJ/kg 258.1 269.2 280.7 267.5 280.7 289.8 280.7 285.2 290.9 289.8 289.8 289.8
h4 kJ/kg 475.7 503.0 530.8 500.5 530.8 553.4 530.8 530.8 530.8 553.4 553.4 553.4
h5 kJ/kg 414.8 437.7 461.2 435.7 461.2 480.3 461.2 464.0 467.5 480.3 480.3 480.3
h6 kJ/kg 184.1 238.1 291.8 279.5 291.8 301.8 291.8 296.6 302.7 301.8 301.8 301.8
h7 kJ/kg 235.3 235.3 235.3 235.3 235.3 235.3 235.3 235.3 235.3 235.3 235.3 235.3
h8 kJ/kg 236.3 236.3 236.3 236.3 236.3 236.3 236.3 236.3 236.3 236.3 236.9 237.8
h9 kJ/kg 300.4 296.0 291.5 283.8 291.5 297.2 291.5 296.7 303.5 297.2 297.2 297.3
h10 kJ/kg 337.4 337.4 337.4 337.4 337.4 337.4 337.4 337.4 337.4 337.4 362.9 399.8
h11 kJ/kg 457.7 457.7 457.7 457.7 457.7 457.7 457.7 457.7 457.7 457.7 466.0 471.8
h12 kJ/kg 428.9 428.9 428.9 428.9 428.9 428.9 428.9 428.9 428.9 428.9 432.6 434.2
W_ net kW 20.88 21.26 21.60 21.35 21.60 21.66 21.60 21.03 20.33 21.66 21.07 19.23
_ net;HT
W kW 14.96 16.56 18.20 18.66 18.20 17.78 18.20 17.40 16.40 17.78 17.78 17.78
_ net;LT
W kW 5.92 4.70 3.40 2.69 3.40 3.88 3.40 3.63 3.93 3.88 3.29 1.45
gth % 18.49 19.60 20.82 21.10 20.82 20.53 20.82 20.11 19.24 20.53 21.44 22.84
EPC $/kW h 0.636 0.625 0.618 0.636 0.618 0.611 0.618 0.631 0.648 0.611 0.634 0.717
Ae1 m2 5.427 5.905 6.423 5.887 6.423 6.959 6.423 6.319 6.187 6.959 6.959 6.959
Ai m2 16.746 11.825 8.671 10.773 8.671 6.509 8.671 7.264 7.169 6.509 6.509 6.509
Ac1 m2 4.096 3.062 2.207 1.954 2.207 2.530 2.207 2.415 2.692 2.530 1.915 1.129
Ae2 m2 2.526 2.250 1.895 1.720 1.895 1.973 1.895 1.903 1.889 1.973 1.816 0.993
Ac2 m2 10.218 9.413 7.408 3.941 7.408 7.751 7.408 7.579 7.790 7.751 7.050 2.967
26 H. Tian et al. / Energy Conversion and Management 136 (2017) 11–26

References [23] Zhou Y, Wu Y, Li F, Yu L. Performance analysis of zeotropic mixtures for the


dual-loop system combined with internal combustion engine. Energy Convers
Manage 2016;118:406–14.
[1] Yu G, Shu G, Tian H, Wei H, Liu L. Simulation and thermodynamic analysis of a
[24] Chys M, Broek MVD, Vanslambrouck B, Paepe MD. Potential of zeotropic
bottoming Organic Rankine Cycle (ORC) of diesel engine (DE). Energy 2013;51
mixtures as working fluids in organic Rankine cycles. Energy 2012;44
(9):281–90.
(1):623–32.
[2] Shu G, Yu G, Tian H, Wei H, Liang X. A multi-approach evaluation system (MA-
[25] Song J, Gu C-W. Parametric analysis of a dual loop Organic Rankine Cycle (ORC)
ES) of organic rankine cycles (ORC) used in waste heat utilization. Appl Energy
system for engine waste heat recovery. Energy Convers Manage
2014;132(11):325–38.
2015;105:995–1005.
[3] Shu G, Li X, Tian H, Liang X, Wei H, Wang X. Alkanes as working fluids for high-
[26] Shu G, Liu L, Tian H, Wei H, Liang Y. Analysis of regenerative dual-loop organic
temperature exhaust heat recovery of diesel engine using organic Rankine
Rankine cycles (DORCs) used in engine waste heat recovery. Energy Convers
cycle. Appl Energy 2014;119(15):204–17.
