Energy Conversion and Management 177 (2018) 718–736
Contents lists available at ScienceDirect
Energy Conversion and Management journal homepage: www.elsevier.com/locate/enconman
Thermodynamic and thermo-economic analysis of dual-pressure and single pressure evaporation organic Rankine cycles
T
Mingtao Wanga, , Yiguang Chenb, Qiyi Liua, Zhao Yuanyuana ⁎
a b
School of Energy and Power Engineering, Ludong University, 186 Middle Honqi Road, Zhifu District, Yantai 264025, China College of Civil Engineering & Architecture, Jiaxing University, Jiaxing 314001, China
ARTICLE INFO
ABSTRACT
Keywords: Dual pressure Organic Rankine cycle Performance evaluation Electricity production cost
The goal of this paper is to study and evaluate the thermodynamic and the thermo-economic performance of dual-pressure and single pressure evaporation organic Rankine cycle (DORC and SORC) using isobutane as the working fluid. The effects of the two-stage pressures and the heat source inlet temperature on the system performances have been investigated. Then the performance comparisons between SORC and DORC have been conducted at the optimized condition. Results show that the DORC yields more net power than the SORC when the heat source temperature is between 100 and 177.2 °C, and the net power output gains of the DORC are higher at lower heat source temperature. The DORC no longer has the performance advantage as the heat source temperature is above 177.2 °C. The thermo-economic evaluations show that the optimized electricity production cost (EPC) decreases as the heat source temperature rises. Furthermore, optimized EPC of SORC and DORC are nearly equal at the same heat source temperature. For the heat source temperature of 100–177.2 °C, the DORC can significantly increase the net power output, but the thermo-economic performance of DORC is not reduced by system complexity compared to the SORC.
1. Introduction The main energy consumption fields of the world are industry [1] and building [2]. Due to the emission regulations and the increasing of energy prices, waste heat recovery [3] and application of renewable energy [4] have received more and more attention. Organic Rankine cycle (ORC) has been considered as an effective and promising method for translating low-media temperature heat to power [5]. Therefore, it has been widely used in the waste heat recovery [6,7]. However, the thermodynamic performance of the ORC is limited by the cycle structure. Lecompte et al. [8] pointed out that there is a mismatch between the working fluid isothermal phase change process and the heat source temperature linear drop process in the subcritical ORC, which means a large heat transfer difference and increases the exergy loss in the evaporator. In order to improve the temperature match in the evaporator, some scholars introduced the modified ORC, such as the recuperated cycles [9]. Maraver et al. [10] investigated the ORC performance with and without a recuperator. The result showed that the recuperated ORC could improve the thermal efficiency, but reduce the utilization rate of heat source. Therefore it could not increase the net power output. Furthermore, using zeotropic mixtures as working fluids is also an ⁎
approach to improve the temperature match between heat source and working fluid, because the zeotropic mixtures have temperature glides during the phase change process [11,12]. Nevertheless, the temperature match between the zeotropic mixtures and heat source is still not satisfactory, especially when the temperature drop of heat source is much larger than the temperature glide of the zerotropic mixtures in the evaporation process, as shown in Ref. [13]. Liu et al. [14] also found that ORC system using zeotropic mixtures showed better cycle performance when the temperature increase of cooling water was nearly equal to condensation temperature glide; however, there is still large exergy loss in the evaporator. In the transcritical ORC (TORC), the working fluid temperature continues to rise by absorbing heat at the pressure above the critical pressure in the evaporator [15,16]. That will improve the temperature match in the evaporator. Therefore, the average heat transfer temperature difference will be decreased, and the cycle performance would be improved, compared with the subcritical ORC [17,18]. Besides the performance advantages, TORC also have some drawbacks. Li et al. [19] investigated the performances of the subcritical and transcritical ORCs using R1234ze (E) as working fluid. They found that the TORC had better thermal efficiency; however its heat absorption capacity decreased compared to the subcritical ORC. Maraver et al. [10]
Corresponding author. E-mail address:
[email protected] (M. Wang).
https://doi.org/10.1016/j.enconman.2018.10.017 Received 15 August 2018; Received in revised form 21 September 2018; Accepted 7 October 2018 0196-8904/ © 2018 Published by Elsevier Ltd.
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
Nomenclature A Ash Cai Cp Cbm Ctot cps Dsh do di F Fb Fcr Fp Fbm Fk Gs h he2 ho hi hd hnb hnc hl hload ht hs Jl Jb Js ji m n Nu Pr Pt q Q Re T U v W y
Xtt x ΔT
heat exchanger surface area, m2 bundle cross flow area, m2 annuity of the investment cost, k$ purchased cost, k$ bare module cost, k$ total investment cost, k$ specific heat, kJ/(kg·K) the shell diameter, m diameters of outside tube, m diameters of inside tube, m cost factor correction factor for the effect of convection capital recovery factor pressure factor bare module cost factor maintenance cost factor shell side mass velocity, kg/(m2·s) enthalpy, kJ/(kg·K) boiling heat transfer coefficient, W/(m2·K) heat transfer coefficients, W/(m2·K) heat transfer coefficients for inside flow, W/(m2·K) ideal transfer coefficient for pure cross flow, W/(m2·K) nucleate boiling coefficient, W/(m2·K) natural convection heat transfer coefficient, W/(m2·K) heat-transfer coefficient of the liquid phase, W/(m2·K) annual operation hours, h heat transfer coefficient of tube-side, W/(m2·K) condensation heat transfer coefficient, W/(m2·K) correlation factor for baffle leakage correction factor for bundle bypassing effects correlation factor for variable baffle spacing Colburn j-factor mass flow rate, kg/s economic life time, year Nusselt number Prandtl number tube pitch heat flux, kW/m2 heat capacity, kW Reynolds number temperature, °C overall heat transfer coefficient, W/(m2·K) velocity, m/s power, kW interest rate
Lockhart–Martinelli parameter vapor mass fraction temperature difference, °C
Greek symbols δbb
k µ
bypass clearance, m thermal efficiency heat recovery effectiveness thermal conductivity, W/(m·K) viscosity, kg/(m·s) density, kg/m3
Abbreviations DORC EPC SORC ORC
dual-pressure ORC electricity production cost single-pressure ORC organic Rankine cycle
Subscript a amb bm c cool con d eva h hs i in LMTD l o pp out pum s sub sh tur tot t w
absorption ambient bare module critical cooling water condensation dual-pressure evaporator High-state pressure Heat source inside inlet log-mean temperature difference low-state pressure outside pinch point outlet pump single pressure sub-cooling shell turbine total tube tube wall
recovering the engine waste heat. They found that the dual-loop ORC system could provide a maximum net power output (23.62 kW) and a minimum EPC (0.41 $/kW h) at the rated condition. The thermal efficiency of the system is between 8.97% and 10.19% over the whole operating range. Indeed, dual-loop ORC was more suitable for two or more heat sources heat recovery, as also shown in Ref. [22]. However, this kind of cycle still had high exergy destruction in the evaporator, especially in the high-stage pressure, on account of the isothermal phase change of the working fluid. Different from the dual-loop ORC, the DORC (dual-pressure ORC) system splits the pure working fluid into two evaporation process with different pressure values, which could reduce the heat transfer
researched performances of subcritical and transcritical ORC systems. It was found that the highest second law efficiency was achieved by transcritical ORC using R134a when the geothermal heat source temperature was 170 °C. However, Maraver et al. [10] also emphasized that there were also some drawbacks for the TORC, such as high pressure and high expansion ratios. Besides, dual-loop ORC, also called cascade ORC, was also proposed for heat recovery. Shu et al. [20] investigated the dual-loop ORC to recovery both the flue gas waste heat and cylinder cooling water waste heat. The proposed cycle achieved a power output higher than that of the Steam Rankin cycle. Yang et al. [21] recently analyzed the thermodynamic and economic performance of dual-loop ORC system for
719
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
temperature difference in the evaporator. The advantages of the DORC have been presented by some researches. Franco et al. [23] evaluated the system performance improvement potential of the DORC system. They found that the DORC using isobutene or n-pentane as working fluid could increase the exergy efficiency by 15–21% compared to the SORC in the recovery of geothermal water at 160 °C. Peris et al. [24] investigated system performances of the 6 ORC layouts using 10 working fluids for recovering the jacket water waste heat from the engines. The results showed that the dual-pressure layout using SES36 as working fluids achieved the maximum efficiency of 7.15%. Stijepovic et al. [25] investigated the potential of multi-pressure ORC for recovering the waste heat of a 320 °C heat source using the exergy composite curves approach. The results indicated that, because of the rise of the heat capacity absorbed from the heat source, the power output of the multi-pressure ORC was higher than that of the SORC. It was also found that the third-pressure ORC improved a marginal system performance compared with the DORC system. Guzovic et al. [26] evaluated the feasibility of the DORC for recovering a 175 °C geofluid. The authors showed that the DORC system increased the exergy
efficiency from 52% to 65%, and brought an obvious increase in the net power output from 5270 to 6371 kW. Li et al. [27] investigated and compared the system performances of SORC and DORC (two-stage ORC) for recovering the geothermal heat source of 100 °C. The results showed that the DORC improved the net power output from 622.7 to 669.7 kW compared to the SORC. And the irreversible loss in the evaporator decreased from 360.8 to 295.2 kW. Thierry et al. [28] analyzed the performance of the DORC using mixture as working fluid for recovering waste heat of 90–110 °C. The results indicated that the DORC improved efficiency from 6.85% to 7.70% compared to the SORC when heat source temperature was 90 °C. Li et al. [29] investigated and evaluated system performances of series two-stage ORC (STORC), parallel two-stage ORC (PTORC) and single pressure ORC (SORC) using 90–120 °C geothermal water as heat source. The results showed that PTORC and STORC could reduce exergy loss in the evaporator and improve the net power output. Moreover, it also found that STORC showed best system performance, and could be widely used in engineering. Shokati et al. [30] compared the SORC, DORC, dual-loop ORC and Kalina cycle from the viewpoints of energy, exergy and
90 80
Isobutane
Temperature/K
70
5
3
4
60 50 40
6
2
7
1 8
30 20 10 1.0
1.2
1.4
a) SORC
1.6 1.8 2.0 Entropy/kJ(kg K)-1
2.2
2.4
c) T-s diagram of SORC 90 80
Isobutane
8
Temperatrue/K
70 60
5
50
3
4
2 1
30 1.2
13
12
1.4
1.6
1.8
2.0
Entropy/kJ kg K)-1
b) DORC
d) T-s diagram of DORC Fig. 1. Schematic and T-s diagrams of SORC and DORC.
