You are on page 1of 25

Journal Pre-proofs

Working fluid selection of dual-loop organic Rankine cycle using multi-ob-


jective optimization and improved grey relational analysis

Xiao Xia Xia, Zhi Qi Wang, Nai Jun Zhou, Yan Hua Hu, Jian Ping Zhang,
Yin Chen

PII: S1359-4311(19)36360-4
DOI: https://doi.org/10.1016/j.applthermaleng.2020.115028
Reference: ATE 115028

To appear in: Applied Thermal Engineering

Received Date: 12 September 2019


Revised Date: 28 January 2020
Accepted Date: 29 January 2020

Please cite this article as: X. Xia Xia, Z. Qi Wang, N. Jun Zhou, Y. Hua Hu, J. Ping Zhang, Y. Chen, Working
fluid selection of dual-loop organic Rankine cycle using multi-objective optimization and improved grey
relational analysis, Applied Thermal Engineering (2020), doi: https://doi.org/10.1016/j.applthermaleng.
2020.115028

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.


Working fluid selection of dual-loop organic Rankine cycle using

multi-objective optimization and improved grey relational analysis

Xiao Xia Xiaa, Zhi Qi Wanga,*, Nai Jun Zhoub, Yan Hua Hua, Jian Ping Zhanga, Yin Chena
a. Institute of Mechanical Engineering, Xiang Tan University, No.13 North Xiangtan University Road,

Xiangtan 411105, PR China

b. School of Energy Science and Engineering, Central South University, No. 932 South Lu Shan Road,

Changsha 410083, PR China

Abstract: Working fluid selection is still a great challenge for a dual-loop organic Rankine cycle (DORC)

to achieve better comprehensive performance. In this paper, a method combing multi-objective

optimization with improved grey relational analysis (GRA) was proposed to select a suitable fluid pair

for the DORC system, which was driven by the flue gas from a rotary kiln with a temperature of 573.15K.

A multi-objective optimization was conducted for the thermodynamic performance (exergy efficiency,

EXE), economic performance (payback period, PBP), and environmental performance (annual CO2

emission reduction, AER) by using non-dominated sorting genetic algorithm-II (NSGA-II). Then,

priorities of 27 candidate pairs were sequenced based on their grey relational degree. The results show

that cyclohexane/butane has the best comprehensive performance among 27 alternatives, and the EXE,

AER, and PBP are 0.53, 3.43 years, and 14100 tons, respectively. Boiling point temperature is a

criterion of fluid selection for the DORC system. The most suitable boiling point temperature of

working fluid in high-temperature (HT) loop and low-temperature (LT) loop is 330-363K and

255-305K, respectively. This research provides a general methodology to evaluate working fluids of

different ORC systems to recover low-medium temperature waste heat.

Key words: Working fluid selection, Dual-loop organic Rankine cycle, multi-objective optimization,

grey relational analysis

1 Introduction

Recovery of low-medium temperature waste heat, which accounts for more than 50% of the total

waste heat, can lead to reducing fossil fuel consumption and greenhouse gas emissions. As a promising

technology to convert low-medium temperature heat into electricity, the organic Rankine cycle (ORC)

Corresponding author. Wang Zhi Qi Tel.: +86 731 58292215; fax: +86 731 58292210.
E-mail addresses: wangzhiqi@xtu.edu.cn
has attracted considerable attention in the past decades [1, 2]. Compared with the simple ORC, the

DORC system can increase the net output power and heat source utilization rate [3-5].

Working fluid selection is one of the focuses of DORC systems research. Song et al. [6] compared

the net power output of a subcritical DORC using different working fluid pairs, and they concluded that

cyclohexane/R236fa was the best pair for the HT/LT loop. R245fa/R134a was recommended by Wang

et al [7] for a subcritical DORC. When water was used for the HT loop, R143a [8] and R1234yf [9]

were the best fluid for the transcritical and subcritical LT loop, respectively. Considering the

thermodynamic and economic performance of a transcritical DORC system, toluene/R143a was selected

as the desirable fluid pair for the HT/LT loop [10, 11]. Additionally, toluene/R143a was recommended

for the HT/LT loop of a regenerative transcritical DORC to achieve the highest net power output and

exergy efficiency [12]. Wang et al. [13] proposed a regenerative supercritical-subcritical DORC system

and compared system performance using four fluid pairs. The results showed that R1233zd/R1234yf

was the best choice for the HT/LT loop. Although the above research has focused on the working fluid

selection for DORC systems, the performance was compared under fixed conditions. Thus, it is

necessary to perform parameter optimization to ensure candidate working fluids are compared on the

basis of the best performance.

Zhi et al. [14] optimized the inlet temperature and pressure of the HT loop, the evaporation

temperature of the LT loop for a transcritical-subcritical DORC system. The results showed that

R600a/R601a (0.2/0.8) and R134a/R245fa (0.4/0.6) produced the maximum power output. To obtain

better thermodynamic and economic performance, many multi-optimizations have been carried out to

find a suitable working fluid for simple ORCs using different algorithms, such as simulated annealing

algorithm [15], genetic algorithm [16] and NSGA-II [17-19]. As to DORC systems, Yang et al. [20]

conducted a multi-optimization using a genetic algorithm for maximizing net power output and

minimizing total investment cost of DORC systems. It indicated that R245fa/R245fa was the desirable

pair for the HT/LT loop.

In many cases, there is no working fluid that has an optimal performance both on thermodynamic

and economic performance [21]. GRA combines different performance criterion values into a quantified

value of the grey relational degree, which provides a possible approach to the multi-criteria

decision-making problems [22]. On the basis of setting an equal weight coefficient for each indicator,

Wang et al [23] calculated the grey relational degree of 14 candidate fluids. According to the priority of
14 alternatives, R141b was selected as the best working fluid for simple ORCs. Zhang et al. [24]

introduced a fast decision-making method based on the GRA and concluded that R245ca was the most

cost-effective working fluid. However, it was complicated to calculate the weight coefficient by a

universal global optimization.