Manage 2013;76(6):234–43.
[4] Fernández FJ, Prieto MM, Suárez I. Thermodynamic analysis of high-
[27] Shu G, Liu L, Tian H, Wei H, Yu G. Parametric and working fluid analysis of a
temperature regenerative organic Rankine cycles using siloxanes as working
dual-loop organic Rankine cycle (DORC) used in engine waste heat recovery.
fluids. Energy 2011;36(8):5239–49.
Appl Energy 2014;113(1):1188–98.
[5] Shu G, Yu G, Tian H, Wei H, Liang X, Huang Z. Multi-approach evaluations of a
[28] Sciubba E, Tocci L, Toro C. Thermodynamic analysis of a Rankine dual loop
cascade-Organic Rankine Cycle (C-ORC) system driven by diesel engine waste
waste thermal energy recovery system. Energy Convers Manage
heat: Part A – thermodynamic evaluations. Energy Convers Manage
2016;122:109–18.
2016;108:579–95.
[29] Vaja I, Gambarotta A. Internal Combustion Engine (ICE) bottoming with
[6] Camporeale SM, Pantaleo AM, Ciliberti PD, Fortunato B. Cycle configuration
Organic Rankine Cycles (ORCs). Energy 2010;35(2):1084–93.
analysis and techno-economic sensitivity of biomass externally fired gas
[30] Colonna P, Guardone A, Nannan NR. Siloxanes: a new class of candidate Bethe-
turbine with bottoming ORC. Energy Convers Manage 2015;105:1239–50.
Zel’dovich-Thompson fluids. Phys Fluids 2007;19(8):75–245.
[7] Tian H, Liu LN, Shu GQ, Wei HQ, Liang XY. Theoretical research on working
[31] Angelino G, Invernizzi C. Cyclic methylsiloxanes as working fluids for space
fluid selection for a high-temperature regenerative transcritical dual-loop
power cycles. J Sol Energy Eng 1993;115(3):130–7.
engine organic Rankine cycle. Energy Convers Manage 2014;86:764–73.
[32] Wilcock DF. Vapor pressure-viscosity relations in methylpolysiloxanes. J Am
[8] Maraver D, Royo J, Lemort V, Quoilin S. Systematic optimization of subcritical
Chem Soc 1946;68(4):691–6.
and transcritical organic Rankine cycles (ORCs) constrained by technical
[33] Shu G, Zhao M, Tian H, Huo Y, Zhu W. Experimental comparison of R123 and
parameters in multiple applications. Appl Energy 2014;117:11–29.
R245fa as working fluids for waste heat recovery from heavy-duty diesel
[9] Yang F, Dong X, Zhang H, Wang Z, Yang K, Zhang J, et al. Performance analysis
engine. Energy 2016;115(Part 1):756–69.
of waste heat recovery with a dual loop organic Rankine cycle (ORC) system for
[34] Babushok VI, Linteris GT, Meier OC, Pagliaro JL. Flame inhibition by CF3CHCl2
diesel engine under various operating conditions. Energy Convers Manage
(HCFC-123). Combust Sci Technol 2014;186(6):792–814.
2014;80:243–55.
[35] Lemmon EW, Huber ML, McLinden MO. NIST standard reference database 23.
[10] Huang H, Zhu J, Yan B. Comparison of the performance of two different Dual-
NIST reference fluid thermodynamic and transport properties—REFPROP,
loop organic Rankine cycles (DORC) with nanofluid for engine waste heat
version, vol. 9; 2010. p. 55.
recovery. Energy Convers Manage 2016;126:99–109.
[36] Zabetakis MG. Flammability characteristics of combustible gases and
[11] Wang T, Zhang Y, Zhang J, Peng Z, Shu G. Comparisons of system benefits and
vapors. Washington DC: Bureau of Mines; 1965.
thermo-economics for exhaust energy recovery applied on a heavy-duty diesel
[37] Garg P, Kumar P, Srinivasan K, et al. Evaluation of isopentane, R-245fa and
engine and a light-duty vehicle gasoline engine. Energy Convers Manage
their mixtures as working fluids for organic Rankine cycles. Appl Therm Eng
2014;84:97–107.
2013;51(1):292–300.