720
9 10 11
40
20
7
6
2.2
2.4
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
performances. In addition, the reasons behind the advantages of DORC system are also investigated.
Table 1 System parameters of the SORC and DORC system models. Parameters
Value
Heat source temperature, Ths Heat source pressure, Phs Heat source mass flow rate, mhs Pinch point temperature difference in the evaporator, ΔTeva Condenser sub-cooling, ΔTsub Cooling water inlet temperate, Tcool Cooling water pressure, Pcool Condensation temperature, Tcon Pinch point temperature difference in the condenser, ΔTcon Turbine efficiency, ηt Pump efficiency, ηp Ambient temperature, Tamb
100–200 °C 0.5–1.6 MPa 5 kg/s 10 °C 2 °C 20 °C 0.101 MPa 35 °C 5 0.85 0.7 25 °C
2. Methodology This section introduces the SORC and DORC systems using isobutane as the working fluid for the following thermodynamic and thermo-economic comparisons. Moreover, boundary conditions, assumptions and optimization process are also presented for the analysis of the cycle performances. 2.1. SORC and DORC Fig. 1 shows the schematic diagrams of SORC and DORC using pure working fluids, and the corresponding T-s diagrams respectively. For the SORC system (Fig. 1a and c), the high pressure working fluid from the pump flows into the evaporator, where it absorbs heat and converts to the superheated vapor (2–5). Then, the superheated vapor enters the turbine to generate power (5–6). The exhausted vapor from turbine goes through the condenser and condenses to the liquid. At last, the liquid working fluid is compressed and enters the evaporator to complete a cycle. For the DORC system (Fig. 1b and d), the working fluid from the preheater 1 is split into two sections. One part enters the vaporizer and absorbs heat (3–4). The other part flows into the pump 2 and is further compressed to the high-stage pressure (3–5). Then it flows into the high-stage evaporator (preheater1, vaporizer 2 and superheater 2), absorbs heat and then becomes superheated (5–8). The superheated vapor expands in the high-pressure turbine (8–9) and then mixes with the working fluid from the vaporizer 1 in the inlet of the turbine 1. The mixed working fluid flows into the turbine 1 and undergoes the second expansion (10–11). The exhaust vapor from the turbine 1 enters into the condenser and condenses to the liquid (12–1). At last, the liquid working fluid is compressed and enters the preheater1, where the whole cycle is completed.
exergoeconomic. They found that the net power output of the DORC was maximum, 15.22% higher than the SORC. Manente et al. [31] proposed and compared the SORC and DORC using geothermal heat source of 100–200 °C. The results indicated whether the thermodynamic performance of the DORC was better than that of SORC depended on the working fluid and heat source temperature. The increment of the power output of the DORC decreased, and at last disappeared as the temperature of heat source increased. Sadeghi et al. [32] analyzed and compared the performances of the simple ORC, PTPRC and STORC using zeotropic working fluids for heat recovery from geothermal water. The minimum turbine size parameter (TSP) and maximum net power output were chosen as the optimized goals. It was observed that STORC yielded the maximum net power and R407A was the most suitable working fluid candidate. Li et al. [33] studied the applicable range of the DORC and investigated the influences of the working fluid critical temperature on the suitable heat source temperature range. Results showed that the SORC and DORC systems had different heat source temperature applicable range, which was closely relevant to the working fluid critical properties. In addition, they found that there was a linear relationship between applicable range of heat source temperature and the critical temperature of the working fluid. Li et al. [34] investigated and evaluated the thermal economic performance of the separate and induction turbine layouts, and analyzed the effects of high-stage and low-stage pressure on the thermal economic performance. They found that induction layout produced more net power, and the maximum decrement of the specific investment cost was 34.2%. However, the existing researches on DORC mainly concentrated on the system parameters optimization or the performance comparisons with other ORC layouts, and few published researches focus on the DORC thermo-economic performance. In fact, the DORC system is more complex because of the two evaporation process, two turbines and pumps. Although the thermodynamic performance of the DORC is better than that of the SORC, components investment of the DORC also increases simultaneously. Therefore, it is essential to investigate the applicability of the DORC from the aspect of the thermo-economic. In this paper, the thermodynamic and the economic performances of DORC and SORC are investigated and compared. The thermodynamic and economic models of the SORC and DORC systems are built by the Matlab software. The influences of the high-stage and low-stage pressures on SORC and DORC systems parameters (Outlet temperature of heat source, working fluid mass flow rate, thermal efficiency, net power output, investment cost and electricity production cost) are investigated. On the basic of this, the optimization processes are carried out to compare the thermodynamic and economic performances of DORC and SORC systems. Then the feasibility and advantage of the DORC system is discussed in view of thermodynamic and economic
2.2. System parameters and assumptions The models of the SORC and DORC are established using Matlab software and the REFPROP 9.0. System parameters of SORC and DORC are listed in the Table 1. For simplifying the analysis, the assumptions of the models are presented as follows. (1) The SORC and DORC systems operate at a steady state; (2) The pressure drop of exchangers and the pipe heat dissipating capacity could be neglected; (3) The input power of the cooling pump is calculated as 1% of the total capacity of the heat release in the condenser [35]; (4) The counter-current flow exchangers are adopted; (5) The maximum pressure (Pe,max) is set as 0.9Pc for the SORC and Table 2 The ranges of the optimized parameters of the SORC and DORC systems [31,33].
721
Cycle layout
Parameters
Maximum value
Minimum value
SORC
Evaporation pressure, Peva Superheat degree, ΔTeva
0.9 Pc /
Pcon + 100 kPa 2K
DORC
High-stage pressure of, P,h Low-stage pressure of, P,l Superheat degree of the highstage, ΔTh Superheat degree of the low stage, ΔTl
0.9 Pc 0.9Pc-100 kPa /
Pe,l + 100 kPa Pcon + 100 kPa 2K
/
0.01 K
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
Heat
Ts1 Temperature
are the enthalpies of the heat source water in the evaporator inlet and outlet respectively, kJ/kg. Thus, the mass flow rate of the working fluid of SORC is
Tin
SORC ce sour
msw =
Tsout Working fluid
3
(2)
The equations used to analyze the performance of SORC are shown in Table 3. The net power of the SORC system can be calculated as follows:
5 4
(3)
Wsnet = Wstur Wspum Wscool
where Wscool is the pump power, assumed as 1% of the heat rejection in the condenser, kW. According to Fig. 2b, the mass flow rate of the working fluid in the vaporizer 2 and vaporizer 1 are
2 Heating capacity
mdh =
mhs (hdin hd2 ) h7 h5
(4)
mdl =
mhs ·(h d2 h dout ) h 4 h2
(5)
a) SORC Tdin
DORC
Temperature
Qa (h5 h2 )
Heat
Td3
s
e ourc
Td2
Td1 7 6
Tdout
where hd2 is the enthalpy of heat source water at the temperature of Td2 in Fig. 2b, kJ/kg. The equations used to analyze the performance of DORC are shown in Table 4. The net power of the DORC system is calculated as:
Working fluid
Therefore, the thermal efficiency of SORC and DORC can be determined by:
4
3
(6)
Wdnet = Wdtur1 + Wdtur2 Wdpum1 Wdpum1 Wdcool
5
(7)
= Wnet / Qa
2
by
Heating capacity
=
b) DORC
Table 3 Equations used in the energy analysis of SORC. Equation
Turbine
Wstur = msw ·(h5 h6) = m sw ·(h5 h6s)·
Pump
Evaporator Condenser
Wspum = msw ·(h2 h1) = msw ·(h2s h1)/
(8)
2.4. Economic models 2.4.1. Heat transfer calculation The shell and tube heat exchanger is used in this paper. The heat source water and cooling water flow in the tube side, and the working fluid flows in the shell side. The heat-transfer surface of the exchangers can be obtained by:
tur pum
Qeva = mhs·(h sin h sout ) Qcon = m sw ·(h7 h1)
A= DORC [33]; (6) The working fluid from turbine is superheated; (7) The ranges of the optimized parameters are shown in Table 2.