From the above literature reviews, working fluid selection using multi-objective optimization

mainly focused on the simple ORC system, but less on the complicated DORC system. Additionally, the

thermal and economic indicators were considered in most multi-objective optimization, while the

environmental index was seldom considered. Furthermore, it is necessary to improve the calculation

method of the weight coefficient in the grey relational analysis.

This paper aims to find out a suitable fluid pair and practical selection criteria using

multi-objective optimization and improved GRA for the DORC system. Taking the EXE, AER, and

PBP as the objective function, a multi-objective optimization model is established firstly. Subsequently,

a series of Pareto optimal solutions of each candidate fluid are obtained using the NSGA-II in which the

optimal compromise one is determined by the TOPSIS method. In the next part, we propose an

improved grey correlation analysis, which uses the entropy theory to calculate the weight coefficient.

The EXE, AER, and PBP are combined into a comprehensive indicator using the improved GRA.

Finally, priorities of 27 candidate fluids are sequenced and correlations between fluid properties and

their grey relational degree are explored.

2. Methodology

2.1 System description


The schematic diagram of the DORC system, which is used to recover the waste heat from a rotary

kiln for petroleum coke calcination, is shown in Fig.1.


3
10

Exhaust gas
Expander 1

HT evaporator HT loop

2 Pump 1
1 4
11
Intermediate
12 heat exchanger

7 6
LT evaporator

8 LT loop
Pump 2
Expander 2 5

Condenser
9

13 Cooling water 14

Fig.1 Schematic diagram of DORC system.


As shown in Fig.1, the heat of exhaust gas is absorbed first by the HT loop, and then further

absorbed by the LT loop. Thus, exhaust gas goes through two heat exchanges and releases into the

environment at a lower temperature finally, resulting in more complete utilization. The HT loop consists

of four different components: a pump, a HT evaporator, an expander, and an intermediate heat

exchanger. The LT loop consists of a pump, an intermediate heat exchanger, a LT evaporator, an

expander, and a condenser. These two loops are coupled through the intermediate heat exchanger which

is a condenser of the HT loop and a preheater of the LT loop. The working fluid of the HT loop is

pressurized by the pump1 into the HT evaporator to absorb exhaust waste heat. Then it flows into the

expander1 as superheated vapor and is condensed to saturated liquid in the intermediate heat exchanger.

In the LT loop, the saturated liquid is pressurized into a high-pressure state by the pump2. Then the

subcooled working fluid absorbs the condensation heat of the HT loop and the waste heat of exhaust gas

sequentially. Working fluid becomes a superheated state at the exit of the LT evaporator and enters the

expander2 to generate power. The expanded working fluid is cooled to saturated liquid by cooling water

in the condenser.

The T-s diagrams of the HT loop and the LT loop are shown in Fig.2.
Fig.2 T-s diagram of HT loop and LT loop
The flue gas temperature at the outlet of the rotary kiln is 1260℃. After heat absorption by a waste

heat boiler, it decreases to 573.15K. The parameters of the exhaust are listed in Table 1.

Table 1 Parameters of exhaust from rotary kiln.


Temperature Flow rate Composition %
K kg/s
N2 CO2 CO O2 H2O

573.15 14.1 66.32 24.25 1.66 5.58 2.19

2.2 Mathematic models


2.2.1 Thermodynamic model of DORC system

(1) HT loop

The consumed power by the pump1 in process 1-2 can be expressed as:

Wp1  mfH  h2 s  h1  /  p (1)

The absorbed heat by the HT evaporator in process 2-3 is given by:

QeH  mfH  h3  h2  (2)

mg c p , g T10  T3'  Tp,e1  (3)


mfH 
(h3  h3' )

The generated power by the expander1 in process 3-4 can be determined as:

Wt1  mfH  h3  h4s t (4)

The absorbed heat by the intermediate heat exchanger in process 4-1 is given by:

Qi H  mfH  h4  h1  (5)

(2) LT loop

The consumed power by the pump2 in process 5-6 can be expressed as:
Wp2  mfL  h6  h5  (6)

The absorbed heat by the intermediate heat exchanger in process 6-7 is defined as:

Qi L  mfL  h7  h6   mfH  h4  h1  (7)

mg cp,g T11  T12   mf  h4  h1  (8)


mfL  H

(h8  h6 )

The absorbed heat by the LT evaporator in process 7-8 is given by:

Qe L  mfL  h8  h7  (9)

The generated power by the expander2 in process 8-9 is expressed as:

Wt2  mfL  h8  h9s t (10)

The absorbed heat by the condenser in process 9-5 can be determined as:

Qc L  mfL  h9  h5   m w cw T14  T13  (11)

(3) DORC system

The net output power of the DORC system is calculated by:

  
Wnet  gen Wt1  Wt2  W p1  W p2  Wp  (12)

where ηt, ηp and ηgen is the efficiency of an expander, a pump and an electric generator respectively,

which are set as 0.8, 0.7 [25] and 0.96 [26].

The consumed power of pumps due to the pressure drop of cooling water is defined as:

mcw pcw (13)


Wp 
cw p

According to the second thermodynamic law, EXE can be expressed as:

Wnet (14)
EXE 
I tot  Wnet

The exergy of state j in a DORC system is given by:

E j  m  hj  h0   T0  s j  s0   (15)

The exergy flow rate balance for component j is:

I j  Ein, j  Eout, j  Wj (16)

The total exergy destruction rate is defined as:


I tot   I j    I e  I c  I t  I p  (17)

2.2.2 Economic model

(1) Heat exchanger surface area calculation

The logarithmic mean temperature difference is used to calculate the heat transfer area.

Q (19)
A
K TLMTD
tmax  tmin
TLMTD  (20)
t
ln  max 
 tmax 
1 A  A 1 (21)
 o  w o   Ri
K h f Ai w Am hg fin

A finned tube heat exchanger is employed as the evaporator of the DORC system. The Dias-Young

correlation is used for exhaust gas heat transfer [15].