[12] Tian H, Shu G, Wei H, Liang X, Liu L. Fluids and parameters optimization for the
[38] Dai B, Li M, Ma Y. Thermodynamic analysis of carbon dioxide blends with low
organic Rankine cycles (ORCs) used in exhaust heat recovery of Internal
GWP (global warming potential) working fluids-based transcritical Rankine
Combustion Engine (ICE). Energy 2012;47(1):125–36.
cycles for low-grade heat energy recovery. Energy 2014;64(1):942–52.
[13] Xie H, Yang C. Dynamic behavior of Rankine cycle system for waste heat
[39] Heberle F, Preißinger M, Brüggemann D. Zeotropic mixtures as working fluids
recovery of heavy duty diesel engines under driving cycle. Appl Energy
in Organic Rankine Cycles for low-enthalpy geothermal resources. Renew
2013;112(4):130–41.
Energy 2012;37(1):364–70.
[14] Roy JP, Mishra MK, Misra A. Performance analysis of an Organic Rankine Cycle
[40] Turton R, Bailie RC, Whiting WB, et al. Analysis, synthesis and design of
with superheating under different heat source temperature conditions. Appl
chemical processes. Pearson Education; 2008.
Energy 2011;88(9):2995–3004.
[41] Walraven D, Laenen B, D’haeseleer W. Optimum configuration of shell-and-
[15] Oellrich DILR, Leisenheimer B, Srinivasan K. Flammability behavior of
tube heat exchangers for the use in low-temperature organic Rankine cycles.
(octafluoropropane + propane) and (octafluoropropane + methane). Chem Ing
Energy Convers Manage 2014;83:177–87.
Tech 1997;69(7):984–6.
[42] Hartnett JP, Irvine TF. Advances in heat transfer, vol. 6. Academic P; 1970.
[16] Yang Z, Liu B, Zhao H. Experimental study of the inert effect of R134a and
[43] Petukhov BS, Krasnoshchekov EA, Protopopov VS. An investigation of heat
R227ea on explosion limits of the flammable refrigerants. Exp Therm Fluid Sci
transfer to fluids flowing in pipes under supercritical conditions. ASME Int Dev
2004;28(6):557–63.
Heat Transf Part 1961;3:569–78.
[17] Li Z, Gong M, Wu J, Zhou Y. Flammability limits of refrigerant mixtures with
[44] Baskov VL, Kuraeva IV, Protopopov VS. Heat-transfer with turbulent-flow of a
1,1,2,2-tetrafluoroethane. Exp Therm Fluid Sci 2011;35(6):1209–13.
liquid at supercritical pressure in tubes under cooling conditions. High Temp
[18] Kondo S, Takizawa K, Takahashi A, Tokuhashi K. Extended Le Chatelier’s
1977;15(1):81–6.
formula for carbon dioxide dilution effect on flammability limits. J Hazard
[45] Chen JC. Correlation for boiling heat transfer to saturated fluids in convective
Mater 2006;138(1):1–8.
flow. Ind Eng Chem Process Des Dev 1966;5(3):322–9.
[19] Shu G, Long B, Tian H, Wei H, Liang X. Flame temperature theory-based model
[46] Yu W, France D, Wambsganss M, Hull J. Two-phase pressure drop, boiling heat
for evaluation of the flammable zones of hydrocarbon-air-CO2 mixtures. J
transfer, and critical heat flux to water in a small-diameter horizontal tube. Int
Hazard Mater 2015;294:137–44.
J Multiphas Flow 2002;28(6):927–41.
[20] Shu G, Long B, Hua T, Wei H, Liang X. Evaluating upper flammability limit of
[47] Bell KJ, Ghaly MA. An approximate generalized design method for
low hydrocarbon diluted with an inert gas using threshold temperature. Chem
multicomponent/partial condenser. AIChE Symp Ser 1973;69(131):72–9.
Eng Sci 2015;138:810–3.
[48] Chen H, Goswami DY, Rahman MM, Stefanakos EK. A supercritical Rankine
[21] Dong B, Xu G, Cai Y, Li H. Analysis of zeotropic mixtures used in high-
cycle using zeotropic mixture working fluids for the conversion of low-grade
temperature Organic Rankine cycle. Energy Convers Manage 2014;84:253–60.
heat into power. Energy 2011;36(1):549–55.
[22] Liu Q, Duan Y, Yang Z. Effect of condensation temperature glide on the
performance of organic Rankine cycles with zeotropic mixture working fluids.
Appl Energy 2014;115(4):394–404.

You might also like