Q U · TLMTD
(9)
Table 4 Equations used in the energy analysis of DORC.
2.3. Thermodynamic models
Component
Equation
Turbine1
Wdtur1 = (mdh + mdl )·(h10 h11) = (mdh + mdl )·(h10 h11s)·
Turbine2 Pump 1
Fig. 2 shows the temperature profiles in the evaporators of the SORC and DORC. According to Fig. 2a, the heat absorption of the SORC from the heat source is
Qa = mhs ·(hsin hsout )
Qa mhs ·(hsin hamb)
where hamb is the enthalpy of heat source water at the ambient temperature, kJ/kg.
Fig. 2. Temperature profiles in the evaporator of SORC and DORC.
Component
And the heat recovery effectiveness of SORC and DORC is calculated
(1)
where mhs is the mass flow rate of the heat source, kg/s; hsin and hsout
722
Wdtur2 = m dh ·(h8 h 9) = mdh (h8 h 9s)·
Wdpum1 = (mdh + mdl)·(h2 h1) = (mdh + mdl )·(h2s h1)/
Pump 2
Wdpum2 = mdh ·(h5 h3) = m dh (h5s h3)/
Evaporator Condenser
Qeva = mhs·(h din h dout ) Qcon = (mdh + m dl )·(h12 h1)
tur
tur
pum
pum
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
where A is the heat transfer area, m2; Q is the heat transfer capacity of the heat exchanger, kW; ΔTLMTD is the log-mean temperature difference of the heat exchanger, given by
Tmax
TLMTD =
ln(
Tmin Tmax ) Tmin
where y is the interest rate, and assumed as 5%; n is economic life time, assumed as 15 years [40]. 2.5. Optimization process
(10)
The net power output and EPC are used to evaluate the system performance. For SORC system, the decision variables of the optimization process are the evaporation pressure and the superheating degree. For the DORC system, the high and low stage evaporation pressures along with the superheat degrees are optimized. The following procedure is the optimization process for the DORC system, which is shown in Fig. 3. The optimization process of the DORC system is similar.
The overall heat transfer coefficient of the heat exchanger is
1 1 1 d d d = + · o + o ·lg o U ho hi di 2µ µ
(11)
where μ is the tube thermal conductivity, W/(m·K); do and di are the outside and inside diameters of the tube, m; ho and hi are the heat transfer coefficients for outside and inside flows, respectively, W/ (m2·K).
(1) The high-stage pressure and low-stage pressures are divided into 50 equal parts between the maximum and minimum values respectively; (2) The initial superheat degree of the high stage evaporator is set as 2 K. The system performances are calculated according to the above conditions; (3) With the calculation results to determine whether the working fluid from turbine 2 is superheated. If the result is satisfactory, go to step 4; otherwise, increasing superheat degree of the evaporator 2 and repeating the calculation until the working fluids is superheated. (4) According to the superheat degree obtained from step 3, the system performances at different pressures are calculated. The operation pressures which could not meet the demand of pinch point temperature difference (PPTD) in the evaporator are ignored.
2.4.1.1. Heat transfer calculation of shell side. The single phase heat transfer coefficient of the shell side is calculated by the correlation of Bell-Dlaware [36]:
ho1 = h id Jc Jl Jb Js
(12)
The boiling heat transfer coefficient of the evaporator is calculated as [37]:
he2 = h nb Fb + h nc
(13)
The condensation heat transfer coefficient for the two-phase fluid in the condenser is calculated as the following [38]:
hs / hl = 1.26Xtt 0.78
(14)
The parameters of Eqs. (12)–(14) are given in Appendix A.1–3. 2.4.1.2. Heat transfer of the tube side. The correlation of Petukhow and Popov [39] required for the calculation of the single phase heat transfer coefficient of the tube side can be shown in the following equation.
ht =
Nu·k di
(15)
The parameters of Eq. (14) are given in Appendix A.4 2.4.2. Investment cost calculation The total investment cost of the systems in 2014 is calculated by [40]:
Ctot2014 =
Ctot2001 CEPCI2014 CEPCI2014
Ctot2001 = Cbm,eva + Cbm,con + Cbm,tur + Cbm,pum
(16) (17)
where CEPCI stands for Chemical Engineering Plant Cost Index; CEPCI2001 = 382, CEPCI2014 = 586.77 [41]. The bare module cost for evaporator, condenser, turbine, and pump are given in Appendix B. The electricity production cost (EPC) can be calculated by:
EPC =
Cai + Fk Ctot2014 Wnet ·hload
(18)
where Fk is the maintenance cost factor, assumed as 1.65% [41]; hload is the annual operation hours, and its value is 7446 h/year [42]; Cai is the annuity of the investment cost, estimated as:
Cai = Ctot2014· Fcr
(19)
where Fcr is the capital recovery factor, and is calculated as:
Fcr =
y (1 + y )n (1 + y )n 1
Fig. 3. Diagram of the optimization process of DORC.
(20)
723
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
position of the pinch point temperature difference (ΔTpp) in the evaporator occurs at ΔTpp1, as shown in Fig. 5. In case of satisfying the ΔTpp, rising evaporation pressure could lead to increase of the heat source outlet temperature, as shown in Fig. 6, which means the utilization rate of the heat source decreases. On the other hand, increasing Pe means the increase of the thermal efficiency. Therefore, there exists an optimal Pe for the SORC system to obtain the maximum Wnet (Wmax). When Ths increases to 180–200 °C, the position of the ΔTppwill move from ΔTpp1 to ΔTpp2, as shown in Fig. 5. Tout depends on the working fluid temperature from the pump1, and it keeps basically constant as shown in Fig. 6. In other words, the utilization of the heat source almost remains unchanged. Therefore, Wnet is basically proportional to the thermal efficiency of SORC, and increases with the rise of Pe. Moreover, it is apparent that Wnet increase with the rise of Ths, due to the increase of Ths and the decrease of Tout. The green five stars points in the Figs. 4 and 6 are Wmax and the corresponding Pe and Tout at different Ths. In this paper, parameters of the system corresponding to the maximum Wnet are called optimized parameters.
Table 5 System input parameters and boundary conditions of the system models [31]. Parameters
Value
Working fluid Heat source temperature, Ths Heat source mass flow rate, mhs Pinch point temperature difference, ΔTeva Condenser sub-cooling, ΔTsub Condensation temperature, Tcon Turbine efficiency, ηt Pump efficiency, ηp Ambient temperature, Tamb
Isobutane 100 or 150 °C 100 kg/s 10 °C 2 °C 35 °C 0.85 0.7 20 °C
3. Model verification The calculation results of this paper were compared with the data presented by Manente et al. [31] for SORC and DORC. The same input parameters and boundary conditions are shown in Table 5. The output parameters and comparison results of the SORC and DORC at heat source temperature of 100 °C and 150 °C are presented in Tables 6 and 7 respectively. It can be seen that the maximum relative error is 3.56%, which is the mass flow rate of the working fluid in the SORC, followed by 3.42%, which is the heat recovery effectiveness. Other relative errors are usually less than 3%. Moreover, it should be noted that the relative errors of the net power output are less than 0.2%. The error reasons are as follows. In this paper, the models of the SORC and DORC are built using Matlab software and the REFPROP 9.0, whereas in Ref. [33], the models are established using the EES (Engineering Equation Solver) software and its internal fluids properties library. Therefore, the relative errors are mainly caused by the different fluids properties library (EES 10.101 vs. REFPROP 9.0) and different calculation step lengths. From the comparison results, it can be seen that the calculation models of this paper are reliable.