0.718
  d o Gmax  0.296

 Prg   HY 


1/3
hg  0.1378   (22)
d o   g 

We use the Gnielinski correlation [21] to calculate the heat transfer coefficient of single-phase fluid in a

smooth tube.
(23)
f / 8 Re f Prf
Nu f 
12.7 f / 8  Prf 2/3  1  1.07

(24)
f  (1.82 lg Re f  1.64) 2

 Nu
h
d (25)
For the evaporating process, Varme Atlas correlation [21] is used to evaluate the heat transfer at

the two-phase region.

2.2 0.5
2

     "    (26)
 f   '  (1  x)  1.2 x 0.4 (1  x)( l )0.37    ' x 0.01 (1  8(1  x)0.7 )( l )0.67  
  g    g  

A shell-and-tube heat exchanger is chosen for the condenser in this work. The Dittus-Boelter

correlation [19] is employed to calculate the heat transfer coefficient of cooling water.

w
hw,i  0.023 Rew 0.8 Prw 0.4 (27)
di

The single-phase heat transfer coefficient of the shell side is calculated using the McAdams
correlation [27].
0.14
 
Nu  0.36 Re 0.55 1/3
Pr  b  (28)
 w 
For the condensation process, the heat transfer coefficient of the two-phase working fluid is

calculated by [15]:
0.25
 f3          gr 
hf,o  0.725   (29)
 f d o Tc  Tb  
(2) Investment calculation

The capital cost of components and the cost of operation and management for the DORC system

are considered in the economic model. The equipment purchase cost of heat exchangers, including an

evaporator, an intermediate heat exchanger and a condenser, can be written as [28]:

lg CHe  F1, He  F2, He lg AHe  F3, He  lg AHe 


2
(30)

The capital cost of the expander and pump is expressed as:

lg Ct , p  F1,t , p  F2,t , p lg Wt , p  F3,t , p  lg Wt , p 


2
(31)

Besides, the generator cost is calculated as:

lg Cgen  1850000(Wnet /11800)0.94 (32)

When the equipment type, system pressure and construction materials deviate from the base

conditions, the capital cost is corrected by the coefficient Fbm.

Cbm  FbmC0 (33)

Fbm  B1  B2 Fm Fp (34)

lg Fp  C1  C2 lg p  C3 (lg p ) 2 (35)

The Chemical Engineering Plant Cost Index (CEPCI) is used to assess the impact of inflation on

investment cost, and the final capital cost of a DORC in the year 2017 is calculated as follows:

CI ,2017 (36)
Cbm,2017  Cbm,2001
CI ,2001

where CI,2001 is set as 397[29], CI,2017 is set as 623.5[30].


The total cost of all equipment can be calculated by:

Ctot  Cbm,He  Cbm,t  Cbm,p  Cbm,ge (37)


Coefficients of cost evaluation of different components are listed in Table 2.
Table 2 Coefficients of cost evaluation [28].
Coefficients Turbine Pump Evaporator Condenser Generator
F1 2.247 3.3892 4.3247 4.033 --
F2 1.49 0.0536 -0.303 0.23 --
F3 -0.1618 0.1538 0.1634 0.0497 --
B1 -- 1.89 1.63 1.96 --
B2 -- 1.35 1.66 1.21 --
C1 -- -0.3935 0.03881 -- --
C2 -- 0.3957 -0.11272 -- --
C3 c -0.002 0.08183 -- --
F -- 1.5 1 1 --
FB 3.5 -- -- -- 1.5

The PBP is obtained as:

ln(1  kCtot / Cn0 ) (38)


PBP  
ln(1  k )

Fn0  CpriWnet tops  f k Ctot (39)

where k is the interest rate with a value of 5% [31] and fk is the maintenance and operation cost factor

with a value of 1.5% [15]. CPri is the electricity price which is set as 0.10$/kWh and top is the full load

operation time which is set as 8000 h [28].

2.2.3 Environmental model


Since greenhouse gas like CO2, CO, CH4, and NOx is a major reason causing global warming,

there is a growing concern about the reduction of greenhouse gas emissions. During the process of

operation, the DORC system converts the waste heat to power. Thus, its operation does not increase

additional greenhouse gas emissions. In this paper, the greenhouse gases emitted by traditional

coal-fired power plants in China are employed as the calculation benchmark of the DORC system.

According to a conversion factor which is used to evaluate the contribution of different gases to the

greenhouse effect, a total reduction of CO2 emission can be calculated by:


r
MCO2
 Wnet topgen (CFCO2 mCO2  CFCO mCO  CFCH4 mCH4  CFNOx mNOx ) (40)

On the other hand, the DORC system is made of stainless steel. Therefore, emissions from steel

production should be considered in evaluating the environmental impact of DORC systems. According

to Ref. [32], the consumption of stainless steel for the expander is set as 1.2 kg per 1 kW output power.

The required steel for the production of pipes, valves, and pumps is negligible. The mass of heat
exchangers is calculated according to the length and thickness of the tube.
  L
 do  2H   do2  
2
M he   (do2  di2 ) L    (41)
4 4  Y
A total CO2 emission by steel consumption of the DORC system is defined as:
em
MCO2
  M he  Mt  (FCO2 mCO
em
2
 FCO mCO
em
 FCH4 mCH
em
4
 FNOx mNOx
em
) (42)

The conversion factors, emissions of electricity and steel production, are listed in Table 3.

Table 3 Parameters in the environmental model [33]


Emission Conversion Traditional coal-fired Emissions of steel
factor power plant (g/kWh) production (g/kg)
CO2 1 877 410
CO 2 1.25 5.5
CH4 25 2.65 0.9
NOx 320 6.34 0.8

The AER is defined as:


em
M CO (43)
AER  M r
CO2  2

Slt
where Slt is the plant lifetime and set as 20 years[32].
3 Multi-objective optimization
3.1 Optimization model and algorithm

Three objectives of maximizing the EXE and AER, and minimizing the PBP of the DORC system

are considered in the proposed multi-objective optimization. Six decision parameters are selected for the

optimization of the DORC system: Pe1: Evaporation pressure of the HT cycle, Pc1: Condensation

pressure of the HT cycle, Pc2: Condensation pressure of the LT loop, Δ Te1: Pinch point temperature of

the HT evaporator, Δ Te2: Pinch point temperature of the intermediate heat exchanger, Δ Te3: Pinch

point temperature of the LT condenser.

The multi-objective optimization model of the DORC system can be expressed as:

max F1  pe1 , pc1 , pc 2 , Te1 , Te 2 , Tc   EXE (44)



min F2  pe1 , pc1 , pc 2 , Te1 , Te 2 , Tc   PBP

max F3  pe1 , pc1 , pc 2 , Te1 , Te 2 , Tc   AER
Based on the energy balance and the definition of pinch point temperature difference, the following

constraints are considered in the optimization.


T10  T3  Te1 , T11  T2  Te1 , T1  T7  Te 2 , T5  T13  Tc , T12  T7  Te 2

Te1  5, Te 2  5, Tc  5 , pe1  pc1  0.101 , pc 2  0.101, Tc1  Tc 2  T0 , T12  373.1  0 (45)
Since different objectives often conflict with each other in a multi-objective optimization problem,

there is no optimal solution that is best for all objective functions. Instead, there exists a set of optimal

solutions, well-known as Pareto optimal solutions [33]. NSGA-II has been widely used to solve

multi-objective optimization problems. It can search for all Pareto optimal solutions and includes

diversity and better elitism. Therefore, NSGA-II is employed to obtain the Pareto optimal solutions of

multi-objective optimization. Details about the procedure of NSGA-II can be found in Ref. [17, 18].

3.2 Optimal solution

The TOPSIS method is employed to select an optimal solution from the Pareto frontier. The final

compromise one in TOPSIS is the solution with a longer distance from the non-ideal solution is

preferred over that with a shorter distance from the ideal solution. The distance from the ideal solution

is expressed as:

m
 Fij 
2 (46)
Di    Fijideal
j 1

The distance from non-ideal solution is expressed as:


m
  Fij 
2 (47)
Di    Fijn ideal
j 1

The ratio coefficient can be calculated as:


Di 
Cc  (48)
Di   Di 

Finally, the solution with the maximum Cc is selected as the optimal solution.

3.3 Algorithm validation

The same operating parameters and energy conservation equations as the reference [34] are used to

verify the accuracy of the algorithm. Taking the exergy efficiency and heat exchanger area per unit of

net power output as the objective function, the population is 40 and the generation size is 100. The

comparison between the Pareto frontier solution and the Fig.3 of the reference [34] shows that the trend

of the curve is consistent.


Fig.3 Pareto frontier of the basic ORC for APR with the exergy efficiency.

4. Decision-making method for working fluid selection

In the present work, an improved GRA is proposed to combine EXE, AER, and PBP into a
quantified indicator of grey relational degree. It provides a possible way to evaluate the DORC system
performance in terms of thermodynamics, economy, and environment.
According to GRA, the relational coefficient between the evaluation sequence and the ideal
optimal sequence is calculated by:
 min   max (49)
crij 
 i  j    max

 
 max  max i  max j Poj  Pij 
  (50)

 
 min  min i  min j Poj  Pij 
  (51)

i  j   Poj  Pij (52)

The grey relational degree of different working fluids is defined as:


Ri   j 1 w j  crij  i  1, 2, , n 
m
(53)

The calculation of the weight coefficient for each objective function is the key part of the grey
relational analysis. In the improved GRA, a weight coefficient is calculated based on the entropy of the
indicator.
According to the entropy theory, the index value proportion of EXE and ARE is defined as:
xij (54)
Pij 
 i 1 xij
n

While the proportion of PBP is expressed as:


1 (55)
xij
Pij 
 i 11 xij
n
The entropy value and weight coefficient of each indicator are calculated as follows:


n
P ln Pij (56)
i 1 ij
EV j 
ln n

1  EV j (57)
wj 
 1  EV 
m
j 1 j

After multi-objective optimization, the fluid that has the highest grey relational degree is the best
choice. The process of working fluid selection is shown in Fig.4.
Input initial parameters

Multi-objective
optimization model

NSGA-II algorithm
Pareto solution
New fluid
Topsis method
Optimal solution

All fluid is No
optimized

Yes

Indicator value for


different fluid
Entropthy method

Weight coefficient
Grey relational
analysis
Grey relational grade
of all selected fluids

Best working fluid

Fig.4 Flow chart of working fluid selection


The multi-objective optimization is conducted using the NSGA-II program written in Matlab and

properties of working fluids are obtained from REFPROP. During the optimization process, the parameters

of the NSGA-II are as follows: population size 120, generation size 1000, and crossover fraction 0.9.

Besides, working fluid is assumed to be saturated liquid at the exit of the condenser.

5. Results and discussion


Previous studies have shown that working fluids of the DORC system can be divided into two

categories. In one case, the HT loop and LT loop use the same working fluid to reduce the complexity

of DORC systems, such as R600 and R245fa in Ref [35-37]. In the other case, the HT loop and LT loop

use different working fluids, such as R245fa/R134a, toluene/R143a in Ref [7, 10, 11]. Therefore, we

evaluate and compare the comprehensive performance of a DORC system in the two situations.
5.1 system performance using the same working fluid
In present study, we select nine working fluids for both the HT loop and the LT loop. The boiling

temperature of these working fluids locates in the ranges of 247 ~398 K. The thermodynamic properties

of these working fluids are listed in Table 4.

Table 4 Thermodynamic properties of selected working fluids.