4.1.2. DORC Compared with the SORC system, the DORC system is more complex which is affected by the high stage pressure (Ph) and the low stage pressure (Pl) simultaneously. As Ths increases, the variations of Tout, thermal efficiency (ηth), working fluid mass flow rate (M) and Wnet are similar. Therefore, take Ths = 100 °C as an example to analyze the parameters variations of the DORC. 4.1.2.1. Outlet temperature of the heat source. Fig. 7 shows the variations of Tout with Ph and Pl of the DORC system when Ths is 100 °C. It is apparent that Tout increases with the increase of Pl, while basically remains unchanged with the increase of Ph. The reason is as follows: for the DORC system, when Ths is 100 °C, the optimized Ph and Pl are 572.9 kPa and 1160.1 kPa respectively, and the position of ΔTpp in low stage evaporator occurs between Point 1 and Point 2 in Fig. 8a. If Pl decreases from 772.9 kPa to 572.9 kPa, the position of ΔTpp will move from 1–2 to 1′–2′. With the limitation of ΔTpp, Tout decreases from 58.6 °C to 49.3 °C. If Ph increases from 1160.1 kPa to 1360.1 kPa, the position of ΔTpp in the higher evaporator will move from 3–4 to 3′–4′. While Tout basically remains unchanged (58.5 °C −58.6 °C), because of the fact that the position of ΔTpp in the high stage evaporator is more far away from the heat source outlet, as shown in Fig. 8b. Therefore, the variations of Ph have little effect on Tout.
4. Results and discussions 4.1. Thermodynamic performance analysis 4.1.1. SORC Fig. 4 presents the influences of the heat source inlet temperature (Ths) and evaporation pressure (Pe) on the variation of the net output power (Wnet) of the SORC. It is obvious that, when Ths is 100–160 °C, Wnet firstly increases and then decreases as Pe increases. When Ths is higher than 160 °C (180 and 200 °C), Wnet increases with the rise of Ths. The reason is as follows: When Ths is between 100 and 160 °C, the
4.1.2.2. Net power output of turbines. Fig. 9 shows net power output of the two turbines over the Ph and Pl when Ths is 100 °C. The variations of Mh, Ml and Mtot are presented in Fig. 10. Increasing Pl has two conflicting influences on Wtur1, one is increasing the pressure
Table 6 Comparison of the results in this paper and the data presented by Manente et al. [31] for SORC. Isobutane
Pe (kPa) m (kg/s) T3 (°C) Tout (°C) Wnet (kW) Φ (%) ηth (%) ηsys (%)
Ths = 100 °C
Ths = 150 °C
This paper
Manente et al.
Relative error
This paper
Manente et al.
Relative error
996.98 35.92 66.05 68.75 755.1 39.16 5.75 2.253
975 37.2 65.1 67.7 756.0 40.5 5.58 2.257
2.2% −3.56% 1.44% 1.53% 0.12% 3.42% 3.0% 0.18%
1861.4 82.35 96.53 72.75 3269.9 59.74 10 5.97
1860 82.3 96.6 72.7 3270 59.7 9.99 5.969
0.07% 0.06% −0.07% 0.07% 0 0.07% 0.1% 0.017%
724
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
Table 7 Comparison of the results in this paper and the data presented by Manente et al. [31] for DORC. Isobutane
Ths = 100 °C
Php (kPa) Php (kPa) mhp (kg/s) mlp (kg/s) Tsat_hp (°C) Tsat_lp (°C) Tout (°C) Wnet (kW) Φ (%) ηth (%) ηsys (%)
450 400 350
This paper
Manente et al.
Relative error
This paper
Manente et al.
Relative error
1160.1 772.89 26.56 21.44 72.99 55 58.6 960.16 51.85 5.53 2.87
1170 795 26.0 20.8 73.5 56.2 59.6 961.5 50.6 5.67 2.87
−0.08% −2.86% 2.1% 2.98% −0.7% −2.18% −1.7% −0.14% 2.4% −2.5% 0
2534.6 1241.1 62.61 33.16 113.5 76.18 60.37 3868.1 69.2 10.21 7.06
2530 1250 62.9 32.8 113.3 76.6 60.5 3871 69.2 10.22 7.066
0.18% −0.7% −0.463% 1.09% 0.18% −0.55% 0.21% 0.075% 0 0.1% 0.09%
Ths=100 oC
Ths=180 oC
Ths=120 oC
Ths=200 oC
420 410
413.9
Ths=140 oC
400
358.3
Ths=160 oC
300
390 380
250 200
206.3
150
1000
Ths=180 oC Ths=200 oC
360
1500
2000
2500
3000
310 300 600 900 1200 1500 1800 2100 2400 2700 3000 3300 Pe/kPa
3500
Fig. 4. Wnet of the SORC system.
Fig. 6. Tout of the SORC system.
180
Tout/oC
1500
160
Isobutane Heat source
140 120
1400 1300 1200
pp1
Ph/kPa
Temperature/oC
Ths=160 oC
320
Pe/kPa
100 80
40
0
400
800
1200
1600
2000
2400
2800
1000 Ths=100 oC
800
pp2
-400
1100
900
Isobutane Heat source
60
20
Ths=140 oC
330
37.8 500
Ths=120 oC
340
75.2
50
370
Ths=100 oC
350
128.8
100
0
Tout/oC
Wnet/kW
Ths = 150 °C
700
3200
48.0 50.6 53.3 55.9 58.5 61.1 63.8 66.4 69.0 71.6 74.3 76.9 79.5 82.1 84.8 87.4 90.0
600 700 800 900 1000 1100 1200 1300 1400
Heat absorption capacity/kW
Pl/kPa
Fig. 5. Temperature profiles in the evaporator of SORC.
Fig. 7. Outlet temperature of heat source of DORC.
difference of turbine1 which has a positive influence on Wtur1, and the other is the reduction of total mass flow rate (Mtot) that has a negative effect on Wtur1. Therefore, Wtur1 increases firstly and then decreases as Pl increases, as shown in Fig. 9a. Moreover, it can be seen that Wtur1 most remain unchanged as Ph increases. This is because the Mtot and pressure difference of pump1
remain basically unchanged, which leads to no considerable variation in Wtur1, as Ph increases. As shown in Fig. 9b, the power output of turbine 2 (Wtur2) increases firstly and then decreases as Ph increases. Same as Pl in Fig. 9a, increasing Ph would lead to the increase of the pressure difference of
725
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
Isobutane Heat source Isobutane Heat source
90
Temperature/oC
80 70
o
Tout=58.6 C
60 50 40
2' Tout=49.3 oC
3' Ph=1160.1 kPa
Pl=772.9 kPa
33.40 36.06 38.72
900
Pl=572.9 kPa
41.38
800
o
Ths=100 C
700
Isobutane Heat source Isobutane Heat source
90 80
4'
Tout=58.6 oC
3 2'
2
o
Tout=58.5 C
1
1'
Wt2/kW
1500
Ph=
1400 1300
3'
1200
Ph=1160.1 kPa
Pe=772.9 kPa
1100 1000 Ths=100 oC
900
40
800
30 -200 -100
0
46.70
a ) Turbine 1
1360.1 kPa
4
44.04
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
Ph/kPa
100
Temperature/oC
30.74
1000
a) Decreasing Pl
50
28.08
1100
-200 -100 0 100 200 300 400 500 600 700 800 900 10001100 Qe/kW
60
25.42
1200
30
70
22.76
1300
3
1'
20.10
1400
4'
4
2
1
Wt1/kW
1500
Ph/kPa
100
700
100 200 300 400 500 600 700 800 900
0.50 2.63 4.75 6.88 9.00 11.13 13.25 15.38 17.50 19.63 21.75 23.88 26.00 28.13 30.25 32.38 34.50
600 700 800 900 1000 1100 1200 1300 1400
Qe/kW
Pl/kPa
b) Increasing Ph
b) Turbine2
Fig. 8. Temperature profiles in the evaporator of DORC.
Fig. 9. The power output of two turbines of DORC over Ph and Pl.
turbine 2 and the decrease of the Mh. It also can be seen that Wt1 decreases with the increase of Pl. This is because of the turbine1 pressure difference and Mtot decrease as the Pl increases, which leads to reduction of Wtur1. 4.1.2.3. Thermal efficiency. Fig. 11 illustrates the effects of Ph and Pl on the thermal efficiency (ηth) of the DORC. As above analysis, Wt2 increases firstly and then decreases, whereas Wt1 (Fig. 9a) and Tout (Fig. 7) most remain unchanged with the increase of Ph. Therefore, ηth increases firstly and then decreases with the rise of Ph. Moreover, it is apparent that ηth increases as Pl increases. The main reason is that the increase of Pl leads to the rise of Tout, which decreases the utilization rate of the heat source.
and Wt2: Wt1 increases firstly and then decreases, whereas Wt2 remains basically unchanged with the increase of the Ph. Increasing of Pl also has the similar effects on the Wt1 and Wt2. Consequence of above mentioned effects leads to change trend in Wnet. The higher Wnet can be seen in the range of 1000 kPa < Ph < 1400 kPa and 650 kPa < Pl < 1000 kPa from Fig. 12a. When Ths is 100 °C, the maximum Wnet (Wmax) is 48.0 kW, and the optimized pressures (Ph and Pl) are 1160 kPa and 773 kPa respectively. Fig. 12b–d are the Wnet contours when Ths are 120 °C, 140 °C and 160 °C respectively. The contours show that it has the same tendency of changes, and the reason is also similar. Moreover, the maximum Wnet markedly increases from 48.0 kW to 241.7 kW as Ths rises from 100 °C to 160 °C due to the rise of the optimized Ph, Pl and Mtot.
4.1.2.4. Net power output. Fig. 12a shows the variations of Wnet of the DORC system over Ph and Pl, when Ths is 100 °C. It is apparent that the Wnet increases firstly and then decreases as Ph and Pl increases. The reason is explained as follows. Increasing Ph has different effects on Wt1
4.1.3. Thermodynamic performance comparison Fig. 13a presents maximized net power outputs (Wmax) of SORC and DORC under the optimized operation conditions. For the Ths of 100–160 °C, Wmax of the SORC increases from 37.8 to 206.3 kW,
726
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
Mh/kg s-1
1400 1300
Ph/kPa
1200 1100 1000 Ths=100 oC
900 800 700
1400 1300 1200
700
Ph/kPa
Fig. 11. Thermal efficiency of DORC.
Ml/kg s-1
1200 1100 1000
Ths=100 oC
900 800 700
0.26 0.43 0.59 0.76 0.93 1.10 1.26 1.43 1.60 1.77 1.93 2.10 2.27 2.44 2.60 2.77 2.94
whereas that of the DORC rises from 48 to 241.7 KW. It is apparent that Wmax of DORC is larger than that of SORC at the same heat source inlet temperature. The reason for the gains of Wnet in DORC is as follows. The Wmax depends on the heat absorption capacity and the thermal efficiency. Fig. 13b shows the thermal efficiency of the SORC and DORC. It is found that the thermal efficiency of DORC show no obvious advantages, even slightly lower than that of SORC at heat source temperature of 100 °C (5.53% vs. 5.75%). However, the heat absorption capacity of the DORC increases significantly compared to the SORC at the same Ths, as shown in Fig. 13c. Therefore, the DORC system yields more net power than SORC system under the same heat source inlet temperature. Moreover, it can be found from Fig. 13a that the increment of Wmax decreases from 21.3% to 14.6% when Ths increases from 100 °C to 160 °C. That is to say that the Wmax increment of DORC decrease with the increase of the Ths, and DORC system has more advantage at lower Ths. From the exergy point of view, the average heat transfer temperature difference in the evaporator of DORC is smaller than that of SORC, which leads to the lower exergy loss in the evaporator. Accordingly, DORC shows better thermodynamic performance than SORC at the same heat source inlet temperature of 100 to 160 °C. It should be noted that when Ths increases to 177.2 °C, the thermal match in the evaporator of DORC and SORC are almost equal, as shown in Fig. 14. When Ths is higher than 177.2 °C, DORC system no longer has performance advantage, while its layout will be more complex. In a word, for the working fluids of isobutene, the SORC system is beneficial for Ths of 177.2 °C to 200 °C, while DORC system would show better performance for Ths of 100 °C to 177.2 °C under the boundary conditions of this paper. Therefore, in the next part, the thermo-economic performance comparisons of SORC and DORC are conducted at Ths of 100–160 °C.
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
b) Low-stage pressure Mtot(kg/s)
1500
0.53 0.70 0.86 1.03 1.20 1.36 1.53 1.70 1.87 2.03 2.20 2.37 2.53 2.70 2.87 3.03 3.20
1400 1300
Ph/kPa
1200 1100 1000 900 800 700
0.016 0.021 0.025 0.029 0.034 0.038 0.043 0.047 0.051 0.056 0.060 0.065 0.069 0.073 0.078 0.082 0.087
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
a) High-stage pressure
1300
Ths=100oC
1000
800
Pl/kPa
1400
1100
900
600 700 800 900 1000 1100 1200 1300 1400
1500
th
1500
0.19 0.36 0.52 0.69 0.85 1.02 1.19 1.35 1.52 1.69 1.85 2.02 2.19 2.35 2.52 2.68 2.85
Ph/kPa
1500
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
4.2. Thermo-economic performance analysis 4.2.1. SORC According to the above analysis, it could be found that Wmax of DORC is larger than that of SORC when Ths is 100–177.2 °C. However, the layout of DORC is more complex and expensive. Therefore, it is essential to make a thermo-economic comparison between DORC and SORC.
c) Total mass flow rate
Fig. 10. Mass flow rate of working fluid of DORC.
727
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
Wnet/kW
1400 1300 48.0 kW
Ph=1160kPa
Ph/kPa
1200 1100 1000 900 800
Ths=100 oC
700 600
Pl=773kPa
25 27 28 30 31 33 34 36 38 39 41 42 44 45 47 48 50
Wnet/kPa
3200 2800 2400
155.1kW
Ph/kPa
1500
2000
Ph=2130 kPa
1600 1200 800 800
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
1200
1600
1800 93.3 kW
Ph/kPa
1600
Ph=1566 kPa
1400 1200 1000
Ths=120 oC
800
Pl=929 kPa 600
2400
2800
c) Ths=140 oC Wnet/kW
3600
40 43 47 50 53 57 60 63 67 70 73 77 80 83 87 90 93
3200
Wnet/kW
Ph=3266.1 kPa 241.7 kW
2800 2400 Ph/kPa
2000
2000
Pl/kPa
a) Ths=100 oC 2200
Ths=140 oC
Pl=1137 kPa
65.0 70.7 76.3 82.0 87.6 93.3 98.9 104.6 110.3 115.9 121.6 127.2 132.9 138.5 144.2 149.8 155.5
2000 1600 1200
Pl=1397.2 kPa
Ths=160 oC
800
800 1000 1200 1400 1600 1800 2000 Pl/kPa
800
1200
b) Ths=120 oC
1600 2000 Pl/kPa
2400
150 156 162 167 173 179 185 190 196 202 208 213 219 225 231 236 242
2800
d) Ths=160 oC Fig. 12. The net power output of DORC over Ph and Pl.
Fig. 15 shows the total investment cost (Ctot) of SORC over Pe for different Ths. As can be seen, the changing trends for Ctot and Wnet are the same as the Pe increases. The reasons are as follows. The investment cost of the turbine accounts for more than 50% of Ctot. Therefore, Wnet is the most important factor affecting Ctot. Consequently, Ctot increases firstly and then decreases as Pe increases. Moreover, it can be seen that Ctot increases dramatically with the increase of Ths due to the increase of the system capacity. Fig. 16 presents the EPC of SORC system over Pe for different Ths. The change tendency of the EPC is opposite to the Ctot, and reasons are similar. It should be noted that EPC decreases with the increase of Ths, which means that the SORC shows better thermo-economic performance at higher Ths. The points of the green five stars in Figs. 15 and 16 are the system parameters corresponding to Wmax. It can be seen than the optimized EPC decreases from 0.24 $/kW h to 0.13 $/kW h, and the optimized Ctot increases from 607.5 k$ to 1785.6 k$ as Ths increases from 100 to 160 °C
4.2.2. DORC Fig. 17a shows the variation of the investment cost of pump1 (Cpum1) over Ph and Pl when Ths is 100 °C. Increasing Pl has two conflicting effects on the input power of pump1 (Wpum1), one is increasing the pressure difference of pump1 which leads to the increase of Wpum1 and the other is that it has a negative impact on Wpum1 by the reduction of the Mtot. And Wpum1 is the main factor affecting Cpum1. Therefore, as Pl increases, Cpum1increases firstly and then decreases. In addition, the increase of Ph has little effect on the pressure difference of pump1 and Mtot. So Wtur1 and Cpum1 basically remains unchanged as Ph increases. Fig. 17b outlines the effect of Ph and Pl on the investment cost of pump 2 (Cpum2). By raising Pl, the pressure difference of pump 2 decreases and Mh basically remains unchanged. Therefore, Cpum2 decreases as Pl increases. Same as Pl in Fig. 16a, the increase of Ph also has two opposite effects on the Wpum2: increasing the pressure difference of pump 2 and decreasing Mh. Therefore, Wpum2 and Cpum2 firstly increase and then decrease with the increase of Ph. The effect of Ph and Pl on the variation of investment cost of two pumps (Cpum) is depicted in Fig. 18a. It is obvious that Cpum firstly
728
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
160
160
128.8
140
17.1%
75.2 93.3 19.4% 37.8
100
48.0 0
50
SORC DORC
37.8
21.3%
48.1
100
Ths1
140
155.5
120
Tin=177.2 oC
180
14.6%
241.7
Temperature/oC
Heat source inlet temperature/oC
200
206.3
150 Wnet/kW
120
3
100 80 60
Tout
40
200
20
250
2 0
400
800
a) Maximum net power output of SORC and DORC
7.53 7.48
100
SORC DORC
5.53 0
2
4
6
8
Ths1
140 Temperature/oC
Heat source inlet temperature /oC
160
9.35
5.75
10
120
80 60 40 20
12
7
6
T T hs2 Tout hs3 2 0
34
5
400
800
b) Thermal efficiency of SORC and DORC
1200 1600 Q/kW
2000
2400
2800
b) DORC Fig. 14. Temperature profiles in the evaporator of SORC and DORC when Tin = 177.2 °C.