M Pcri Tcri Tnbp
fluid GWP ODP Ref.
/g·mol-1 /kPa /K /K

R134a 102.03 4.05 374.2 247.1 1430 0 [38]

Butane 58.12 3.80 425 272.7 4 0 [38]

R245fa 134.05 3.65 427.21 288.3 1030 0 [38]

R141b 116.95 4.21 477.5 305.2 725 0.12 [38]

Cyclopentane 70.13 4.57 511.7 322.4 very low 0 [39]

Cyclohexane 70.13 4.07 553.6 353.8 very low 0 [39]

Benzene 78.11 4.90 561.9 353.1 very low 0 [6]

Toluene 92.14 4.12 591.7 383.75 3.3 0 [6]


Octane 114.2 2.49 569.3 398.77 very low 0 [39]

Pareto frontiers and best solutions for different working fluids are shown in Fig. 5 and Table 5,

respectively. Figure 5 shows that Pareto fronts of different fluids almost do not overlap, which indicates

that the system performance is quite different. It can be seen from Table 5 that R141b/R141b is the

optimal fluid pair, which has the highest EXE and AER, and the lowest PBP. Its EXE, AER, and PBP

are 0.44, 3.49 years and 12000 tons, respectively. In addition to R141b/R141b, R245fa/R245fa and

cyclopentane/cyclopentane are also suitable fluid pairs for the DORC system.
12
benzene benzene
1.4
R245fa R245fa
R141b R141b
10 butane butane
1.2
octane octane
toluent toluent
R134a 1.0 R134a
8
AER /104 t

cyclopentane
PBP /y

cyclopentane
cyclohexane cyclohexane
0.8
6
0.6

4
0.4

0.20 0.25 0.30 0.35 0.40 0.45 0.50 4 6 8 10 12


Exergy efficiency PBP /y

(a)exergy efficiency and PBP (b) PBP and ARE


Fig.5 Pareto frontier results of DORC system with the same working fluid.
Table 5 Optimal parameters and performance of DORC system using the same working fluid
pe1 pc1 pc2 ∆Te1 ∆Te2 ∆Tc1 PBP AER
Fluid pairs EXE
/MPa /MPa /MPa /K /K /K /y /104t
R134a/R134a 3.7 1.10 0.626 84.0 7.1 6.8 0.31 5.73 0.78

Butane/Butane 3.5 0.735 0.226 88.0 8.5 6.0 0.41 3.95 1.09

R245fa/R245fa 3.59 0.40 0.17 91.0 7.9 6.9 0.46 3.88 1.16

R141b/R141b 3.9 0.387 0.101 50.3 7.7 10.0 0.44 3.49 1.20

Cyclopentane/Cyclopentane 3.91 0.37 0.101 30.1 7.6 16.5 0.43 3.54 1.13

Cyclohexane/Cyclohexane 2.2 0.44 0.101 14.2 8.5 32.1 0.35 4.86 0.8

Benzene/Benzene 2.3 0.53 0.101 9.6 8.7 32.3 0.37 4.48 0.83

Toluene/Toluene 0.29 6.15 0.57


1.37 0.42 0.101 10.2 8.8 5.0
Octane/Octane 0.24 9.0 0.42
1.20 0.4 0.101 13.1 4.4 9.1

5.2 system performance using different working fluid


In this part, five organic fluids with a high decomposition temperature, including cyclopentane,

cyclohexane, benzene, toluene, and octane, are selected as candidate fluids for the HT loop. While these

working fluids with low boiling temperature, including R245fa, R600, R134a, and R141b, are used for

the LT cycle. Eighteen pairs of working fluids are obtained by combining fluid of the HT loop with that

of the LT loop.

Based on the multi-objective optimization of the DORC system, Pareto frontiers and best solutions

for all fluid pairs are shown in Fig. 6 and Table 6, respectively.

Cyclohexane/Butane
Cyclopentane/R134a
Cyclohexane/R141b
Cyclopentane/R141b
4.5 Cyclohexane/R134a
Cyclopentane/Butane
Cyclohexane/R245fa 3.6 Cyclopentane/R245fa
Toluene/Butane
Benzene/R134a
Toluene/R245fa
Benzene/butane
Toluene/R141b
Benzene/R245fa
4.0 Octane/R141b
Benzene/R141b
Octane/R245fa 3.4
PBP /y

PBP /y

Octane/Butane

3.5
3.2

3.0 3.0
0.45 0.50 0.55 0.60 0.48 0.51 0.54 0.57
Exergy efficiency Exergy efficiency

(a)exergy efficiency and PBP


1.6 1.5
cyclohexane/butane cyclopentane/butane
cyclohexane/R141b cyclopentane/R141b
cyclohexane/R134a cyclopentane/R134a
1.5
cyclohexane/R245fa cyclopentane/R245fa
toluent/ butane benzene butane
toluent/ R245fa benzene/ R245fa
1.4 toluent/ R141b 1.4 benzene/ R141b
octane/R141b bemzene/R134a
AER /104 t

AER /104 t
octane/R245fa
1.3 octane/butane

1.2 1.3

1.1

1.0 1.2
3.0 3.5 4.0 4.5 5.0 3.1 3.2 3.3 3.4 3.5 3.6
PBP /y PBP /y

(a) AER and PBP


Fig. 6 Pareto frontier results of DORC system with different working fluids.

Table 6 Optimal solutions and performance of DORC system using different working fluids
pe1 pc1 pc2 ∆Te1 ∆Te2 ∆Tc1 PBP AER
Fluid pairs EXE
/MPa /MPa /MPa /K /K /K /y /104t