2141.4 1875.8 656.13kW 1423.1
140
1657.9
2800
998.4
120
1246.1
SORC DORC
867.7 0
400
800
1200
Ths=120oC Ths=140oC
2000
656.1
100
Ths=100oC
2400
1600
2000
Ths=160oC
1600 Ctot/k$
Heat sourece inlet temperature/oC
2800
100
Thermal efficiency/%
160
2400
Tin=177.2 oC
Isobutane
180
9.0
120
2000
200
11.29
140
1200 1600 Q/kW
a) SORC
11.0
160
7
2400
Heat absorption capacity/kW
1785.6
1200
1320
800
c) Heat absorption capacity of SORC and DORC Fig. 13. Wmax and corresponding heat recovery capacity and thermal efficiency of SORC and DORC.
939.5
400 0
increases and then decreases with the increase of Ph and Pl. As discussed above, it is the combined effect of Cpum1 and Cpum2. Fig. 18b presents the effect of Ph and Pl on the variation of the investment cost of turbine (Ctur). It is apparent that the variation tendency of Ctur is similar to that of Wtur, and the reason is also analogous.
607.5 400
800
1200
1600 2000 Pe/kPa
Fig. 15. Ctot of SORC over Pe.
729
2400
2800
3200
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
0.7 0.6
Ths=100 oC
Ths=140 oC
1500
Ths=120 oC
Ths=160 oC
1400 1300
0.5 EPC($ kWh)
Cpum/k$
1200 Ph/kPa
0.4
1100
0.3
1000
0.24 0.19
0.2
0.16
900
0.13
800
0.1
Ths=100 oC
700
600
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
900 1200 1500 1800 2100 2400 2700 3000 3300 Pe/kPa
a) Pump
Fig. 16. EPC of SORC over Pe.
Ctur/k$
1500
1500
Cpum1/k$
1400
1400 1300
48.30
1300
51.01
1200
53.72
1200
Ph/kPa
56.42
1100
59.13
1100
Ph/kPa
61.84
1000
64.55
1000
67.26
900
69.97
900
72.67
800
75.38
800
80.80
600
700
800
b) Turbine
a) Pump 1 Cpum2/k$ 38.30 41.46 44.62 47.77 50.93 54.09 57.25 60.41 63.57 66.72 69.88 73.04 76.20
Ph/kPa
1100 1000 900 800 700 700
800
1400
70.90
1300
77.05
73.98 80.13
1200
83.20
1100
86.28
1000
92.42
900
98.58
Ph/kPa
1200
600
Cexc/k$
1500
1500
1300
123.0 134.9 146.8 158.6 170.5 182.4 194.3 206.1 218.0 229.9 241.8 253.6 265.5 277.4 289.3 301.1 313.0
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
900 1000 1100 1200 1300 1400 Pl/kPa
1400
Ths=100 oC
700
78.09
700
63.60 65.18 66.76 68.34 69.92 71.51 73.09 74.67 76.25 77.83 79.41 80.99 82.58 84.16 85.74 87.32 88.90
89.35 95.50 101.7
800
Ths=100 oC
700
104.7 107.8
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
900 1000 1100 1200 1300 1400 Pl/kPa
c) Heat exchanger
b) Pump 2
Fig. 18. Investment cost of DORC component.
Fig. 17. Investment cost of pump 1 and pump 2.
The variation of the investment cost of the heat exchanger by increasing Ph and Pl is demonstrated in Fig. 18c. It can be seen that Cexc maintains within a narrow range as Ph increases, and decreases as Pl decreases. Cexc is mainly determined by the heat exchanger surface
areas. According to Eq. (9), there is a direction correlation between the heat exchanger area and absorbed heat from the evaporator. The absorbed heat is decided by Tout, which is shown in Fig. 7. Therefore, Cexe has the opposite variation trend with Tout, whereas the reason is similar.
730
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
1500
403.0 426.0 449.0 472.0 495.0 518.0 541.0 564.0 587.0 610.0 633.0 656.0 679.0 702.0 725.0 748.0 771.0
1300 1200
770 k$
Ph=1160 kPa
1100 1000 900 800
o
Ths=100 C
Pl=773kPa
700 600
700
800
Ctot(k$)
2400
1587 k$
1600
Pl=1137 kPa
800
Pl/kPa
a) Ths=100 C
c) Ths=140oC
2200
Ctot/k$
1800
Ph/kPa
Ph=1566 kPa
1400 1200 Ths=120oC Pl=929 kPa
800 600
482 524 566 608 651 693 735 777 819 861 903 945 988 1030 1072 1114 1156
Ph=3266.1kPa
3300
Ctot(k$)
2172 k$
3000 2700 2400 Ph/kPa
2000
1000
Th=140 oC
600 900 1200 1500 1800 2100 2400 2700 3000
o
1600
Ph=2130 kPa
2000
1200
900 1000 1100 1200 1300 1400 Pl/kPa
1154 k$
670.0 727.8 785.6 843.4 901.3 959.1 1017 1075 1133 1190 1248 1306 1364 1422 1479 1537 1595
2800
Ph/kPa
1400
Ph/kPa
3200
Ctot(k$)
2100 1800 1500 Ths=160 oC
1200 900
800 1000 1200 1400 1600 1800 2000 Pl/kPa
600
b) Ths=120oC
Pl=1397 kPa
985 1059 1134 1208 1283 1357 1431 1506 1580 1654 1729 1803 1878 1952 2026 2101 2175
900 1200 1500 1800 2100 2400 2700 3000 Pl/kPa
d) Ths=160oC Fig. 19. Contours of Ctot of the DORC.
Fig. 19a shows the investment cost of DORC (Ctot) variations with the increase of Ph and Pl when Ths is 100 °C. Consequence of above mentioned reasons leads to the variations trend of Ctot: Ctot firstly increases and then decreases with the increase of Ph and Pl. Fig. 19b–d present variations of Ctot with the increase of the Ph and Pl when Ths is 120 °C, 140 °C and 160 °C respectively. The variation trend of Ctot is similar, and the reason is also same. It can be seen than the optimized Ctot increases from 770 k$ to 2172 k$ as Ths increases from 100 °C to160 °C The effect of Ph and Pl on the EPC variation is shown in Fig. 20. From Fig. 20a, it can be observed that EPC decreases firstly, and then increases with the increase of Ph and Pl. The lower values of EPC can be seen in the range of 1000 kPa < Ph < 1400 kPa and 680 kPa < Pl < 1050 kPa. Consequently, the value of EPC is 0.241 at the corresponding optimized conditions of Ph = 1160 kPa and Pl = 773 kPa. Combined with Fig. 12, we could found that the range of Ph and Pl of higher Wnet is basically included by that of lower EPC. That is to say,
the DORC could show better thermodynamic and thermo-economic performances in the optimized two-stage pressures. Fig. 20b–d show the variations of EPC with the increase of Ph and Pl when Ths is 120 °C, 140 °C, and 160 °C respectively. It is obvious that EPC decreases from 0.24 $/kW h to 0.14 $/kW h when Ths increases from 100 °C to 160 °C. In a word, DORC system possesses better thermoeconomic performance at higher Ths. 4.2.3. Thermo-economic performance comparison Fig. 21a presents the optimized Ctot of SORC and DORC over different Ths. It is apparent the optimized Ctot of DORC is larger than that of the SORC. Moreover, optimized Ctot increases as Ths increases on account of the increase of the system capacity. Fig. 21b shows the optimized EPC of SORC and DORC over different Ths. It can be seen that the optimized EPC of SORC and DORC show little change. From Figs. 21b and 12, we could find that the DORC can significantly increase net power output, but the thermo-economic
731
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
EPC($/kWh)
1500 1400
0.2592 0.2773 0.2864
Ph/kPa
Ph/kPa
1160 kPa
1100
0.2955
1000
0.3046
900
0.3227
800
Ths=100 C
773 kPa
700
0.1860
2100 1800
0.3409
0.2980 0.3140
Pl=1137 kPa
0.3460
Pl/kPa
EPC($/kWh)
0.2142 0.2278
1400
0.2549 0.2685 0.2821
1200
Ph/kPa
Ph=1566 kPa
0.2413
0.2957
o
Ths=120 C
0.1320
2700
0.1368
2400
0.1465
2100
0.1562
0.1417 0.1513 0.1610
1800
0.1658 0.1707
1500
0.3092
1000
EPC($/kWh)
0.14 $/kWh
3000
0.2006
1800
Ph=1397 kPa
3300
0.1870
2000
Ph/kPa
Th=140 C
c) Ths=140oC
2200
600
0.3300
o
600 900 1200 1500 1800 2100 2400 2700 3000
a) Ths=100 C
800
0.2500 0.2820
o
Pl=929 kPa
0.2340 0.2660
900
600 700 800 900 1000 1100 1200 1300 1400 Pl/kPa
1600
0.2180
Ph=2130 kPa
1200
0.3318 0.3500
0.19$/kWh
0.2020
0.16 $/kWh
1500
0.3137
o
0.1700
2400
0.2682
0.24$/kWh
1200
0.1540
2700
0.2501
1300
EPC($/kWh)
3000
0.2410
0.1755
1200
0.3228 0.3364
Pl=1397 kPa
900
0.3500
600
800 1000 1200 1400 1600 1800 2000 Pl/kPa
0.1803
Ths=160oC
0.1852 0.1900
900 1200 1500 1800 2100 2400 2700 3000 Pl/kPa
d) Ths=160oC
b) Ths=120oC
Fig. 20. Contours of EPC of DORC system.
performance is not reduced by increased system complexity compared to the SORC. This well confirms that DORC has a good prospect for application at lower temperature (100–177.2 °C) heat source for isobutane.
21.3% to 14.6% when Ths increases from 100 °C to 160 °C. DORC system has more advantage at lower Ths. (c) Both SORC and DORC show better thermo-economic performances at higher heat source inlet temperature. Optimized EPC of SORC decreases from 0.24 $/kW h to 0.13 $/kW h, that of DORC decreases from 0.24 $/kW h to 0.14 $/kW h as Ths increases from 100 to 160 °C (d) Optimized EPC of SORC and DORC at the same heat source inlet temperature are nearly equal. The DORC can significantly increase net power output, but the thermo-economic performance is not reduced by increased system complexity compared to the SORC.
5. Conclusions The thermodynamic and thermo-economic performances of SORC and DORC using isobutane as working fluids were investigated and evaluated. The effects of the two-stage pressures and the heat source temperature on the system performances are also analyzed. Main conclusions obtained are listed as follows: (a) Thermodynamic performance of DORC using isobutane as working fluid is better than that of SORC as the heat source temperature is lower than 177.2 °C; DORC system no longer has performance advantage if the heat source temperature is higher than 177.2 °C. (b) The optimized net power output increment of DORC decreases from
Acknowledgement This work has been supported by the Natural Science Foundation of Shandong Province (Nos. ZR2014EEP026 and ZR2016AP07) and the Scientific Research Foundation of Ludong University (No. 27860301).
732
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
Heat source temperature/oC
2795.1
1785.6
160
2172.3
2605
1320
140
1586.6 939.5
120
1154.2 607.5
100
770 0
SORC DORC
770
300 600 900 1200 1500 1800 2100 2400 2700 3000 Ctot/k$
a) Ctot
Heat source temperature/oC
160
140
0.1312
DORC SORC
0.1362 0.1553 0.1875
0.1551
0.1875 0.1894
120
0.1875
0.1875 0.2438
100
0.1875
0.2431
0.00 0.03 0.06 0.09 0.12 0.15 0.18 0.21 0.24 0.27 0.30 EPC($/kWh)
b) EPC Fig. 21. Thermo-economic comparisons of SORC and DORC.
Appendix A. . Heat transfer A.1. Single phase heat transfer coefficient in the shell side The single phase heat transfer coefficient of shell side is calculated by the correlation of Bell-Dlaware [36]: (A1)
ho = h id Jc Jl Jb Js
where Jc is the correction factor for baffle cut and spacing; Jl is the correlation factor for baffle leakage; Jb is the correction factor for bundle bypassing effects; Js is the correlation factor for variable baffle spacing at the inlet and outlet; The combined effects of all these correction factors for a reasonable well-designed shell-and-tube heat exchanger is of the order of 0.60 [36]. hd is the ideal transfer coefficient for pure cross flow,
hd = ji cps
m sh Ash
k sh c psh µsh
2/3
µsh µsh,w
0.14
(A2)
where cps is the specific heat of the working fluid, kJ/(kg·K); sh stands for shell; msh is the mass flow rate of working fluid, kg/s; ksh is the thermal conductivity of the working fluid, W/(m·K); μsh is the dynamic viscosity of working fluid, Pa·s; Ash is the cross flow area at the centerline of the shell for one cross flow between two baffles, m; ji is the Colburn j-factor for an ideal tube bank. A.2. The boiling heat transfer coefficient in the shell-side The boiling heat transfer coefficient in the evaporator of the shell-side is given as follows [37]: (A3)
he = h nb Fb + h nc 2
where Fb is a correction factor for the effect of convection; hnb is the nucleate boiling coefficient, W/(m ·K); hnc is the natural convection heat transfer coefficient, W/(m2·K). The parameters are calculated as follows:
733
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al. 0.7 h nb = 0.00417p0.69 c q Fp
Fp = 0.7 + 2pr (4 +
1 ) 1 pr
(A5)
0.785Dotl 1 Ct (Pt / do )2d o
Fb = 1 + 0.1
Dotl = Dsh
(A4)
0.75
(A6) (A7)
bb
(A8)
= 0.017Ds + 0.256m
bb
2
where Pc is the working fluid critical pressure, MPa; q is the heat flux, kW/m ; pr = P/Pc; Pt is the tube pitch, m; Dsh is the diameter of the shell, m; δbb is the bypass clearance; Ct is a constant decided by the tube bundle configuration [37]. A.3. The condensation heat transfer coefficient in the shell side The condensation heat transfer coefficient for the two-phase fluid in the condenser is calculated by [38] (A9)
hs / hl = 1.26Xtt 0.78 hl = ji Gs· cps·Pr (
Xtt =
1 x x
2/3) (µ
0.25
(A10)
1 /µ w )
v
µl
l
µv
0.1
(A11) 2
where hl is the heat-transfer coefficient for the liquid phase flowing alone through the bundle, W/(m ·K); Xtt is the Lockhart–Martinelli parameter, Eq. (A11); Gs is the shell side mass velocity of the working fluid, kg/(m2·s); cps is the specific heat capacity of the working fluid, kJ/(kg·K); Pr is the Prandtl number of the working fluid; x is the vapor mass fraction in the liquid-vapor mixture; V and L stand for vapor and liquid, respectively; μ is the viscosity of the working fluids, kg/(m·s). The Colburn j-factor for an ideal tube bank is given as follows:
ji = a1
1.33 PT/ d o
a
(Resh )a2
(A12)
where Res is the Reynolds number of working fluids:
Resh =
m sh d 0 Ash µsh
(A13)
where Ash is the bundle cross flow area, m2. where
a=
a3 1 + 0.14(Resh ) a4
(A14)
Table A.1 gives the coefficients of Eqs. (A12) and (A14).