Cyclopentane/R134a 3.9 0.17 0.665 30 9.5 7.0 0.49 3.26 1.29

Cyclopentane/Butane 2.25 0.15 0.232 30 8.4 6.8 0.50 3.26 1.33

Cyclopentane/R245fa 2.38 0.15 0.141 30 7.8 7.0 0.50 3.24 1.33

Cyclopentane/R141b 3.9 0.27 0.101 30 8.3 10.0 0.48 3.21 1.28

Cyclohexane/R134a 1.53 0.105 0.64 14.3 8.6 7.1 0.51 3.38 1.34

Cyclohexane/Butane 1.46 0.101 0.224 14.8 8.8 6.5 0.53 3.43 1.41

Cyclohexane/ R245fa 1.44 0.101 0.142 13.7 8.2 7.2 0.53 3.41 1.39

Cyclohexane/ R141b 1.48 0.101 0.101 13.7 7.8 9.9 0.51 3.31 1.34

Benzene /R134a 1.40 0.101 0.65 11.3 9.6 7.5 0.51 3.25 1.36

Benzene /Butane 1.47 0.103 0.234 10.6 9.7 7.2 0.52 3.31 1.39

Benzene /R245fa 1.49 0.101 0.145 10.6 9.9 7.2 0.52 3.28 1.38

Benzene/R141b 1.55 0.109 0.101 10.7 8.4 10.1 0.51 3.2 1.35

Toluene /Butane 0.904 0.101 0.232 12.2 11.8 7.1 0.5 3.49 1.28

Toluene / R245fa 0.987 0.108 0.144 13.2 12.3 7.1 0.49 3.47 1.25

Toluene / R141b 0.881 0.101 0.101 12.7 10.9 10.5 0.50 3.27 1.27

Octane /Butane 1.06 0.105 0.232 16.7 5.1 7.0 0.48 3.8 1.16

Octane / R245fa 0.846 0.101 0.141 16.6 9.9 7.3 0.48 3.75 1.18

Octane /R141b 0.915 0.104 0.101 16.3 10.1 10.4 0.49 3.51 1.18

Fig. 6 shows that the Pareto front of 18 working fluid pairs intersects with each other. In this case,

no fluid pair can obtain the optimal EXE, AER, and PBP simultaneously. It is quite different from the

results of the same fluid used in the HT loop and LT loop. Table 6 shows that Cyclohexane /Butane and
Cyclohexane/R245fa have the highest exergy efficiency, while Benzene/R141b has the shortest PBP. To

solve this multi-objective decision-making problem, an improved GRA is employed to analyze the

optimization result. Pareto optimal solutions of 27 working fluid pairs are processed, and the results are

presented in Table 7.
Table 7 Grey relational degree of different working fluid pairs.
Fluid pairs R Fluid pairs R
Cyclohexane/Butane 0.874 R141b/R141b 0.566

Cyclohexane/ R245fa 0.839 Toluene / R245fa 0.559

Benzene /R245fa 0.836 Octane /R141b 0.518

Benzene /Butane 0.835 Octane / R245fa 0.459

Benzene/R141b 0.821 Octane/Butane 0.446

Benzene /R134a 0.793 Cyclopentane/Cyclopentane 0.424

Cyclopentane/R245fa 0.741 R245fa/R245fa 0.391

Cyclohexane/ R141b 0.734 Butane/Butane 0.344

Cyclopentane/Butane 0.727 Benzene/ Benzene 0.255

Cyclohexane/R134a 0.703 Cyclohexane/ Cyclohexane 0.233

Cyclopentane/R141b 0.686 R134a/ R134a 0.214

Toluene / R141b 0.675 Toluene/ Toluene 0.182

Cyclopentane/R134a 0.673 Octane/ Octane 0.151

Toluene /Butane 0.593

As can be seen from Table 7, Cyclohexane/Butane has the largest grey relational degree of 0.874.

Therefore, it is the best choice among 27 candidate pairs to achieve the optimal comprehensive

performance. Compared with R141b/R141b, the EXE, AER, and PBP of Cyclohexane/Butane increase

by 20.5%, 17.5%, and -1.7%, respectively. Besides, Cyclohexane/ R245fa, Benzene /R245fa, Benzene

/Butane, and Benzene /R134a are also suitable fluid pairs for the HT/LT loop since their grey relational

degree is larger than 0.8. By contrast, the grey correlation degree of Toluene/Toluene and Octane/

Octane are lower than 0.2, which is not suitable for the HT/LT loop. It indicates that different fluids

used in the HT/LT loop are more beneficial to the comprehensive performance of the DORC system.

In order to quickly and accurately select several suitable pairs from a large number of working

fluids, we explore correlations between the fluid properties and the grey relational degree, as shown in

Figure 7.
400

Boiling temperature of working fluids in LT loop /K


380
0.2302

360 0.2302
0.3104

340 0.3907
0.4709
320 0.5511

0.7918
300

0.6313 0.7116
280 0.3907
0.4709
0.4709
260 0.3104
0.5511

260 280 300 320 340 360 380 400


Boiling temperature of working fluids in HT loop /K

Fig.7 The relationship between boiling point temperature and grey relational degree.

Fig.7 shows that the boiling point temperature is an important criterion for the fluid selection of

DORC systems. When the boiling point temperature of working fluids in the HT loop is higher than that

in the LT loop, the DORC system can obtain better performance. The optimal boiling point temperature

of fluids in the HT loop and LT loop locates in the ranges of 330-363K and 255-304K, respectively.

6 Conclusions
A multi-objective optimization is conducted using NSGA-II for the DORC system, which is driven

by the waste heat from a rotary kiln. An improved GRA is proposed to evaluate the DORC system

performance in terms of thermodynamics, economy, and environment. The optimal working fluid pair

and fluid selection criteria for DORC system are presented. The conclusions are as follows:

(1) Among the 9 selected candidate pairs, which uses the same fluid in the HT loop and LT loop,

R141b/R141b is the optimal pair.

(2) When different working fluids are used for the HT loop and LT loop respectively,

cyclohexane/butane is the best choice among 18 selected candidate pairs. Compared with R141b/R141b,

the EXE, AER, and PBP of cyclohexane/butane increases by 20.5%, 17.5%, and -1.7%, respectively.

(3) Boiling temperature is a criterion of fluid selection for DORC systems. When the heat source

temperature is 573.15K, the suitable boiling point temperature range of working fluid in the HT loop is

330-363K, and that of working fluid in the LT loop is 255-304K.