Table A1 Correlation coefficients for ji in Eqs. (A12) and (A14). Layout angle
Reynolds number
a1
a2
a3
a4
30°
4
10 -10 104-103 103-102 102-10 < 10
0.321 0.321 0.593 1.360 1.4
−0.388 −0.388 −0.477 −0.657 −0.667
1.450
0.519
45°
105-104 104-103 103-102 102-10 < 10
0.370 0.370 0.73 0.498 1.550
−0.396 −0.396 −0.500 −0.667 −0.667
1.930
0.500
90°
105-104 104-103 103-102 102-10 < 10
0.370 0.107 0.408 0.900 0.97
−0.395 −0.266 −0.460 −0.631 −0.667
1.187
0.370
5
734
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
A.4. Single phase heat transfer coefficient in the tube side The correlation of Petukhow and Popov [39] is used to calculate the single phase heat transfer coefficient and is given by:
ht =
Nu·k di
(A15)
where the Nusselt number can be calculated by
Nu =
(f /8) Ret ·Prt [1.27·(f /8)0.5·(Pr 2/3 1) + 1.07]
(A16)
And
Ret =
Prt =
t vt d i
(A17)
µt
cpt µt
(A18)
kt
f = (0.782· lgRet 1.51)
(A19)
2
where t stands for tube; Ret and Prt are the Reynolds number and the Prandtl number of the heat source water, respectively; ρt, vt, kt, μt are the density, mass velocity, thermal conductivity and viscosity of the heat source water respectively. Appendix B. . Total investment cost The bare module cost for each component is calculated by (B1)
Cbm = Cp Fbm
where Cp is the purchased cost for base conditions: equipment made of carbon steel and operating at near ambient pressures, which is calculated by (B2)
log10 Cp = K1 + K2log10 (X ) + K3 (log10 X ) 2
where X refer to the output net power for the turbine, input net power for the pump, or surface area for heat exchanger respectively. The values for K1, K2 and K3 are given in the Table B1. Fbm is the factor of bare module cost, which is given as follows: (B3)
Fbm = B1 + B2 FM FP where the values of B1and B2 are given in Table B1; Fp is the pressure factor, which is calculated by the following equation:
(B4)
log10 Fp = C1 + C2 log10 P + C3 (log10 P ) 2 where P is the operating pressure and the units of P is bar gauge. The values of C1, C2 and C3 are given in Table B1. Table B1 Component bare module cost parameters [41]. Components
K1
K2
K3
C1
C2
C3
B1
B2
FM
FBM
Pump Turbine Condenser Evaporator
3.87 2.705 4.325
0.316 1.44 −0.303
0.122 −0.177 0.163
−0.245 / −0.039
0.259 / 0.082
−0.014 / −0.012
1.89 / 1.63
1.35 / 1.66
2.35 / 1.35
/ 6.2 /
cycle. Appl Energy 2014;119:204–17. [7] Chagnon-Lessard N, Mathieu-Potvin F, Gosselin L. Geothermal power plants with maximized specific power output: optimal working fluid and operation conditions of subcritical and transcritical Organic Rankine Cycles. Geothermics 2016;64:111–24. [8] Lecompte S, Huisseune H, Broek MVD, Vanslambrouck B, Paepe MD. Review of organic Rankine cycle (ORC) architectures for waste heat recovery. Renew Sust Energy Rev 2015;47:448–61. [9] Braimakis K, Karellas S. Energetic optimization of regenerative Organic Rankine Cycle (ORC) configurations. Energy Convers Manage 2018;159:353–70. [10] Maraver D, Royo J, Lemort V, Quoilin S. Systematic optimization of subcritical and transcritical organic Rankine cycles (ORCs) constrained by technical parameters in multiple applications. Appl Energy 2014;117:11–29. [11] Li J, Liu Q, Duan YY, Yang Z. Performance analysis of organic Rankine cycles using R600/R601a mixtures with liquid-separated condensation. Appl Energy 2017;190:376–89. [12] Sadeghi M, Nemati A, Ghavimi A, Yari M. Thermodynamic analysis and multi objective optimization of various ORC (organic Rankine cycle) configurations using zeotropic mixtures. Energy 2016;109:791–802. [13] Heberle F, Bruggemann D. Thermo-economic evaluation of organic Rankine
References [1] Tchanche BF, Lambrinos G, Frangoudakis A, Papadakis G. Low-grade heat conversion into power using organic Rankine cycles – a review of various applications. Renew Sust Energy Rev 2011;15(8):3963–79. [2] Wang MT, Chen YG, Liu QY. Experimental study on the gas engine speed control and heating performance of a gas engine-driven heat pump. Energy Build 2018;178:84–93. [3] Yang MH, Yeh RH. Economic performances optimization of the transcritical Rankine cycle systems in geothermal application. Energy Convers Manage 2015;95:20–31. [4] Shi LF, Shu GQ, Tian H, Deng S. A review of modified Organic Rankine cycles (ORCs) for internal combustion engine waste heat recovery (ICE-WHR). Renew Sust Energy Rev 2018;92:95–110. [5] Li J, Liu Q, Ge Z, Duan YY, Yang Z. Thermodynamic performance analyses and optimization of subcritical and transcritical organic Rankine cycles using R1234ze (E) for 100–200°C heat sources. Energy Convers Manage 2017;149:140–54. [6] Shu GQ, Li XN, Tian H, Liang XY, Wei HQ, Wang X. Alkanes as working fluids for high-temperature exhaust heat recovery of diesel engine using organic Rankine
735
Energy Conversion and Management 177 (2018) 718–736
M. Wang et al.
[14] [15] [16]
[17] [18] [19] [20] [21] [22] [23] [24] [25] [26]
cyclesfor geothermal power generation using zeotropic mixtures. Energies 2015;8(3):2097–124. Liu Q, Duan YY, Yang Z. Effect of condensation temperature glide on the performance of organic Rankine cycles with zeotropic mixture working fluids. Appl Energy 2014;115:394–404. Tian H, Chang LW, Gao YY, Shu GQ, et al. Thermo-economic analysis of zeotropic mixtures based on siloxanes for engine waste heat recovery using a dual-loop organic Rankine cycle (DORC). Energy Convers Manage 2017;136:11–26. Chagnon-Lessard N, Mathieu-Potvin F, Gosselin L. Geothermal power plants with maximized specific power output: optimal working fluid and operating conditions of subcritical and transcritical Organic Rankine Cycles. Geothermics 2016;64:111–24. Zhang SJ, Xu XX, Liu C, Zhang YD, Dang CB. The buoyancy force and flow acceleration effects of supercritical CO2 on the turbulent heat transfer characteristics in heated vertical helically coiled tube. Heat Mass Transf 2018;125:274–89. Baik Y, Kim M, Chang KC, Kim SJ. Power-based performance comparison between carbon dioxide and R125 transcritical cycles for a low-grade heat source. Appl Energy 2011;88:892–8. Li J, Liu Q, Ge Z, Duan YY, Yang Z. Thermodynamic performance analyses and optimization of subcritical and transcritical organic Rankine cycles using R1234ze (E) for 100–200 °C heat sources. Energy Convers Manage 2017;149:140–54. Shu G, Wang X, Tian H. Theoretical analysis and comparison of rankine cycle and different organic rankine cycles as waste heat recovery system for a large gaseous fuel internal combustion engine. Appl Therm Eng 2016;108:525–37. Yang FB, Cho H, Zhang HG, Zhang J. Thermoeconomic multi-objective optimization of a dual loop organic Rankine cycle (ORC) for CNG engine waste heat recovery. Appl Energy 2017;205:1100–18. Huang H, Zhu J, Yan B. Comparison of the performance of two different Dual loop organic Rankine cycles (DORC) with nanofluid for engine waste heat recovery. Energ Convers Manage 2016;126:99–109. Franco A, Villani M. Optimal design of binary cycle power plants for water dominated, medium-temperature geothermal fields. Geothermics 2009;38:379–91. Peris B, Navarro-Esbrí J, Moles F. Bottoming organic Rankine cycle configurations to increase Internal Combustion Engines power output from cooling water waste heat recovery. Appl Therm Eng 2013;61:364–71. Stijepovic MZ, Papadopoulos AI, Linke P, Grujic AS, Seferlis P. An exergy composite curves approach for the design of optimum multi-pressure organic Rankine cycle processes. Energy 2014;69:285–98. Guzovic Z, Raskovic P, Blataric Z. The comparision of a basic and a dual pressure ORC (organic rankine cycle): geothermal power plant Velika Ciglena case study. Energy 2014;76:175–86.
[27] Li T, Yuan Z, Li W, Yang J, Zhu J. Strengthening mechanisms of two-stage evaporation strategy on system performance for organic Rankine cycle. Energy 2016;101:532–40. [28] Thierry DM, Flores-Tlacuahuac A, Grossmann IE. Simultaneous optimal design of multi-stage organic Rankine cycles and working fluid mixtures for low temperature heat sources. Comput Chem Eng 2016;89:106–26. [29] Li TL, Zhang ZG, Lu J, Yang JN, Hu YJ. Two-stage evaporation strategy to improvesystem performance for organic Rankine cycle. Appl Energy 2015;150:323–34. [30] Shokati N, Ranjbar F, Yari M. Exergoeconomic analysis and optimization of basic, dual-pressure and dual–fluid ORCs and Kalina geothermal power plants: a comparative study. Renew Energy 2015;83:527–42. [31] Manente G, Lazzaretto A, Bonamico E. Design guidelines for the choice between single and dual pressure layouts in organic Rankine cycle (ORC) systems. Energy 2017;123:413–31. [32] Sadeghi M, Nemati A, Ghavimi A, Yari M. Thermodynamic analysis and multi–objective optimization of various ORC (organic Rankine cycle) configurations using zeotropic mixtures. Energy 2016;109:791–802. [33] Li J, Ge Z, Duan YY, Yang Z, Liu Q. Parametric optimization and thermodynamic performance comparison of single-pressure and dual-pressure evaporation organic Rankine cycles. Appl Energy 2018;217:409–21. [34] Li J, Ge Z, Liu Q, Yang Z, Duan YY. Thermo-economic performance analyses and comparison of two turbine layouts for organic Rankine cycles with dual-pressure evaporation. Energy Convers Manage 2018;164:603–14. [35] Seider WD, Seader JD, Lewin DR. Product and process design principles: synthesis, analysis, and evaluation. 2nd ed. New York: John Wiley and Sons; 2004. [36] Kakac S, Liu H. Heat exchanger selecting, rating, and thermal design. 3rd ed. New York: CRC Press; 2002. [37] Hewitt GF. Hemisphere hand book of heat exchanger design. New York: Hemisphere Publishing Corporation; 1990. [38] Serth RW. Process heat transfer principles and applications. Elsevier; 2007. [39] Irvine TF, Hartnett JP, editors. Advances in heat transfer, vol. 6. New York: Academic Press; 1970. [40] Turton R, Bailie RC, Whiting WB, Shaeiwitz JA. Analysis, synthesis and design of chemical processes. Pearson Education; 2008. [41] Shu GQ, Yu GP, Tian H, Wei HQ, Liang XY. A multi-approach evaluation system (MA-ES) of organic Rankine cycles (ORC) used in waste heat utilization. Appl Energy 2014;132:325–38. [42] Kolahi M, Yari M, Mahmoudi SMS, Mohammadkhani F. Thermodynamic and economic performance improvement of ORCs through using zeotropic mixtures: case of waste heat recovery in an offshore platform. Case Stud Therm Eng 2016;8:51–70.
736