This work provides a general methodology to evaluate working fluids for ORC systems. Based on

this method, we intend to perform a fluid selection for the DORC system using mixtures in the future.
Acknowledgments
This work was supported by the Natural Science Foundation of Hunan Province, China

(Grant No. 2018JJ2399), the National Natural Science Foundation of China (Grant No. 51975503) and

the Science and Technology Major Project of Tibet of China (Grant No.XZ201801-GA-03).

References
[1] K. Rahbar, S. Mahmoud, R. K. Al-Dadah, et al, Review of organic Rankine cycle for small-scale
applications, Energy Convers. Manage. 34(2017) 135-55.
[2] A. Mahmoudi, M. Fazli, M. R. Mora, A recent review of waste heat recovery by Organic Rankine Cycle,
Appl. Therm. Eng. 143(2018) 660-675.
[3] H. G. Zhang, E. H. Wang, B. Y. Fan, A performance analysis of a novel system of a dual loop bottoming
organic Rankine cycle (ORC) with a light-duty diesel engine, Appl. Energy 102 (2013) 1504-1513.
[4] L. F. Shi, G. Q. Shu, H. Tian, et al., A review of modified Organic Rankine cycles (ORCs) for internal
combustion engine waste heat recovery (ICE-WHR), Renew. Sust. Energ. Rev. 92 (2018) 95-110.
[5] G. Q. Shu, L. N. Liu, H. Tian, et al., Analysis of regenerative dual-loop organic Rankine cycles (DORCs)
used in engine waste heat recovery, Energy Convers. Manage. 76 (2013) 234-243.
[6] J. Song, C. W. Gu, Parametric analysis of a dual loop Organic Rankine Cycle (ORC) system for
engine waste heat recovery, Energy Convers. Manag. 105 (2015) 995-1005.
[7] E. H. Wang, H. G. Zhang, Y. Zhao, et al., Performance analysis of a novel system combining a dual
loop organic Rankine cycle (ORC) with a gasoline engine, Energy 43 (2012) 385-95.
[8] G. Q. Shu, L. N. Liu, H. Tian, et al., Parametric and working fluid analysis of a dual loop organic
Rankine cycle (DORC) used in engine waste heat recovery, Appl. Energy 113 (2014) 1188-98.
[9] G. Q. Shu, L. N. Liu, H. Tian, et al., Performance comparison and working fluid analysis of subcritical
and transcritical dual-loop organic Rankine cycle (DORC) used in engine waste heat recovery, Energy
Convers. Manag. 74 (2013) 35-43.
[10] G. Q. Shu, G. P. Yu, H. Tian, et al., Multi-approach evaluations of a cascade-Organic Rankine
Cycle (C-ORC) system driven by diesel engine waste heat: Part A -Thermodynamic evaluations,
Energy Convers. Manag. 108 (2016) 579-595.
[11] F. Mohammadkhani, M. Yari. A 0D model for diesel engine simulation and employing a
transcritical dual loop Organic Rankine Cycle (ORC) for waste heat recovery from its exhaust and
coolant: Thermodynamic and economic analysis, Appl. Therm. Eng. 150 (2019) 329-347.
[12] H. Tian, L. N. Liu, G. Q. Shu, et al., Theoretical research on working fluid selection for a
high-temperature regenerative transcritical dual-loop engine organic Rankine cycle, Energy Convers.
Manag. 86 (2014) 764-73.
[13] E. H. Wang, Z. Yu, H. G. Zhang, et al., A regenerative supercritical-subcritical dual-loop organic
Rankine cycle system for energy recovery from the waste heat of internal combustion engines. Appl.
Energy, 190 (2017) 574-90.
[14] L. H. Zhi, P. Hu, L. X. Chen, et al., Parametric analysis and optimization of transcritical-subcritical
dual-loop organic Rankine cycle using zeotropic mixtures for engine waste heat recovery, Energy
Convers. Manag. 195 (2019) 770-787.
[15] Z. Q. Wang, N. J. Zhou, J. Guo, et al., Fluid selection and parametric optimization of organic
Rankine cycle using low temperature waste heat, Energy, 40 (2012) 107-115.
[16] F. Yang , H. Zhang , S. Song, et al., Thermoeconomic multi-objective optimization of an organic
Rankine cycle for exhaust waste heat recovery of a diesel engine, Energy 93 (2015) 2208-2228.
[17] J. F. Wang, M. Wang, M. Q. Li, et al., Multi-objective optimization design of condenser in an organic
Rankine cycle for low grade waste heat recovery using evolutionary algorithm, Int. Commun. Heat mass
45 (2013) 47-54.
[18] M. Imran, M. Usman, B. S. Park, et al., Multi-objective optimization of evaporator of organic Rankine
cycle (ORC) for low temperature geothermal heat source, Appl. Therm. Eng. 80 (2015) 1-9
[19] Y. Q. Feng, T. C. Hung, Y. N. Zhang, et al., Performance comparison of low-grade ORCs (organic
Rankine cycles) using R245fa, pentane and their mixtures based on the thermo-economic multi-objective
optimization and decision makings, Energy 93 (2015) 2018-2029.
[20] F. B. Yang, H. Cho, H. G. Zhang, et al., Thermoeconomic multi-objective optimization of a dual loop
organic Rankine cycle (ORC) for CNG engine waste heat recovery, Appl. Energy 205 (2017) 1100-1118.
[21] P. Linke, A. I. Papadopoulos, P. Seferlis, Systematic methods for working fluid selection and the design,
integration and control of organic Rankine cycles-A review, Energies 8(6) (2015) 4755-4801.
[22] X. X. Xia, Z. Q. Wang, Y. H. Hu, et al., A novel comprehensive evaluation methodology of organic
Rankine cycle for parameters design and working fluid selection, Appl. Therm. Eng. 143 (2018) 283-292.
[23] Y. Z. Wang, J. Zhao, Y. Wang, et al., Multi-objective optimization and grey relational analysis on
configurations of organic Rankine cycle, Appl. Therm. Eng. 114 (2017) 1355-1363.
[24] X. Zhang, H. Bai, X. C. Zhao, et al., Multi-objective optimization and fast decision-making method for
working fluid selection in organic Rankine cycle with low-temperature waste heat source in industry, Energy
Convers. Manag. 172 (2018) 200-211.
[25] L. Xiao, S. Y. Wu, T. T. Yi, et al., Multi-objective optimization of evaporation and condensation
temperatures for subcritical organic Rankine cycle, Energy 83 (2015) 723-733.
[26] A. Toffolo, A. Lazzaretto, G. Manente, et al. A multi-criteria approach for the optimal selection of
working fluid and design parameters in organic Rankine cycle systems, Appl. Energy 121(2014)
219-232.
[27] A. Erdogan, C. O. Colpan, D. M.Cakici, Thermal design and analysis of a shell and tube heat exchanger
integrating a geothermal based organic Rankine cycle and parabolic trough solar collectors, Renew. Energy
109 (2017) 372-391.
[28] X. Q. Wang, X. P. Li, Y. R. Li, et al. Payback period estimation and parameter optimization of
subcritical organic Rankine cycle system for waste heat recovery, Energy 88 (2015) 734-745.
[29] G. Yu , G. Q. Shu, H. Tian, et al., Multi-approach evaluations of a cascade-Organic Rankine Cycle
(C-ORC) system driven by diesel engine waste heat: Part B-techno-economic evaluations, Energy
Convers. Manag. 108 (2016) 596-608.
[30] D Mignard, Correlating the chemical engineering plant cost index with macro-economic indicators,
Chem. Eng. Res. Des. 92 (2) (2014) 285-294.
[31] G. Q. Shu , G. Yu, H. Tian, et al., A Multi-Approach Evaluation System (MA-ES) of Organic Rankine
Cycles (ORC) used in waste heat utilization, Appl. Energy 132 (2014) 325-338.
[32] C. Liu, C. He, H. Gao, et al., The environmental impact of organic Rankine cycle for waste heat
recovery through life-cycle assessment, Energy 56 (2013) 144-154.
[33] X. Di, Z. Nie, B. Yuan, et al., Life cycle inventory for electricity generation in China. Int. J. Life
Cycle Ass. 12 (2007) 217-224.
[34] Y. Feng, Y. Zhang, B. Li, et al., Sensitivity analysis and thermoeconomic comparison of ORCs
(organic Rankine cycles) for low temperature waste heat recovery, Energy 82 (2015) 664-677.
[35] F. B. Yang, X. R. Dong, H. G. Zhang, et al., Performance analysis of waste heat recovery with a
dual loop organic Rankine cycle (ORC) system for diesel engine under various operating conditions.
Energy Convers. Manag. 80 (2014) 243-255.
[36] F. B. Yang, H. G. Zhang, Z. Yu, et al., Parametric optimization and heat transfer analysis of a dual loop
ORC (organic Rankine cycle) system for CNG engine waste heat recovery, Energy 118 (2017) 753-775.
[37] B. F. Yao, F. B. Yang, H. G. Zhang, et al., Analyzing the Performance of a dual loop organic
Rankine cycle system for waste heat recovery of a heavy-duty compressed natural gas engine, Energies,
7(11) (2014) 7794-7815.
[38] X. C. Wang, E. K. Levy, C. J. Pan, et al. Working fluid selection for organic Rankine cycle power
generation using hot produced supercritical CO2 from a geothermal reservoir, Appl. Therm. Eng. 149
(2019) 1287-1304.
[39] G. Q. Shu, X. N. Li, H. Tian, et al., Alkanes as working fluids for high-temperature exhaust heat
recovery of diesel engine using organic Rankine cycle, Appl. Energy 119 (2014)204-217.

Nomenclature
HT high temperature DORC Dual-loop organic Rankine cycle
LT low temperature CEPCI Chemical Engineering Plant Cost Index
PBP payback period NSGA-II Non-dominated sorting genetic
AER annual emission reduction GRA algorithm
Grey relational analysis
W power, kW m mass flow rate, kg·s-1
h specific enthalpy, J·kg-1 cp specific heat,kJ·kg-1·K-1
T temperature, K ∆T temperature difference, K
Q heat flow rate, W E exergy, W
s entropy, J·kg-1·K-1 C cost,$

F function A heat transfer surface area,m2


B buy factor t time, h
K heat transfer coefficient, W·m-2·K-1 f cost related to the operation

k interest rate ∆p pressure drop, Pa


p Pressure, Pa w weight coefficient
P Proportion of index value r correlation coefficient
Slt lifetime, year CF conversion factor

Re Reynolds number Pr Prandtl number

Y Fin pitch, m Nu Nusselt number

H Fin height, m G mass velocity, kg·m-2·s-1


d Diameter, m λ Thermal conductivity, W·K-1·m-1
x Vapor quality r Latent heat, kJ·kg-1
M Mass, kg D Distance

R Grey relational degree cr relational coefficient


L Length, m I exergy destruction, W
Greek
η efficiency, % Ω Distinguish coefficient

𝛿 fin height, m λ Thermal conductivity, W·K-1·m-1


𝜌 Density, kg·m-3 μ Dynamic viscosity
Superscripts
g exhaust gas e evaporator
c condenser f working fluid
H high temperature loop L low temperature loop
s isentropic p pump
in inlet out outlet
t turbine cw cooling water
pw water pump pf pump of working fluid
gen generator ops operation
pri price om operation and management
bm buy money he heat exchanger

max maximum min minimum


tot total o outside

w wall i inside
Superscripts
r reduction em emission
' saturation liquid " saturation vapor
Declaration of interests

√ The authors declare that they have no known competing financial interests or personal
relationships that could have appeared to influence the work reported in this paper.

☐The authors declare the following financial interests/personal relationships which may be
considered as potential competing interests:
Highlights

 Multi-objective optimization and improved GRA are used to working fluid selection.
 System performance with different fluids are compared under the optimized conditions.
 Boiling temperature is a working fluid selection criteria for DORC system.
 Cyclohexane/butane is the optimal working fluid when flue gas temperature is 573.15K.

You might also like